Appreciating Physical Landscapes: Three Hundred Years of Geotourism 1862397244, 9781862397248


329 117 124MB

English Pages 256 [245] Year 2016

Report DMCA / Copyright

DOWNLOAD PDF FILE

Table of contents :
0a front-matter
0b Table of Contents
0c About this title
1 Three centuries (1670–1970) of appreciating physical landscapes
2 Appreciating geology and the physical landscape in Scotland
3 Waterfalls and the Romantic traveller
4 The artist as geotourist- Eugene von Guérard
5 Landscape and geotourism on the Dutch coast
6 Visitors to ‘the northern playgrounds’- Norway
7 The role of local societies in early modern geotourism
8 Geotourism-an early photographic insight
9 The contribution of maps to appreciating physical landscape
10 Three centuries of open access to the caves in Stoney Middleton Dale Site
11 Geology and landscape in SW England in the late eighteenth century
12 The role of Carclaze tin mine in eighteenth and nineteenth century geotourism
13 From tourism to geotourism - French Alpine
14 Rediscovering geoheritage, reinventing geotourism
15 Appreciating loess landscapes through history
16 back-matter
Recommend Papers

Appreciating Physical Landscapes: Three Hundred Years of Geotourism
 1862397244, 9781862397248

  • 0 0 0
  • Like this paper and download? You can publish your own PDF file online for free in a few minutes! Sign Up
File loading please wait...
Citation preview

Appreciating Physical Landscapes: Three Hundred Years of Geotourism

The Geological Society of London Books Editorial Committee Chief Editor

Rick Law (USA) Society Books Editors

Jim Griffiths (UK) Dave Hodgson (UK) Phil Leat (UK) Nick Richardson (UK) Daniela Schmidt (UK) Randell Stephenson (UK) Rob Strachan (UK) Mark Whiteman (UK) Society Books Advisors

Ghulam Bhat (India) Marie-Franc¸oise Brunet (France) Martin Hand (Australia) Jasper Knight (South Africa) Mario Parise (Italy) Satish-Kumar (Japan) Marco Vecoli (Saudi Arabia) Gonzalo Veiga (Argentina)

Geological Society books refereeing procedures The Society makes every effort to ensure that the scientific and production quality of its books matches that of its journals. Since 1997, all book proposals have been refereed by specialist reviewers as well as by the Society’s Books Editorial Committee. If the referees identify weaknesses in the proposal, these must be addressed before the proposal is accepted. Once the book is accepted, the Society Book Editors ensure that the volume editors follow strict guidelines on refereeing and quality control. We insist that individual papers can only be accepted after satisfactory review by two independent referees. The questions on the review forms are similar to those for Journal of the Geological Society. The referees’ forms and comments must be available to the Society’s Book Editors on request. Although many of the books result from meetings, the editors are expected to commission papers that were not presented at the meeting to ensure that the book provides a balanced coverage of the subject. Being accepted for presentation at the meeting does not guarantee inclusion in the book. More information about submitting a proposal and producing a book for the Society can be found on its website: www.geolsoc.org.uk.

It is recommended that reference to all or part of this book should be made in one of the following ways: Hose, T. A. (ed.) 2016. Appreciating Physical Landscapes: Three Hundred Years of Geotourism. Geological Society, London, Special Publications, 417. Cayla, N., Gauchon, C. & Hoble´a, F. 2016. From tourism to geotourism: a few historical cases from the French Alpine foreland. In: Hose, T. A. (ed.) Appreciating Physical Landscapes: Three Hundred Years of Geotourism. Geological Society, London, Special Publications, 417, 199–213. First published online February 11, 2015, http://dx.doi.org/10.1144/SP417.10

GEOLOGICAL SOCIETY SPECIAL PUBLICATION NO. 417

Appreciating Physical Landscapes: Three Hundred Years of Geotourism

EDITED BY

T. A. HOSE University of Bristol, UK

2016 Published by The Geological Society London

THE GEOLOGICAL SOCIETY The Geological Society of London (GSL) was founded in 1807. It is the oldest national geological society in the world and the largest in Europe. It was incorporated under Royal Charter in 1825 and is Registered Charity 210161. The Society is the UK national learned and professional society for geology with a worldwide Fellowship (FGS) of over 10 000. The Society has the power to confer Chartered status on suitably qualified Fellows, and about 2000 of the Fellowship carry the title (CGeol). Chartered Geologists may also obtain the equivalent European title, European Geologist (EurGeol). One fifth of the Society’s fellowship resides outside the UK. To find out more about the Society, log on to www.geolsoc.org.uk. The Geological Society Publishing House (Bath, UK) produces the Society’s international journals and books, and acts as European distributor for selected publications of the American Association of Petroleum Geologists (AAPG), the Indonesian Petroleum Association (IPA), the Geological Society of America (GSA), the Society for Sedimentary Geology (SEPM) and the Geologists’ Association (GA). Joint marketing agreements ensure that GSL Fellows may purchase these societies’ publications at a discount. The Society’s online bookshop (accessible from www.geolsoc. org.uk) offers secure book purchasing with your credit or debit card. To find out about joining the Society and benefiting from substantial discounts on publications of GSL and other societies worldwide, consult www.geolsoc.org.uk, or contact the Fellowship Department at: The Geological Society, Burlington House, Piccadilly, London W1J 0BG: Tel. +44 (0)20 7434 9944; Fax +44 (0)20 7439 8975; E-mail: [email protected]. For information about the Society’s meetings, consult Events on www.geolsoc.org.uk. To find out more about the Society’s Corporate Affiliates Scheme, write to [email protected]. Published by The Geological Society from: The Geological Society Publishing House, Unit 7, Brassmill Enterprise Centre, Brassmill Lane, Bath BA1 3JN, UK The Lyell Collection: www.lyellcollection.org Online bookshop: www.geolsoc.org.uk/bookshop Orders: Tel. +44 (0)1225 445046, Fax +44 (0)1225 442836 The publishers make no representation, express or implied, with regard to the accuracy of the information contained in this book and cannot accept any legal responsibility for any errors or omissions that may be made. # The Geological Society of London 2016. No reproduction, copy or transmission of all or part of this publication may be made without the prior written permission of the publisher. In the UK, users may clear copying permissions and make payment to The Copyright Licensing Agency Ltd, Saffron House, 6– 10 Kirby Street, London EC1N 8TS UK, and in the USA to the Copyright Clearance Center, 222 Rosewood Drive, Danvers, MA 01923, USA. Other countries may have a local reproduction rights agency for such payments. Full information on the Society’s permissions policy can be found at: www.geolsoc.org.uk/permissions British Library Cataloguing in Publication Dataw A catalogue record for this book is available from the British Library. ISBN 978-1-86239-724-8 ISSN 0305-8719 Distributors For details of international agents and distributors see: www.geolsoc.org.uk/agentsdistributors Typeset by Techset Composition India (P) Ltd, Bangalore and Chennai, India Printed and bound by CPI Group (UK) Ltd, Croydon CR0 4YY

Foreword The appreciation of landscape physically has been an applied human attribute since the beginning of civilization in regard to agriculture and irrigation, the location of building materials and metals and the selection of sites for habitation and the routes between them. This is implied by archaeology and artefacts. Recording the physical landscape in guides and on maps has a more recent history preceding the modern scientific approach and then enabling a popular access to understanding the new and developing discipline of geology and appreciating what underlies and forms the physical landscape. Increasing literacy, travel and available time extended the opportunities and aspirations for the wider public, beyond the small world of the professional and academic geologist, to visit landscapes by armchair, road, and later railway as ‘geotourists’. Geotourism is a modern term, coined and defined by Dr. Tom Hose, for a relatively recent phenomenon – travel to enjoy and understand scenery.

The conference, Appreciating Physical Landscapes: Geotourism 1670–1970, was convened and organized by Tom at the Geological Society apartments on 23rd October 2012. It assembled an international cast of some 15 speakers from Australia, South Africa, North America and Europe; their presentations were supplemented by some nine poster presentations. Examples of geotourism and the dissemination of particular landscape descriptions through art, literature and local natural history societies were among the papers and posters presented, a selection of which appear in this volume. The contribution of maps and guidebooks is presented in a commissioned paper. The conference was followed by a convivial dinner most ably organized by Dick Moody. The following day the conference field excursion, with an accompanying guidebook, arranged and led by Tom went to Margate on the Kent coast. Margate was one of the earliest resorts for health and appreciation of scenery, so that we travelled sustainably, by train

Conference geotourists braving the rain in front of the modernised entrance to Margate’s shell grotto, an eighteenth century decorated ‘cave’ that still attracts the idle and curious.

viii

FOREWORD

and bus, and literally in the footsteps of nineteenth century geotourists. Margate was fittingly partly second home to the great English Romantic landscape artist J. M. W. Turner and appropriately the excursion took afternoon tea at the Turner Contemporary gallery. In typical field excursion practice, the anticipated opening vista of The Bay from the railway station, depicted in our guidebook’s nineteenth century engraving of Margate, was completely shrouded by fog on our arrival; this gradually dispersed and in the late afternoon sunlight, something much appreciated by Turner, The Bay was at last revealed to the excursionists. It is the appreciation of landscapes in all weathers that marks out the true geotourist! As chair of the History of Geology Group, I would like to thank Naomi Newbold and Georgina Worrall of the Geological Society’s conference

office for their vital assistance in organizing the conference. We are also grateful to the staff of the Geological Society Library for their excellent display of conference related material in the Lower Library. The generous support of sponsors Arup, the British Society for the History of Science and Rockhounds Welcome! is very much appreciated. The conference is deeply indebted to our keynote speakers, John Gordon and David Norman, the session speakers and chairs, and also the presenters of the posters. Finally, the conference and this publication would not have come about without the vision, connections, organization and perseverance of Tom Hose, for which HOGG sincerely thanks him. C. John Henry July 2015

Contents HENRY, C. J. Foreword HOSE, T. A. Three centuries (1670–1970) of appreciating physical landscapes

vii 1

GORDON, J. E. & BAKER, M. Appreciating geology and the physical landscape in Scotland: from tourism of awe to experiential re-engagement

25

HUDSON, B. J. Waterfalls and the Romantic traveller

41

PULLIN, R. The artist as geotourist: Eugene von Gue´rard and the seminal sites of early volcanic research in Europe and Australia

59

VAN DEN

ANCKER, J. A. M. & JUNGERIUS, P. D. Landscape and geotourism on the Dutch coast in the seventeenth century as depicted by landscape artists

71

WHALLEY, W. B. & PARKINSON, A. F. Visitors to ‘the northern playgrounds’: tourists and exploratory science in north Norway

83

BUREK, C. V. & HOSE, T. A. The role of local societies in early modern geotourism: a case study of the Chester Society of Natural Science and the Woolhope Naturalists’ Field Club

95

LARWOOD, J. G. Geotourism: an early photographic insight through the lens of the Geologists’ Association

117

HENRY, C. J. & HOSE, T. A. The contribution of maps to appreciating physical landscape: examples from Derbyshire’s Peak District

131

COPE, M. A. Three centuries of open access to the caves in Stoney Middleton Dale Site of Special Scientific Interest, Derbyshire

157

MATHER, J. D. Geology and landscape in SW England in the late eighteenth century, as recorded in the travel journals of William George Maton (1774– 1840)

171

BRISTOW, C. The role of Carclaze tin mine in eighteenth and nineteenth century geotourism

187

CAYLA, N., GAUCHON, C. & HOBLE´A, F. From tourism to geotourism: a few historical cases from the French Alpine foreland

199

MIGON´, P. Rediscovering geoheritage, reinventing geotourism: 200 years of experience from the Sudetes, Central Europe

215

VASILJEVIC´, D. A., MARKOVIC´, S. B. & VUJICˇIC´, M. D. Appreciating loess landscapes through history: the basis of modern loess geotourism in the Vojvodina region of North Serbia

229

Index

241

Geotourism, as a form of sustainable geoheritage tourism, was defined and developed, from the early 1990s, to contextualize modern approaches to geoconservation and physical landscape management. However, its roots lie in the late seventeenth century and the emergence of the Grand Tour and its domestic equivalents in the eighteenth century. Its participants and numerous later travellers and tourists, including geologists and artists, purposefully explored wild landscapes as‘geotourists’. The written and visual records of their observations underpin the majority of papers within this volume; these papers explore some significant geo-historical themes, organizations, individuals and locations across three centuries, opening with seventeenth century elite travellers and closing with modern landscape tourists. Other papers examine the resources available to those geotourists and explore the geotourism paradigm. The volume will be of particular interest to Earth scientists, historians of science, tourism specialists and general readers with an interest in landscape history.

Three centuries (1670 –1970) of appreciating physical landscapes THOMAS A. HOSE School of Earth Sciences, University of Bristol, Wills Memorial Building, Queens Road, Clifton, Bristol, BS8 1RJ, UK (e-mail: [email protected]) Abstract: Although modern geotourism, as a form of sustainable geoheritage tourism, was only recognized as such in the 1990s, its roots lie in the seventeenth century and the Grand Tour with its domestic equivalents. At that time, a few elite travellers recorded their experiences of landscapes, natural wonders, quarries and mines. Such travellers’ observations were supplemented by those of the antiquarians for much of the eighteenth century; at that century’s close, the first modern geologists were recording their observations. The nineteenth century witnessed an explosion in public interest and engagement with geology, and field excursions were provided by the burgeoning natural history and geology societies. By the close of the nineteenth century, the Romantic movement had successfully promoted wild landscapes to a newly expanding urban population. The development of the Grand Tour and the landscape aesthetic movements, the various influential institutions, key personalities and locations are considered insofar as they provide an overview of the background to historical geotourism. All are underpinned by a theoretical consideration of the geotourism paradigm and how geotourism historical studies can contextualize modern geotourism.

Many of those who go into this field are fundamentally romantics – about being close to nature in their field work, or about pondering the great events that took place millions of years ago and the evolution of the scenery that we are now privileged to walk in. Interest in most of us is quickened by the very idea of the vast panorama of organic evolution: the feeling of being witness to a brief glimpse of a tremendous story encompassing some 5000 million years of earth history, the feeling of being close to forces powerful enough to thrust up mountain ranges and pull apart the ocean basins. (Baird 1968, p. 223)

Many Earth scientists journey and spend time ‘in the field’ for the purposes of their employment and, if they are fortunate, sometimes in places that have some grandeur. Even their holidays might partly be spent ‘in the field’ in pursuit of their interests. However, few will consider how their holiday activities align with patterns of leisure travel, to common locations, established some considerable time ago. Most will know little of the development of their disciplines, although they will be familiar with some locality names in the, especially stratigraphical, nomenclature. No more than a handful will make any connection between, say, geology and tourism and being ‘in the field’. Again, only a handful will recognize the nature and value of the data on Earth science localities and phenomena within the accounts of travellers and tourists intent on recording their impressions (in print and image), to inform others about what they had seen and done, rather than empirical observation. Some Earth scientists will have heard about ‘geotourism’, even if they do not understand its historical and cultural

implications; they will, if coming to it for the first time, deduce it has ‘something’ to do with tourism and either geology or geography. It was with that knowledge and understanding gap and the general lack of historical geotourism literature in mind that the Appreciating Physical Landscapes: Geotourism 1670– 1970 conference was organized by the History of Geology Group, or HOGG. Its presentations and posters, although not all are represented herein, form the basis of this volume, which is also supplemented by commissioned papers. The conference particularly sought material that moved beyond mere description of past events to provide critical analysis and contextualization of modern geotourism provision. Inevitably, the material has a western European bias (Fig. 1), but the places considered and the approaches adopted by various authors have a wider interest and application. Given the eclectic mix of papers and locations, coupled with the likelihood that many of the volume’s readers will posses limited knowledge and understanding of the development of tourism in aesthetically attractive landscapes, this introductory paper seeks to provide that essential background; its time frame reflects major political, social, cultural and scientific events in Britain and Europe (Fig. 2) and the countries over which they had influence elsewhere in the world. Although there has been increasing interest in tourism as a practical and an academic discipline since its emergence from geography in the mid1970s (Hall & Page 2008), the literature on the historical study of tourism is surprisingly sparse (Towner 1984, p. 215) and naturally more so for

From: Hose, T. A. (ed.) 2016. Appreciating Physical Landscapes: Three Hundred Years of Geotourism. Geological Society, London, Special Publications, 417, 1–22. First published online September 11, 2015, http://dx.doi.org/10.1144/SP417.15 # 2016 The Author(s). Published by The Geological Society of London. All rights reserved. For permissions: http://www.geolsoc.org.uk/permissions. Publishing disclaimer: www.geolsoc.org.uk/pub_ethics

2

T. A. HOSE

Fig. 1. Map of Europe and the regions covered in this volume’s papers. This map shows the main areas (excepting northern Norway and Australia) encompassed by the volume’s papers.

geotourism. However, a recent volume (Wyse Jackson 2007) has explored the historical travel component of geological inquiries. This sparseness is despite widely available textual and visual sources that, once their limitations are accepted, can be used to reconstruct key phases of past landscape tourism provision and experiences. It was particularly to promote the potential of such sources that Appreciating Physical Landscapes: Geotourism 1670– 1970 was organized; the introduction to its abstracts volume (Hose 2012a) noted that ‘Travellers, tourists and scientists practised activities nowadays subsumed within the modern geotourism paradigm long before geotourism itself was formally recognised’, and the papers in this volume underscore this truth.

Defining geotourism: a new geological paradigm The long engagement of some countries in mainland Europe and the British Isles in geological study has resulted in many of their geosites and geomorphosites, rocks, minerals and fossil (or geodiversity)

prominently featuring in modern Earth sciences literature – particularly as type material and global stratotypes. Their museums, libraries, archives and universities house the legacy of much of this collected and published geological study. Today, this material is sometimes poorly regarded and even considered a costly liability with no practical use by the stakeholders of the institutions in which it is housed. It was partly to address such issues that geotourism was developed. Geotourism was first employed as a discrete term in the mid-1990s for ‘The provision of interpretive and service facilities to enable tourists to acquire knowledge and understanding of the geology and geomorphology of a site (including its contribution to the development of the Earth sciences) beyond the level of mere aesthetic appreciation’ (Hose 1995a, p. 17). Hence, it is expected that its participants (or ‘geotourists’) have some interest, however limited, in understanding what they have seen. Its initial recognition and definition followed studies (Hose 1995a) on some geosites with interpretation funded by English Nature (the precursor of Natural England). Whilst several European geologists had fleetingly mentioned tourism and geology (De Bastion 1994;

THREE CENTURIES OF APPRECIATING PHYSICAL LANDSCAPES

3

Fig. 2. A European geotourism timeline. This summary shows the major events and influences, with some key British publications, on geotourism’s development from around 1670–1970.

Martini 1994; Spiteri 1994; Page 1998), they had neither defined their understanding of geotourism nor discussed its participants (that is, geotourists). Hence, the first published definition, with some of its associated concepts, was that which was fittingly cited in the Geoparks Programme Feasibility Study (Patzak & Eder 1998; UNESCO 2000); the study also included the essential elements of the later redefinition to ‘The provision of interpretative facilities and services to promote the value and societal benefit of geologic and geomorphologic sites and their materials, and ensure their conservation, for the use of students, tourists and other recreationalists’ (Hose 2000, p. 136). Thus, at the outset, geotourism encompassed geosite interpretative and promotional provision, together with the artefacts, places and memorials of their associated Earth scientists; these were intended to underpin the conservation of geosites and geomorphosites, as

well as geological collections and archives. This approach was incorporated, demonstrably building upon the previously widely accepted and cited definitions (Hose 1995b, 2000), including recent landscape studies (Hose 2008, 2010a, 2010b), in the updated definition of geotourism as ‘The provision of interpretative and service facilities for geosites and geomorphosites and their encompassing topography, together with their associated In situ and Ex situ artefacts, to constituency-build for their conservation by generating appreciation, learning and research by and for current and future generations’ (Hose 2012b, p. 11); it employs an easily and globally accurately translatable vocabulary for the nature, focus and location of modern geologybased geotourism with a geoconservation purpose; see Wilson (1994) for a summary and Burek & Prosser (2008) for a history of geoconservation. This approach (Hose 2012b) and its development

4

T. A. HOSE

have already been examined (Hose 2011) and need not be covered in detail herein, but it is re-presented as the ‘4Gs model of geotourism’ (Fig. 3). Other European authors have employed broader geotourism definitions. For English Nature’s geologists, it was ‘travelling in order to experience, learn from and enjoy our Earth heritage’ (Larwood & Prosser 1998, p. 98). Frey and her colleagues, from experience in Germany’s Vulkaneifel Geopark (where geoscientific, economic and political considerations combined to develop a somewhat commercially orientated approach) suggested, ‘Geotourism means interdisciplinary cooperation within an economic, success-orientated and fast moving discipline that speaks its own language. Geotourism is a new occupational and business sector. The main tasks of geotourism are the transfer and communication of geoscientific knowledge and ideas to the general public’ (Frey et al. 2008, pp. 97 –98). The introductory paper of the inaugural issue of Poland’s Geoturystyka defined geotourism as an ‘offshoot of cognitive tourism and/or adventure tourism based upon visits to geological objects (geosites) and recognition of geological processes integrated with aesthetic experiences gained by the contact with a geosite’ (Slomka & Kicinska-Swiderska 2004, p. 6). As Gray noted, ‘The fundamental principle of tourism is that places are different and provide different experiences and changes of environment. It follows, therefore, that geotourism must be based

on geodiversity, i.e. geotourism provides the opportunity to experience different geologies, geological environments and landscapes and/or take part in geological activities’ (Gray 2008, p. 295). However, his definition of geotourism as ‘Tourism based on an area’s geological or geomorphological resources that attempts to minimise the impacts of this tourism through geoconservation management’ (Gray 2008, p. 295) seemingly excludes secondary geosites (Hose 2003; Hose & Vasiljevic´ 2012) typically displaying material collected from many sources. In Australia, Dowling & Newsome (2008) produced the first text entitled Geotourism and clearly promoted geotourism’s geological approach. Their later text, Geotourism: The Tourism of Geology and Landscape (Newsome & Dowling 2010), noted ‘Geotourism is a form of natural area tourism that specifically focuses on geology and landscape. It promotes tourism to geosites and the conservation of geo-diversity and an understanding of earth sciences through appreciation and learning. This is achieved through independent visits to geological features, use of geo-trails and view points, guided tours, geoactivities and patronage of geosite visitor centres’, p. 232. Dowling (2013) promoted geotourism as an emerging form of sustainable tourism but emphasized its geological basis. In the USA, a more geographical approach to geotourism has developed; this is due to National Geographic ignoring the significant volume of published European

Fig. 3. The four Gs of geotourism model. In this visualization of geotourism, the locations and areas of the individual elements, together with their linking pointers, indicate their interrelationships and relative significances. It is a development of that published as the three Gs in Hose (2012b) but now includes geoconservation, geohistory, geo-interpretation and geosites/geomorphosites (or scenery).

THREE CENTURIES OF APPRECIATING PHYSICAL LANDSCAPES

geotourism work and then erroneously claiming to have singularly coined the term as a ‘destination’s geographic character – the entire combination of natural and human attributes that make one place distinct from another’ (Stueve et al. 2002, p. 1); this approach is essentially sustainable tourism with a holistic approach to aesthetic landscapes. In Europe, some confusion ensued when some members of the European Geoparks Network issued, without consulting any other interested parties, the Arouca Declaration in 2011; this accepted the National Geographic approach, with a geoheritage emphasis, when its organizing committee indicated ‘that there is a need to clarify the concept of geotourism. We therefore believe that geotourism should be defined as tourism which sustains and enhances the identity of a territory, taking into consideration its geology, environment, culture, aesthetics, heritage and the well-being of its residents. Geological tourism is one of the multiple components of geotourism.’ The committee, probably unwittingly, had actually embraced ‘ecotourism’. Ecotourism is defined by the World Tourism Organization (WTO) as ‘tourism which leads to management of all resources in such a way that economic, social and aesthetic needs can be fulfilled while maintaining cultural integrity, essential ecological processes, biological diversity and life support systems’ (WTO 1997). It was a somewhat disputed break with the purely geological approach adopted by the majority of Europe’s governmental agencies, non-governmental organizations (NGOs) and authorities. Such fragmentation of the established consensus on geotourism is at best unhelpful to its stakeholders, and at worst divisive; it is also confusing to governments and funding bodies. Many geotourism practitioners and researchers, whilst accepting its geological focus, appreciate the benefits of cooperating with other heritage and nature conservation parties; geotourism’s encompassing

5

of aesthetic landscapes, or scenery (Hose 2010b), would seem to make this inevitable. Geotourism, however and by whomsoever it is defined (see Fig. 4), is a form of ‘niche’ (Hose 2005; Novelli 2005) or ‘special interest’ (Weiler & Hall 1992) tourism in which the ‘traveller’s motivation and decision-making are primarily determined by a particular special interest . . . [that] implies “active” or “experiential” travel’ (Hall & Weiler 1992, p. 5). There is an obvious link with special interest travel (such as was required for the Grand Tour) which is for ‘people who are going somewhere because they have a particular interest that can be pursued in a particular region or at a particular destination’ (Read 1980, p. 195). As an element of niche tourism it offers participants a ‘meaningful set of experiences in the knowledge that their needs and wants are being met’ (Novelli 2005, p. 1). Modern geotourism provision meets geotourists’ needs by attracting them to particular localities with spectacular or readily appreciated, and usually (on-site and/or off-site) interpreted, geological and/or geomorphological features. At the participant level, it is ‘recreational geology’ that, unlike many forms of countryside (and for that matter, urban) recreation, is not limited by the seasons (Hose 1996, p. 211). It is relatively easy to define and categorize geotourists on their level of engagement, together with what and where they undertake an expressed (that, is observable and recordable) geotouristic activity. Simplistically, two geotourist groups can be recognized, ‘casual’ and ‘dedicated’ (Hose 2000). The former occasionally visit geosites mainly for recreation, pleasure and some limited intellectual stimulation; provision for them in the form of populist guides, trails and visitor centres is relatively recent. The latter intentionally visit geosites for the purpose of personal educational or intellectual improvement and enjoyment; provision for them in the

Fig. 4. Table of the summarized content of some geotourism definitions and their associated discussions. The summary, because it is necessarily based upon an interpretation of the associated discussions, is a subjective evaluation. It was developed by examining the definitions and any supporting or explanatory texts. The sources for the definitions can be found in the references. The shaded definition (Hose 2012b) is that which has been adopted for this volume. An extended summary table of definitions can be found in Hose (2012b, table 1).

6

T. A. HOSE

form of field guides and journal papers is longstanding. Geotourists can also be split into ‘recreational’ and ‘educational’ geotourists (Hose 1997). A typology of field excursionists, based upon their expressed behaviour, has been published (Hose 2006) and in modified form is re-presented herein (Fig. 5); it suggests how, with further refinement, historic geotourists might be categorized. In terms of location, we can split them into ‘rural’ and ‘urban’ geotourists and then refine the localities they visit on their physiographical characteristics. The locations visited by the various categories can overlap, although their usages and understandings are often very different, and can be chiefly split into ‘primary’ and ‘secondary’ geosites (Hose 2003; Hose & Vasiljevic´ 2012). Primary geosites have geological and/or geomorphological features, either natural or artificial and generally permanently exposed, within a delimited area and of some significance for their scientific, educational or interpretative value; they range from quarries and natural cliffs to mines and caves (Cope 2014) requiring husbandry rather than strict preservation, for much of their value lies in the access they provide to in situ rocks and their fossils and minerals. They can be refined on the nature of the localities at which geotouristic activities are focussed; for example coastal (van den Ancker & Jungerius, this volume, in press), mountainous/alpine (Cayla et al. 2015; Gordon & Baker 2015; Migon´ 2014; Whalley & Parkinson, this volume, in press), volcanic (Hose 2010a; Pullin 2014) and mining localities (Bristow

2015), and waterfalls (Hudson 2015). Tourists visiting waterfalls have a long history, and in the eighteenth century they were briefly termed ‘cataractists’, perhaps an interesting descriptor worth resurrecting. Secondary geosites have some feature(s) and/or item(s), within or on a structure or delimited area, of at least local significance to the history, development, presentation or interpretation of geology or geomorphology; these include museum, library and archive (Larwood 2014) collections, heritage/visitor centres, geologists’ residences, memorials along with commemorative plaques and monuments. What is much harder to determine is why, unless they record their reasons for particular actions, geotourists undertake a specific activity. However, inferences from their previous actions or from the context of their visits can be made. Thus, it might (probably erroneously) be inferred that participants on geology field trips are intent on learning about geology, when the prime motive might well be social, as Hose (2006) has suggested in drawing attention to the expressed, and possible motivations for, the behaviours of geology excursionists. It could be argued that geotourism, particularly of a more dedicated nature (Hose 2000, p. 136), is a restricted market dependent upon better-educated and wealthier tourists, broadly corresponding to Plog’s (1974) ‘allocentric’ tourists (Hall & Weiler 1992, p. 4), since their interest in geosites is mainly self-education and intellectual improvement. Alternatively, casual geotourists are drawn to geosites

Fig. 5. A ternary diagram of geology field excursionist types. This graphical summary is based upon several qualitative and observational studies completed, but mainly unpublished (Hose 2003), by the author; it is a development of an earlier model published in Hose (2006).

THREE CENTURIES OF APPRECIATING PHYSICAL LANDSCAPES

for pleasure and social interaction; they commonly visit for informal educational experiences for themselves and accompanying children (Hose 1996). It was almost solely due to European academics and practitioners that geotourism emerged in the 1990s as a field of study, publication and practice (Hose 2008). Due to the volume of published material, general usage (especially in policy documents), practice and growth (|Dowling 2011 2013), it can be argued that geotourism has a substantial enough theoretical and conceptual status that it now qualifies as a new geological paradigm. The Oxford English Dictionary defines the basic meaning of a paradigm as ‘a typical example or pattern of something; a pattern or model’, whilst the MerriamWebster Online Dictionary defines it as ‘a philosophical and theoretical framework of a scientific school or discipline within which theories, laws, and generalizations and the experiments performed in support of them are formulated; broadly: a philosophical or theoretical framework of any kind’. In planning the Appreciating Physical Landscapes: Geotourism 1670–1970 conference, and in editing this volume, the author considers that in accepting such definitions ‘geotourism’ unquestionably has attained the status of a modern geological paradigm. It has benefitted from several quite recent volumes (Dowling & Newsome 2008; Erfurt-Cooper & Cooper 2010; Newsome & Dowling 2010; Hose in press) and widely published, if sometimes divergent or vague, definitions. As a geologically based contemporary approach to essentially aesthetic landscape promotion, geotourism was first recognized and defined in England (Hose 1995a 2011). It was developed partly as a response to the loss, for geological field study, of an increasing number of the UK’s quarries and mines due to landfill and reclamation schemes (Ellis 1996, p. 83), as well as unsympathetic after uses that sought to recreate aesthetically pleasing vistas. Natural geosites and geomorphosites were concomitantly facing losses consequent upon planning decisions and design considerations for road and coastal defence construction (Ellis 1996, p. 83). For the former, soil covering and netting cuttings (Baird 1994), and for the latter, the pouring of much concrete (Leafe 1998), were particular problems. Similarly, changing agricultural practices in the UK’s uplands, such as drainage improvements and afforestation (Ellis 1996, p. 83) and new active leisure provision such as skiing and mountain biking (Ruff & Mellors 1993) have, or at least have been perceived to (Brown 2014), deleteriously affect geomorphosites. Gray (2008) has noted that ‘Just as “geodiversity” was developed as the abiotic equivalent of “biodiversity”, so “geotourism” has become a popular topic in recent years as the abiotic parallel

7

of “ecotourism”. Many definitions of “ecotourism” exist, including: “A tourism market based on an area’s natural resources that attempts to minimize the ecological impact of the tourism” and “Tourism supported by natural ecological attributes of an area”.’ (Gray 2008, p. 295). Hence, geotourism is a geoheritage promotional approach with parallels in ecotourism. Further, its antecedents lie in the landscape aesthetic movements that promoted travel, especially into ‘wild’ areas, popularly followed from the mid-eighteenth century onwards (Hose 2008) by the social elite of the UK and Europe. These movements partly built upon the experiences and publications that had arisen from the Grand Tour. On this, Stoye (1989) has noted the increasing number of travellers who, as part of their employment, accompanied merchants and the gentry but neglected the artisan journeymen; this omission undoubtedly reflected the nature of the available written resources and the limited market for travel publications – a constraint to consider when the Grand Tour is examined. Dann (1999) has suggested that, prior to the twentieth century, travel writing was essentially real-world reportage; this was usually a narrative of an actual journey in a foreign setting, ideally suited to the, especially imperial, explorations of the late Victorian age. ‘Hence the travel book dealt with facts, and this scientific emphasis was confirmed by an index, footnotes and a bibliography, similar to a treatise in geography, history or some allied discipline’ (Dann 1999, p. 162). He further suggested that such reportage lacked interest because it was merely a precise chronicle of observations and discoveries, lacking a critical or interpretative element. In 1811, the architect Henry Holland (1745–1806) disparagingly commented that ‘nobody you know, travels now a days without writing a quarto to tell the world where he has been, etcetera, what he has beheld’ (in Barton 1998, p. 3) and clearly would have welcomed the well-researched focussed accounts that began to appear within a decade – especially for England’s Lake District, if not for the Grand Tour.

The Grand Tour: the precursor to aesthetic landscape tourism The Grand Tour is Europe’s first widespread tourism phenomenon for which there is a considerable written and published record in the form of personal artwork, diaries, journals and letters; the information quality varies widely but usually does enable the elucidation of tourists’ itineraries, their impressions of the countries through which they passed, and details of the people, customs and accommodations. The Grand Tour’s accounts (Towner 1984), its

8

T. A. HOSE

participants and the possible reasons for its demise have been critically discussed by Towner (1985), who traced some 900. He cautioned that plagiarism, with the incorporation of material from guidebooks and journals written by other travellers suggesting thoroughness of observation but without any actual first-hand study, is an issue in both published journals and unpublished diaries and journals of the seventeenth and early eighteenth centuries. An illustrated, mainly descriptive introductory history has been published (Hibbert 1969) and further accounts cover its Italian (Black 1993; Wilton & Bignamini 1996) component and its cities (Sweet 2012), landscapes (Chard 1999) and British participants (Black 2003) in some detail. Its female participants (Chard 1999; Dolan 2001) have also been considered. The journey was first described by the Roman Catholic priest Lassels; he had tutored several young English noblemen with whom he travelled through Italy on five separate tours. His book’s introduction listed four arenas in which travelling through France and Italy would benefit young noblemen: the intellectual; the social; the political; and the ethical – by drawing moral instruction from what was seen. Thomas Nugent’s four-volume The Grand Tour . . . of 1749 was perhaps unusual for such a text in that it included topographical and geomorphological information; for example, volume 1, on Holland, has the sort of agrarian reportage – ‘The soil of Holland is hollow, soft and fenny, and not very fit for the plough; the country is flat and even, for that one sees neither hill nor mountain, except those little sandy hillocks, which are a barrier against the ocean’ (Nugent 1749, p.8) – that would have interested Arthur Young. In volume 3, on Italy, he notes on Naples that ‘there cannot in all respects be a more agreeable place to live in, did not the eruptions of neighbouring mount Vesuvius, together with the earthquakes sometimes disturb their quiet’ (Nugent 1749, p. 353). His preface to the four volumes stated that ‘Tho’ most gentlemen are presumed to have some knowledge of geography, yet as this is not always the case, a general description of the several countries is prefixed to each volume, with an account of the situation, extent, climate, soil, seas, rivers, and mountains’ (Nugent 1749, p. iv). Thomas Nugent (c. 1700–1772) was an Anglo-Irish antiquary who also published Travels through Germany . . . in 1768 following a 1766 journey. His accounts include much about antiquities, customs, fashions, manufactures and politics, together with sound practical advice to travellers. The Grand Tourists primarily visited those European cities considered major centres of culture, especially, Paris, Rome, and Venice; however, Florence and Naples were also popular destinations.

The latter meant a visit to Mount Vesuvius was inevitable; when antiquarians began to excavate Herculaneum and Pompeii, in 1738 and 1748, respectively, they became major destinations. A classic account, in the form of published letters, of a geographically extended (to Greece and Turkey) 1794–96 Grand Tour by John Bacon Sawrey Morritt (1985) is significant for its mention of Naples. Whilst his observations are mainly on cultural heritage and practical matters, he provides brief but useful information on the effect of the volcanic activity of Vesuvius on Torre del Greco, where ‘The place smokes still, and six months after the eruption the fire was seen running under it’ (Morritt 1985, p. 267). Of course, Vesuvius is where the origins of modern vulcanology can be traced. Sir William Hamilton (1731–1803), an antiquarian who studied both Vesuvius and Etna, was awarded the Copley Medal by the Royal Society in 1770 for ‘An Account of a Journey to Mount Etna’. He was posted to Naples as a diplomat from 1764 to 1800; his early years there coincided with Vesuvius becoming quite active. In 1766 he sent an account of the eruption, together with drawings and samples of salts and sulphurs, to the Royal Society, following which he was elected a Fellow. In 1767 Vesuvius again erupted and he duly sent a second report to the Royal Society; the two reports were published in the Society’s Philosophical Transactions. A recent biography (Constantine 2001) provides a summary of his vulcanological activities. The Grand Tour of some 2 to 4 years, as described by Lassels and Nugent, had begun in the sixteenth century and reached its acme in the eighteenth century. Its most popular itinerary, followed for some 300 years, began in Paris, thence through the Rhone valley and southern France to view classical Roman remains, after which came a crossing of the Alps followed by a tour of the cities of northern Italy (including Turin, Milan and Venice) before visiting Florence, Rome and Naples, and returning through Germany, the Rhine and Holland to England. However, whilst initially the tourists were chiefly focussed on art objects and architecture, by the late eighteenth century their interests and recorded responses had shifted towards the Romantic consideration of townscapes and landscapes; they were passionate about medieval structures and wild nature with its sublime and picturesque scenery. Indeed, much of the revised tour route can now be viewed as a form of ‘scenic tourism’; that is, early or historical geotourism – the focus of this volume. The tourists were mainly, but not exclusively, young English bachelors and a few women (Dolan 2001) of means; they were seeking to broaden their horizons whilst specifically learning about architecture, art, geography and

THREE CENTURIES OF APPRECIATING PHYSICAL LANDSCAPES

culture. Lady Mary Coke (1727–1811) undertook various European tours between 1756 and 1791 and despite subsequently publishing a 26-volume journal (recounting her experiences between 1766 and 1774), much of her observations remains in private hands and unpublished. She was not, as perusal of her gossip-ridden private life indicates, a typical Grand Tourist – most only did the one tour. Like the journals of Celia Fiennes, the great English domestic grand tourist, hers were originally intended only for family consumption. Unlike art and social reportage, landscape appreciation was initially a minor, generally accidental, commonly ignored element of the Tour. However, journal entries and letters provide much useful information on the perceptions and perils, particularly of mountainous country travel. As Towner (1985) has noted, the influence of the general cultural environment on travel motives can be seen through the increasing interest in the seventeenth century of Europe’s fine and decorative arts. Similarly there was an increase from the seventeenth century in the number of scientific travellers. Empirical factual accounts padded out many travel journal entries. This was especially stimulated in England by the establishment of the Royal Society in 1662. Following the rise of the professional, or at least gentlemanly, scientist this aspect of the tour declined after the 1780s. However, ‘An interest in natural scenery was of little consequence for the spatial pattern of the Grand Tour in the seventeenth century. While some interest was shown in rural scenery (particularly fertile, humanized landscapes), these sentiments were mainly confined to incidental observations while traveling to the major cultural centers’ (Towner 1985, p. 314). The principle cultural centre was Rome. It is to Rome that the beginnings of antiquarianism, the systematic study of sites and their artefacts, can be traced, which would later help to underpin geological discovery and reportage. Of course, getting to Italy meant a crossing of the Alps, and it is accounts of these crossings that give cultural historians the first glimpses of attitudes to and perceptions of such wild ‘frightful’ places. Other locations visited by the more adventurous Grand Tourists included Germany and east Europe, the Balkans and the Baltic, and Spain and Portugal. As Towner (1985) has noted the quest for reliable routes between the major centres that also provided scenic vistas explains why many focussed on the Rhine valley, especially between Mainz and Cologne, and its major tributary valleys such as the Moselle; indeed, such valley routes evolved from a convenient route to a scenic attraction in their own right. Further, ‘The desire to reach the Rhine quickly can be seen in the development of routes from Ostend to Brussels and Cologne’ (Towner 1985,

9

p. 314). This, along with changing tastes, might well explain the virtual abandonment of some formerly popular routes and areas, such as the Loire valley in the seventeenth and Holland in the eighteenth century. In alpine Europe, the significance of road improvements can be especially noted when ‘Wild scenery could be appreciated in the Jura and Italy and its northern lakes reached via a route made fit for carriages by Napoleon in the early 1800s’ (Towner 1985, p. 314). For those Grand Tourists sojourning in Switzerland, Lake Geneva was a focal point for tours to Chamonix and Lauterbrunnen and Grindelwald. Lake Lucerne was the centre from which an ascent of Mount Rigi (1798 m) commenced, a key element of the Grand Tour itinerary. Interestingly from an historical geotourism perspective, the mountain’s name is derived from ‘riginen’ which is the stratification clearly visible on its northern aspect. The principal route across the Alps into Italy was the Simplon Pass. This was the route taken by Wordsworth and a companion in 1790, after they had visited Chamonix, and described in letters to his sister and in the much later posthumously published The Prelude of 1850. Their mid-August crossing was prior to the Napoleonic widening, and they initially took the wrong route. Until the widespread publication of reliable topographic maps and detailed guidebooks (with instructions other than to hire a local guide!) such confusions were not uncommon. Wordsworth’s poetical account in book VI of The Prelude of his alpine ramblings is a little confusing, and the note to his sister that ‘At Brig we quitted the Valais, and passed the Alps at Simplon, in order to visit part of Italy. The impressions of three hours of our walk among the Alps will never be effaced’, at least makes it clear he found the journey memorable despite the low cloud and mist that made his route finding difficult. The Grand Tour was an institution that survived virtually unchanged until the French Revolution and the Napoleonic Wars, from 1789 to 1815, made it particularly hazardous for young British aristocrats, as well as the offspring of the rich mercantile class, to venture abroad. The effective closure of continental Europe to the British Grand Tourists provided the stimulus for the discovery and literary promotion of Britain’s upland landscapes; these were then promoted, perhaps astonishingly to the modern mind, as substitutes for the Alps with which Wordsworth and his literary and artistic contemporaries were themselves familiar from their own Grand Tours. At the same time the south and southwest coasts of England were promoted as the new Riviera, and Italianate villas and formal, often terraced, gardens were constructed; the beginnings of the English scenic tour

10

T. A. HOSE

had been established. Following the peace and relative political stability brought to Europe by the outcome of the Battle of Waterloo in 1815 and the rapid development of railways from the 1840s, the equivalent of the Grand Tour experience was available in a much shorter time frame to a much wider range of participants; elite tourism was giving way to the first wave of mass tourism.

Geotourism contextualized within European landscape tourism The first published history of UK geotourism (Hose 2008), although an earlier consideration of the historical aspects of Europe’s published geoheritage had been provided (Hose 2000), noted that ‘Since many of the activities it encompasses have antecedents in considerably earlier natural science and aesthetic movements, the development of tourism that can be directly attributed to the promotion of landscape and geology must be examined; the former dating from the late seventeenth and the latter from the early nineteenth centuries’ (Hose 2008, p. 37). The second such account (Hose 2010a), focussed on volcanic geotourism on Scotland’s west coast, also noted that ‘From the early 19th century, commercial tourist literature generated specific expectations of landscapes. Landscapes are social and cultural constructs with tourists’ perceptions and their values ascribed to them an admixture of direct observation and cultural interpretation’ (Hose 2010a, p. 260). It has been specifically noted that mining locations ‘must be analyzed

within the context of the scenery that surrounds them, not only in terms of monuments or buildings, but also as part of a series of relics that indicate a continuous process of change, which in turn is a result of the inevitable interaction of all human activities with the surrounding environment. In other words, scenery is the mirror of society, and it expresses the physical and symbolic effects of successive human actions on nature. However, at the same time, this scenery is not neutral, for each person adopts his/her own point of view. One consequence of this is the reductive and unilateral attitude that interprets and values scenery from a strictly visual and aesthetic perspective’ (Edwards & Llurdes 1996, p. 358). Accordingly, this paper touches upon these, particularly insofar as they are explored by other papers in this volume. Encompassing some 300 years of landscape-focussed tourism it is best approached by examining locations, personalities, publications and images key to its development. From the Renaissance, leisure travellers in Europe and farther afield selected the landscapes that best matched their quest for the novel, exotic and authentic. Specific expectations about the places and landscapes they planned to or did visit were created and influenced by images in art galleries and their reproduction in postcards and guidebooks together with the latter’s place descriptions and instructions for travellers. The perceptions of landscapes and the values ascribed to them by travellers are a combination of their processing of direct observation and subsequent interpretation influenced by their education and expectations, the cultural filter of the ‘tourist

Fig. 6. Model of the tourist gaze. This model, developed from that published in Hose (2010a, 2010b), notes the input of artistic and literary material that influences the content of tourist guidebooks; it also notes that artists and authors are partially a subset of travellers and tourists. It stresses the significance of the tourist guidebook acting as a cultural filter on what the tourist sees and understands about landscape. It emphasizes that what the tourist actually sees lies beyond the mere physical landscape that is viewed, suggesting for some an aesthetic spiritual element – much as was envisaged by the Picturesque movement.

THREE CENTURIES OF APPRECIATING PHYSICAL LANDSCAPES

gaze’ (Fig. 6). Landscapes are an admixture of elements ordered and bounded by travellers’ knowledge and experiences. Across Europe and elsewhere different regions with similar landscapes were visited and recorded by travellers (Bristow 2015; Hudson 2015; Mather 2014; Vasiljevic´ et al. 2014), tourists (Cayla et al. 2015), artists (Ancker & Jungerius, this volume, in press; Pullin 2014), writers and poets (this paper), geographers (Henry & Hose 2015), geologists and geomorphologists (Bristow 2015; Larwood 2014; Whalley & Parkinson, this volume, in press), and naturalists (Burek & Hose 2015) who defined, delineated, described and depicted landscapes from their different mindsets. Within Britain three mountainous areas (the Peak District, the Lake District and the Scottish Highlands) and one coastal area (central southern England) were significant in the development of geotourism. Within mainland Europe the key areas were mainly mountainous, not necessarily truly alpine, volcanic and latterly coastal places that were initially visited as a component of the Grand Tour (Hibbert 1969). Whilst the historical approaches to landscape-based tourism are well covered for the UK, elsewhere in Europe and beyond the coverage is rather patchy. Meanwhile, it is worth noting that the term ‘tourist’ first appeared as an English synonym for ‘traveller’ in the late eighteenth century; it has been asserted that ‘The traveller exhibits boldness and gritty endurance under all conditions (being true to the etymology of “travel” in the word “travail”); the tourist is the cautious, pampered unit of a leisure industry’ (Buzzard 1993, p. 2). It was incorporated within the title of what is probably the UK’s first national guidebook, Mavor’s multi-volume The British Tourists; or, Traveller’s Pocket Companion, through England, Wales, Scotland, and Ireland published between 1798 and 1800 (although the guidebook proper was only established in the early nineteenth century). More adventurous travellers did venture to Scandinavia; Barton (1998) provides an account of those who visited between 1765 and 1815 through the accounts of some 30 travellers of whom three were Scots and 11 English. Amongst the former was the mineralogist Thomas Thomson (1773–1852); he visited Sweden in 1812 and made detailed geological observations. Amongst the latter was the accomplished amateur artist John Thomas James (1786–1828); he visited Sweden for much of 1813 and 1814 and provided an illustrated account of his travels. Both recorded aspects of Swedish life and customs, as well as the landscapes in which they travelled – not unlike the mid-nineteenth-century visitors to northern Norway reported by Whalley & Parkinson (this volume, in press), who also notes the mineralogist Leopold von Buch’s 1810 Reise

11

durch Norwegen und Lappland. Buch (1774– 1853) was the first foreign member of the Geological Society of London and introduced the term ‘caldera’ into the geological literature. He visited Norway and Sweden between 1806 and 1808. As Barton (1998) has noted, his travellers were mainly British and ‘Wealth and leisure . . . had made them pioneers of tourism. So that the foreign traveler abroad would be almost axiomatically taken for an English milord’ (Barton 1998, p. 13). The opening years of the 1670s were a period of political turmoil within Europe, with numerous wars; by their end at least, through the Treaties of Nijmegen, the Franco-Dutch war was ended and touring Europe for pleasure became somewhat safer. By then, the Grand Tour was an established feature of European travel, although the term was only introduced into the literature in 1670; this was by Richard Lassels in his posthumously published (initially in Paris and later London) two-volume Voyage to Italy, which went to several editions. By then some pioneering geological work in the Italian city states had been completed and published by antiquarian scholars such as Pirro Ligorio (c. 1510–1583) and Bartolomeo Marliano (1488– 1566); they had published maps and illustrations of Rome’s topography and ancient buildings. Following the 1570 Ferrara earthquake, Ligorio was placed in charge of a group summoned to the city to study seismological events and conduct earthquake research – the first such European scientific effort. Ligorio’s subsequent treatise, Rimedi contra terremoti per la sicurezza degli edifici (Remedies against earthquakes for building security), included plans for an earthquake-proof building; this was the first such design anywhere. Such Renaissance scholars did not focus on a single narrow field but followed multiple interests. The Danish scholar Ole Worm (1588–1655), who preferred the study of antiquity, assembled at his Copenhagen home a large collection of curiosities; these ranged from American native artefacts to stuffed animals and fossils. Significantly for early geotourism he compiled engravings of his collection, along with his speculations about their meanings and origins, into the posthumously published Museum Wormianum of 1655; its frontispiece is one of the best-known illustrations of a ‘cabinet of curiosities’ – the forerunner of the modern natural history museum. By the mid-seventeenth century, antiquarians such as Worm had embarked upon the study of antiquities as a discrete discipline to generate a theoretical framework for them; progressing from merely describing monuments to explaining their functions; in a well-illustrated account Schnapp (1996) has summarized their original approach and their later adoption of methodology pioneered by geologists. It required that

12

T. A. HOSE

they examined the ground and excavated sites as a means of envisioning the lifestyles of vanished peoples without any written history. They also toured and personally recorded sites. Their work was commonly published in county studies and then from 1751 in the Society of Antiquaries journal. In seventeenth-century England Robert Plot and Edward Lhuyd, the Ashmolean Museum’s first and second keepers, were typical of the new generation of antiquarians who regarded the study of antiquities as part of natural history. Fossils, including the first if unrecognized dinosaur bone, were amongst the antiquities illustrated in Plot’s Natural History of Oxford-shire (1677). His later Natural History of Stafford-shire (1686) is noteworthy for its account of pottery clays. Published guides began to appear in the late eighteenth century. After the mid-eighteenth century scientific the study of geology was a mainly British occupation with significant French and German contributions, although there were still some important Italian contributions. Before then, from the seventeenth century the groundwork of scientific geology was laid in the Italian city states by either native or domiciled scholars. The best known of the latter is Nicholas Steno (1638–1686), originally from Copenhagen. After dissecting in 1666 the head of a huge shark caught off Livorno he published his findings, accompanied by one of the most widely known natural history images, in Elementorum Myologiae Specimen (1667). He recognized its teeth were similar to the Glossopetrae or ‘tongue stones’ unearthed from Malta’s Miocene limestones. However, the Italian naturalist Fabio Colonna had already stated in De glossopetris dissertatio (1616) that they were sharks’ teeth. Because they seemingly looked like either the forked tongue or fangs of snakes from which Saint Paul was supposed to have removed the venom they were considered in medieval and Renaissance Europe to protect against poisoning. There was a profitable trade in Glossopetrae, and Steno’s revelations are almost certainly the first by a geologist to undermine economically significant local geotourism! Steno also suggested that rocks formed when particles in fluids, such as water, settled out into horizontal layers; any deviations from this were due to later disturbances – his ‘principle of original horizontality’. He suggested that the youngest layers must be those at the top, with the oldest at the bottom – his ‘law of superposition’. Steno noticed that in the Apennines near Florence the upper rocks were richly fossiliferous and the lower ones lacked fossils; he suggested the former were created in the biblical Flood (after the creation of life) and the latter before life existed – the first geological explanation distinguishing different periods of Earth history. This work led to his epithet – ‘Father of Stratigraphy’.

His ideas were summarized in De solido intra solidum naturaliter contento dissertationis prodromus (1669), which was widely circulated and even translated into English; it was intended as an introduction to a much larger work that, due to his conversion to Roman Catholicism in 1667, was never completed. Italian mercenaries, scholars and travellers traversed Europe, especially within the AustroHungarian Empire, during the seventeenth and eighteenth centuries. The specimens they collected and texts they published on their observations form the core of several major Italian museums and libraries. The most significant of such travellers for geotourism was the army officer, also a naturalist and hydrographer, Luigi Ferdinando Marsigli (1658–1730) from Bologna (see Vasiljevic´ et al. 2014). His duties within the Austro-Hungarian Empire took him into what are present-day Albania, Austria, Bosnia, Bulgaria, Croatia, the Czech Republic, Hungary and Poland, where he recorded and published observations on topography, rivers, lakes and natural history. His seminal geomorphological work was the mapping and description of the Danube basin, finally published in his multivolume Danubius Pannonico-Mysicus (1726), the main reason he was elected a Fellow of the Royal Society. This work included the significant recognition of the thick loess deposits seen along the banks of the Danube (Vasiljevic´ et al. 2014). He founded several art and science institutions (such as the Instituto delle Scienze), and others (especially the Museo di palazio Poggi) now also house his natural history, art and book collections. From the late eighteenth century onwards geology field excursions were, like those of the earlier naturalists and antiquarians such as Marsigli, almost exclusively undertaken by men of high social and economic standing, sometimes members of the Church; this indicates a shift away from the previous restrictive sociocultural conventions to permit their venturing into ‘wild’ landscapes for pleasure rather than necessity. This shift was part of a developing wider focus on natural history studies somewhat fostered by the realization that ‘even the amateur could hope to make significant contributions and participate in important national, even international scientific endeavours’ (Bedell 2001, pp. 4– 5); the post-Enlightenment move towards romanticism was also a contributory factor. They generally trod in the footsteps and were guided by the publications of earlier travellers engaged in antiquarian, agricultural, industrial and socio-economic reportage. Up to the mid-eighteenth century the ‘preferred rural landscape was generally a humanised scene of cultivation, evidence of the successful control of nature’ (Towner 1996, p. 138). The ‘wild’ areas, like those geologically mapped in the

THREE CENTURIES OF APPRECIATING PHYSICAL LANDSCAPES

Lake District in the nineteenth century’s first half by Otley and Sedgwick (Oldroyd 2002) and in Scotland by MacCulloch, were seen as waste places without any real economic value. The establishment in 1830 of the Royal Geographical Society (RGS) of London followed several others in the UK devoted to promoting field study and travel; these included the Linnean Society (established in 1788), the Geological Society (1807), the Zoological Society (1826) and the Raleigh Club of Travellers (1826). The latter, at a meeting on 24th May 1830, approved the motion ‘that a new and useful Society might . . . be formed, under the name of The Geographical Society of London.’ and was thus the parent body of the RGS. The Raleigh Club was a dining club, the members of which claimed collectively to have visited every part of the known world. At a special meeting of the Raleigh Club in 1854 it was dissolved and replaced by the Geographical Club with Sir Roderick Impey Murchison as its first president. Meanwhile the RGS increasingly became the main focus of global exploration partly due to its considerable and influential membership – 800 Fellows in 1850 (twice that of its Berlin and eight times that of its Paris equivalents) and 2400 in 1870. It published general advice from 1854 through its Hints to Travellers and established probably the world’s largest private topographical map collection. Its success reflected the strength of British amateur natural science, the wealth of the country’s upper middle class (which provided the bulk of the fellowship), and the emergence of professional field scientists. Most of its Fellows were amateur scholars, but some prominent scientists also joined its ranks, including Charles Darwin. Its dominant figure during the mid-nineteenth century was Murchison, who was its president in 1843–45 and thrice again in 1851–53, 1856–59 and 1862 –71. This was the same Murchison whose sole fieldwork in ‘wild’ Wales and the Welsh Borderland had established the Silurian system and whose rapid coach traverse of Russia the Permian system; he had also, following joint fieldwork with Sedgwick, established the Devonian system. Murchison was appointed director-general of the British Geological Survey and director of the Royal School of Mines and the Museum of Practical Geology in 1855, succeeding Sir Henry De la Beche, who had first held these offices. De la Beche was a prolific author of such lengthy texts on stratigraphy and geological methods as A Geological Manual (1831), How to Observe (1835) and The Geological Observer (1851). He was also well travelled, as can be seen in his Sections and Views Illustrative of Geological Phenomena (1830) with its many plates and descriptions of alpine and also exotic locations (such as Jamaica) – the geological equivalent of the

13

‘Grand Tour’ guidebook. However, by then the Grand Tour had almost vanished in its classical form as an education for young noblemen. Murchison was an inveterate traveller in Europe with an interest in promoting exploration for trade purposes. He was president of the Geological Society in 1831 having been its secretary for five years. It established a museum in 1808 and a library in 1809; today the museum has gone but the library is the UK’s largest such establishment. It has one the finest collections, approximately 3500 volumes, of antiquarian books on geology and related subjects; most were published after 1800, but some eighteenthcentury volumes plus a few from the sixteenth century and impressive early volumes annotated by those nineteenth-century geologists who originally owned them are in the collection. In 1858 the Geologists’ Association was founded, with a more inclusive membership than the Geological Society, and from its outset promoted field excursions; at first in the London area and then, taking advantage of the expanding railway network, further afield – including the Peak District and the Lake District and eventually the Continent. By the late eighteenth century interest and pleasure in wild areas and some geological matters are found in travellers’ writings, including those on the Grand Tour. Nineteenth-century to Great War guidebooks and travel literature for Europe, as they reflected and influenced emerging mass tourism, have been critically explored by Buzzard (1993). Withey (1998) has provided a mid-eighteenthcentury to Great War overview of European leisure travellers that essentially encompasses the Grand Tour in both elite and populist forms as well as travel beyond Europe. With the various political upheavals and wars affecting Europe from the late eighteenth to the early nineteenth centuries, British travellers increasingly found their journeys restricted to the British Isles. These early leisure travellers were directed by the published selections of other travellers, diarists, poets and artists promoting various landscape aesthetic movements that had some inspiration from the experiences of the Grand Tour. The first of these, the Sublime movement, was promoted in mid-eighteenth-century England by Edmund Burke in A Philosophical Enquiry into the Origin of Our Ideas of the Sublime and Beautiful of 1757; in this book he equated the Sublime with astonishment, fear, pain, roughness and obscurity. However, the later Romantics associated it with the tumultuous chaos of mountains lying beyond rolling foothills, deep valleys and dangerous rocky precipices. The poet William Wordsworth (1770– 1850) suggested in the fifth and last edition of A Guide Through the District of the Lakes . . . that it was ‘the result of Nature’s first great dealings with

14

T. A. HOSE

the superficies of the earth’ (Wordsworth 1835, p. 35). The wild ruggedness of his beloved Lake District’s great rock masses, fells and lakes were to solicit from travellers, accustomed to appreciate southern England’s agricultural and parkland landscapes as evidence of Man’s control over Nature, awe and wonder. Travellers in search of the Picturesque movement’s landscapes sought locations that showed the softer effects of Nature’s subsequent operations that led to the variegation and harmony expressed by a meandering river’s curve or a lake shore, the grouping of their flanking trees, the interplay of light and shade and the subtle colour gradations that framed the vista. That vista was also filtered by many artists and travellers by a Claude Glass. This was a small, slightly convex, dark-tinted mirror, carried in a fitted carrying case, which limited the breadth of the vista and reduced and simplified a landscape’s natural colours and tonal range to that which could readily be painted. From the late eighteenth century the Picturesque topographical approach was literally appreciated from scenic ‘stations’ such as those established by Thomas West (1720–1779) from 1778 in the Lake District. West advised in his A Guide to the Lakes: Dedicated to the Lovers of Landscape Studies, and to All Who Have Visited, or Intend to Visit the Lakes in Cumberland, Westmorland and Lancashire of 1778 that ‘The person using it ought always to turn his back to the object that he views. It should be suspended by the upper part of the case . . . holding it a little to the right or the left (as the position of the parts to be viewed require)’. The Picturesque, from around 1780 to 1850, was an all embracing aesthetic approach that sought to provoke travellers’ emotional reflections on landscapes (see Hebron 2006) and their evocation in visual art and literature. These aesthetic movements were influential in the later years of the Grand Tour. They reflected three interrelated elements: the travellers’ type and purpose; the meanings ascribed to, and understandings of, natural phenomena; the shift from an essentially agrarian to a majorly industrial society and the concomitant rise of the middle classes in numbers and influence. From the mid-nineteenth century, European travel was increasingly opened up through, and literally speeded up by, the spread of the railway network, which wrought as much social as economic change (Faith 1994). The rail network was most conveniently centred on London and Paris; this permitted a new form of the Grand Tour in which the same locations were visited, but stays were shorter, and the tourists were mainly drawn from the middle classes. The railways also often provided the newest well-equipped hotels that accommodated the large numbers of travellers initially disgorged at their main termini. The speed and general

reliability of the railways meant that as much ground could be covered in a few weeks or a month as previously would have taken tourists half a year or more. Undoubtedly, the golden age of long-distance Continental rail travel, especially aboard trains of the Compagnie Internationale des Wagons-Lits (et des grands express europe´ens), was represented by the closing and opening decades of the nineteenth and twentieth centuries. Continental holidays came to an abrupt end with the outbreak of the Great War, when the railways turned to shipping troops rather than tourists. With the coming of peace the reformed political map of interwar Europe and the poor economic situation meant that British and especially German middle-class Continental travel was much reduced. Whereas in the original Grand Tour the journey was part of the educational and cultural experience, it was no longer a significant element but a necessary evil to reach the tourist locations. These were increasingly recorded photographically, especially with the advent of roll-film cameras in 1888, rather than in sketches and water-colours. The private tutor was very much replaced by the local guide and the comprehensive illustrated tourist guidebook; few ventured afar without their Murray or Baedeker. However, the keeping of comprehensive journals declined, as did their publication – their place was taken by the commercially produced tourist guidebooks. Parsons (2007) provides a populist extended history of the guidebook. Sillitoe (1995) summarizes their nineteenth-century to Great War history and usage. Vaughn (1974) provides a well-illustrated account with some fair analysis of late eighteenth- to late nineteenth-century English guidebooks. Because the British and Germans were the first to have the money, leisure and intellectual curiosity to travel in any numbers, their needs drove the nineteenth-century development of the tourist guidebook (Sillitoe 1995). From the 1820s, educated and elite travellers carried John Murray’s ‘handbooks’; his compendia focussed on visitors’ perspectives of what was important, including where they should bank, eat and stay. By the mid-nineteenth century, elite travellers were a less important market compared to the burgeoning middle classes of more modest means and limited, especially cultural, education; the Ward Lock Guides published from 1854 were intended to meet their less demanding needs. Karl Baedeker’s Handbuchlein, published from 1828, adopted Murray’s term rather than guidebook; they were the first to use asterisks as recommendations for significant sites and sights. By the late nineteenth century, tourists were well provided with good-quality affordable guidebooks, many of which mentioned landscape and geological features; however, such mentions were not always accurate,

THREE CENTURIES OF APPRECIATING PHYSICAL LANDSCAPES

as can be gauged from ‘took the ferry to northern Ireland for a day trip to the Giant’s Causeway. Formations among the huge granite [they are actually basalt!] blocks seemed like pipes of a great church organ’ (Hindley 1983, pp. 52–53). Domestic tourism became most significant in the interwar years of the twentieth century. Across Europe an interest in the outdoors saw leisure and touring cycling in the 1930s enjoy perhaps even greater popularity than it had achieved in the 1880s and 1890s. Those two decades saw improvements in roads and recommended accommodations as well as topographical maps (Henry & Hose 2015), which would later be used by motorists, to meet the needs of touring cyclists; in the UK, the founding of the Cyclists’ Touring Club in 1883 helped to push forward these improvements. Hiking (see Marples 1959) and hostelling (see Porter 1992) were also widespread leisure pursuits in the 1930s that persisted veritably unchanged into the 1960s. The former was promoted by the laying on of special excursion trains from major towns and cities, and many of the railway companies’ publicity departments also produced hikers’ guides and pamphlets. The railways were then facing competition from the developing motor-coach routes and private motorists – the latter would eventually dominate domestic tourism in the post-war years, especially from the 1960s. By then, the middle classes could readily tour with car and caravan, especially in the scenic uplands of central and northern Britain. Contemporaneously, Watkins-Pitchford (1959) wrote and finely illustrated a classic account of a tour to Scotland completed before restrictions were placed on where caravans could be parked overnight; access to the UK’s open countryside, although better than in pre-war days, was already adapting to issues around demand and sustainable management. In the UK, the National Parks and Access to the Countryside Act of 1949 led to the establishment in the 1950s of 10 national parks, with the first being the Peak District and the second the Lake District in 1951; similar provision emerged in Europe, although the first had been established as early as 1909 in Sweden and 1914 in Switzerland. From the mid-1960s there was considerable investment in the UK and Europe in environmental management for leisure (Fairbrother 1970) and interpretation, including some geological interpretation – the volume and nature of which is worthy of a volume in its own right. In the UK this was stimulated in 1963 by National Nature Week, repeated in 1966. It was by then recognized that visitors to nature reserves and the National Parks wanted information, and the managers of these lands needed to manage visitors to reduce their environmental impact. Environmental interpretation,

15

often employing the lessons learned and applied by the US National Parks Service, were readily adopted and adapted in the UK by the Countryside Commission (Aldridge 1975), countryside NGOs (Beazley 1971), and subsequently also in Europe – the basis of modern geotourism was being established. Environmental interpretation (Knudson et al. 1995) has been shown to benefit, in terms of engaging and building empathy with its audience, by incorporating human stories and history – a practical outcome of historical geotourism studies.

British landscapes, romanticism and geotourism Consideration of areas key to the development of British scenic tourism, such as those in northern England (Cope 2014; Henry & Hose 2015) and Scotland (Gordon & Baker 2015) is useful in elucidating generic European and global themes together with issues in historical geotourism. Initially the Peak District, considered the birthplace of British geotourism (Hose 2008), was favoured by Britons determined to venture in wild landscapes. Even the odd Continental visitor made his or her way to the Peak District. Perhaps the best known of these is Karl Phillip Moritz (1756–1793) who published Reisen eines Deutschen in England im Jahre 1782 (Journeys of a German in England in 1782) in Germany in 1783 and 1785; an English translation of 1795 was Travels, chiefly on Foot, through several parts of England in 1782, described in Letters to a Friend. Starting in London he, unusually for the times, walked via Leicester and Derby to the Peak District, where he visited Peak Cavern; he is therefore probably the earliest recorded foreign geotourist in England. England’s own first recorded geotourist was Celia Fiennes (Morris 1949), a privileged late seventeenth-century horseback traveller who visited the Ashbourne copper mines in 1698. Whilst Fiennes can be credited as the first English geotourist in that she purposefully visited and wrote about sites of some geological interest, a few earlier travellers incidentally visited and recorded their observations on such places. One of the earliest of these is John Taylor (1580– 1653), the ‘water poet’, who recorded his 1618 visit to the undersea coal mines of Culross in Scotland (Taylor 1888, pp. 43–46). The Peak District is also the region for which the first tourist guides were published; the development of its topographic and geological maps has been examined by Henry & Hose (2015). Its major attractions were organized, described and promoted by Thomas Hobbes (De Mirabilibis Pecci: Being the Wonders of the Peak in Darby-shire . . . of 1678) and Charles Cotton

16

T. A. HOSE

(The Wonders of the Peake of 1681); both recognized seven ‘wonders’, of which five had a geological basis as caves and springs. However, the perceived and sometimes reported overt commercialization of the Peak District’s major sites led travellers to look elsewhere, at first to north Yorkshire, then the Lake District and finally the Scottish Highlands and Islands (Gordon & Baker 2015; Hose 2010a). The Lake District is particularly significant to historical geotourism because of the means by which its exploration opened up a supposedly remote wild region to leisure travellers. Both amateur and professional artists were particularly attracted to its antiquities, lakes and mountains. Their landscape paintings and sketches were occasionally published as sets of engravings, sometimes with accompanying prose or poetry, or were included in several guidebooks. The relationship between English landscape painting and geology in the nineteenth century has been explored by Pointon (1978) in a seminal illustrated essay; she noted that travel was as important to artists as to geologists and that for the former their visual representations varied, according to purpose and employment, between accurate depiction and artistic creation. In 1806 on his sole Lake District visit John Constable (1776–1837) sketched and painted around Brathay, Kendal Skelwith, Thirlmere, and Windermere; his water-colours Windermere and The Castle Rock, Borrowdale show some familiarity with geology. The young Joseph William Mallard Turner (1775–1851) earned a living as a topographical artist; in 1798 he exhibited two Lake District paintings, with Morning amongst the Coniston Fells, Cumberland being noteworthy, at the Royal Academy in London. John Glover (1767– 1849), a rival of Turner, had his own gallery in London’s fashionable Old Bond Street and exhibited Lake District paintings from 1795 onwards; his c. 1820 painting Thirlmere and c. 1831 pen-and-wash Derwentwater typify his faithfully detailed depictions. He actually lived in the Lake District from around 1818 to 1820. He eventually settled in Tasmania in 1831 and helped to establish Australian landscape painting. Original landscape paintings were at first viewed by the social elite in private, including the Royal Academy in London, and commercial galleries. Following the establishment of Local Authority public art galleries, which was enabled by legislation passed in the 1840s, some landscape artworks could be viewed in the provinces by the increasing numbers of the urban middle classes. Those unable to view the originals due to the constraints imposed by either social class or geography benefitted from the 1840s with the introduction of chromolithography that facilitated the printing of reasonably priced colour reproductions. Gillett (1990) provides

a summary of the characteristics of the art public, although with something of a metropolitan focus, in late nineteenth-century Britain. From the 1850s monochrome photography was increasingly employed as an art medium to record landscapes. A generally little-known, especially when compared with the Abraham brothers, Lake District photographer was Ambleside-based Moses Bowness, who operated a studio between 1856 and 1894. Apart from the usual and lucrative portraiture work he also photographed the region’s mining and industrial sites and landscapes; some of these were reproduced as postcards with one being a particularly well-known image of the Bowder Stone in Borrowdale. Joseph Lowe (1865–1934) was a Patterdalebased landscape photographer who produced and sold postcards. His particular hobby just happened to be geology, and he often gave lectures illustrated with his own lantern slides on the region’s geology. Travellers unfamiliar with the Lake District had their expectations of what they would and should see influenced by various colour original artworks and reproduced images, some of which appeared as monochrome engravings in guidebooks; from the nineteenth century’s last quarter photo postcards were also quite widely available. Social historians, writers of literary criticisms and poets have charted the Lake District’s emergence over some 200 years from comparative obscurity to one of the UK’s most visited regions (see Victoria & Albert Museum 1984); likewise its geological exploration has been critically charted (Oldroyd 2002). Celia Fiennes had ridden through it in 1698 but thought it an unprofitable wilderness and merely recorded its potted char and some bread recipes. Daniel Defoe in the 1720s also thought it a barren and frightful place. From the 1750s travellers (popularly called ‘Lakers’ due to it also being known as The Lakes) visited the Lake District because its landscapes and antiquities were by then perceived to be of some quality. The poet Thomas Gray (1716–1771) toured in 1767 and 1769; his letters describing the 1769 tour were published in William Mason’s posthumous edition of his work in 1775 and helped to establish the region’s main visitor points, later to become tourist ‘stations’. From Penrith, Gray visited Ullswater, and from Keswick, he explored Derwentwater, Borrowdale, Bassenthwaite and the Castlerigg stone circle; then, after journeying beneath Helvellyn’s foot, he explored Ambleside, Grasmere and Windermere, before finishing at Kendal – routes followed by modern (geo)tourists. Mirroring contemporary travellers’ mindsets he wrote that the Derwentwater to Borrowdale journey, which he had undertaken on 3 October 1769, was akin to those alpine passes where travellers were imperilled by avalanches; he actually coined, reflecting his perceived peril on

THREE CENTURIES OF APPRECIATING PHYSICAL LANDSCAPES

the journey, the term the ‘jaws of Borrowdale’ (Toynbee & Whibley 1971, p. 1079) for the point where the steep ice-carved slopes of Kings How and Castle Crag confine the route from Keswick into the valley. The artist William Gilpin (1724–1804) toured in 1772, and the agrarian observer Arthur Young (1741–1820) in 1768. In his Observations on the River Wye and several parts of South Wales, etc. relative chiefly to Picturesque Beauty; made in the summer of the year 1770 of 1782, Gilpin remarked, similarly to Gray on the road into Borrowdale, that it ‘grew wilder, and more romantic . . . riding along the edge of a precipice, unguarded by any parapet, under impending rocks, which threaten above’ (Gilpin 1786, vol.1, p. 187); such an actual alpine scene is captured in Turner’s painting The Passage of the St Gothard of 1704. Young described Derwentwater as elegant but noted that its surrounding mountains were wild with dreadful chasms; likewise he observed that ‘Twelve of the fifteen miles from Shapp to Kendal are a continued chain of mountainous moors, totally uncultivated one dreary prospect’ (Young 1770, vol. 3, p. 169). Yet in 1774 William Hutchison justified publication of An Excursion to the Lakes, in Westmoreland and Cumberland, August 1773, because ‘When ever I have read the descriptions given by travellers of foreign countries, in which their beauties and antiquities were lavishly praised, I have always regretted a neglect which has long attended the delightful scenes at home’ (Hutchinson 1774, p. 1). Thomas West ordered the Lake District’s principal locations and sights, with recommended ‘stations’, into its first tourist guidebook, A Guide to the Lakes. . . . It was, when published in 1778, one of the first guides to promote the Picturesque movement and was a major commercial success. It mainly consists of descriptions with the precision required before detailed maps were available to find his numbered stations. Coniston’s first station above Nibthwaite, for example, could be found as the lake came into view ‘by observing an ash tree on the west side of the road, and passing that till you are in a line with the peninsula’ (West 1778, pp. 50 –51). West was one of the first writers to challenge the widely held perception that the north of England was a wild and savage place. Arthur Young, after his 1786 tour, suggested that paths should be cut and resting places created for travellers because ‘many edges of precipices, bold projections of rock, pendent cliffs, and wild romantic spots, which command the most delicious scenes, but which cannot be reached without the most perilous difficulty’ (Young 1770, p. 155); he further suggested the pruning or felling or trees obscuring views. Peter Crosthwaite published from 1783 maps showing his own and West’s stations. He

17

improved access to some of his stations; for example for one of his stations above Derwentwater he had steps cut and a cross marked on the ground. Derwentwater was very popular with early tourists. West’s guidebook recorded that its view from the Cockshott Hill station was close to the ideal requirements of the Picturesque because in ‘a spacious amphitheatre, of the most picturesque mountains imaginable, an elegant sheet of water is spread out before you, shining like a mirror, and transparent as chrystal; variegated with islands . . . clothed with forest verdure’ (West 1778, pp. 89 –90). Popular stations commonly had shelters provided for travellers, but they were not universally welcomed; in 1799 James Plumptree (1771–1832), the playwright and author of The Lakers: A Comic Opera in Three Acts, was of the opinion that Windermere’s first station somewhat lacked bucolic appeal. The Lake District was promoted to elite tourists by William Green (1760–1823), who published The Tourist’s New Guide in 1819; it is a volume packed with detail and was fulsomely praised by Wordsworth in his own guidebook. Beginning as a surveyor in 1778, Green established a studio in Keswick after about 1800. As a successful topographic artist he produced a significant body of detailed scientifically observed aquatints, etchings and water-colours. Wordsworth’s innovative guidebook linked landscape features to natural history (including geology), history and people alongside descriptions of what could be seen. It evolved from his anonymous text that accompanied the Rev. Joseph Wilkinson’s (1764–1831) Select Views in Cumberland, Westmoreland and Lancashire . . . 1810 volume of engravings. In 1820 it became an appendix to his River Duddon sonnets. By 1822 it was a guidebook in its own right and was successful enough to go to several editions up to the 1840s that eventually included three geology letters commissioned from Adam Sedgwick. This was not a novel or singular natural history inclusion, since the 1842 edition also included botanical notes and John Hudson’s 1842 Complete Guide to the Lakes was supplemented by geological notes also written by Adam Sedgwick. Wordsworth’s relationships with the leading geologists of his day, and how his poetry reflects this, has been examined by Wyatt (1995), who noted that Sedgwick had corresponded with Wordsworth, concerned that his geology letters might lack interest (Wyatt 1995, p. 169). The poet’s 1814 The Excursion has the stanza ‘He who with pockethammer smites the edge of luckless rock . . . detaching by the stroke a chip or splinter – to resolve his doubts; And, with that ready answer satisfied, he substance classes by some barbarous name, And hurries on . . .’ suggesting a familiarity with geology field practice and difficult petrological

18

T. A. HOSE

nomenclature. Wordsworth’s approach to the picturesque as developed in his guidebook has been examined by Nabholtz (1964), who noted that its last section was an attempt to prove that the Lake District’s landscapes are better suited to painting than those of the Alps. A few geology guides to the Lake District were also published in the nineteenth century. Jonathan Otley published A Concise Description of the English Lakes . . . and Observations on the Mineralogy and Geology of the District, the first populist account, in 1823. The style of the region’s later guidebooks shifted from mere station descriptions to holistic accounts for landscape students. For example Charles Mackay’s (1814– 1889) illustrated The Scenery and Poetry of the English Lakes: A Summer Ramble of 1846 was noteworthy for a text that focussed the readers’ attentions upon site-centred poetry and other literary allusions; he used his account of Stockghyll Force as an excuse to incorporate the waterfall description in Shelley’s 1816 Alastor; or, The Spirit of Solitude, in which the speaker recounts the life of a poet zealously pursuing the most obscure part of nature whilst seeking out the peculiar truths in obscure places. The creation of such literary landscapes was an enduring legacy of these early travel writers pervading populist tourist publications into the late twentieth century. The Bowder Stone (a 1250-tonne perilously balanced boulder) was an early example of a commercial geotourism enterprise in the region. It was described by West as ‘a mountain in itself, the road winds round its base. Here rock riots over rock, and mountain intersecting mountain, form one grand sweep of broken pointed crags’ (West 1778, p. 100). Forty years later, in 1798, Joseph Pocklington developed it as a tourist attraction. By 1807 as well as providing a cottage for the resident guides he had also erected a druid stone and constructed a small chapel together with an affixed ladder for visitors to clamber to its top; rock fragments around its base were also cleared. Such works were not universally admired; for example: a single rock called the Bowder Stone, a fragment of great size which has fallen from the heights. The game person who formerly disfigured the island in Keswick Lake with so many abominations, has been at work here also; has built a little mock hermitage, set up a new druidical stone, erected an ugly house for an old woman to live in who is to show the rock, for fear travellers should pass under it without seeing it, cleared away all the fragments round it, and as it rests upon a narrow base, like a ship upon its keel, dug a hole underneath through which the curious may gratify themselves by shaking hands with the old woman. The oddity of this amused us greatly, provoking as it was to meet with such hideous buildings in such a place, – for the place is as beautiful as eyes can behold,

or imagination conceive. (Espriella 1814, vol. 2, pp. 164– 165)

Green’s guidebook noted a further deterioration of its natural appeal with the guide presenting visitors with an ‘exordium preparatory to the presentation of a written paper, specifying the weight and dimensions of the stone’ (Green 1819, vol. 2, p. 134). The region’s mines, from the eighteenth century onwards, were popular with some travellers. For example, Defoe in his tenth letter of A tour thro’ the while island of Great Britain . . . mentions ‘Derwent Fells, where the ancient copper mines were found in Queen Elizabeth’s time . . .’ and notes that ‘Here are still mines of black lead found . . . the only place in Britain where it is to be had’. In Espriella’s account it was recorded that ‘Above this lies the mine of black-lead of which those pencils so famous over all Europe are made, it is the only one of the kind which has yet been discovered. We could not see it, as it is worked only occasionally, and had just been shut’ (Espriella 1814, vol. 2, 167). It was the poet Robert Southey who penned in 1808, under the pseudonym Don Manuel Alvarez Espriella, the three-volume Letters from England, with a third edition in 1814. Its second volume has various Lake District accounts and records that We had consulted tourists and topographers in London, that we might not overpass any thing worthy of notice, and our Guide to the Lakes was with us. They told us of tracts of horrible barrenness, of terrific precipices, rocks rioting upon rocks, and mountains tost together in chaotic confusion; of stone avalanches rendering the ways impassable, the fear of some travellers who had shrunk back from this dreadful entrance into Borrodale, and the heroism of others who had dared to penetrate into these impenetrable regions: into these regions, however, we found no difficulty in walking along a good road, which coaches of the light English make travel every summer’s day. (Espriella 1814, vol. 2, p. 162)

The last sentence is significant in that it refutes the accounts of perilous ventures into Borrowdale of only 50 years earlier by Gilpin and Gray. Southey (1774–1843), as well as being a Romantic poet, was a prolific man of letters and essayist, a biographer, literary scholar and historian – excepting the lack of geological training, a fair combination of abilities for a writer on geotourism!

The past is the key to the present In defining modern geotourism and promoting its research, the seminal UK work underpinned the first national geotourism conference – indeed the first anywhere – held at Belfast’s Ulster Museum

THREE CENTURIES OF APPRECIATING PHYSICAL LANDSCAPES

in 1998 (Robinson 1998). Few of its wholly unpublished presentations made any attempt to define geotourism and most, perhaps with the exception of Hose (1998), were focussed on its practice and provision. This approach was reiterated a decade later by the presentations at two ‘global’ geotourism conferences in 2008 and 2010. Nowadays hardly a year passes without a geotourism conference, global or otherwise, held somewhere. However, until the 2012 conference, from which this volume is a partial record, none had examined the historical basis of geotourism. Consequently, few of modern geotourism’s practitioners and stakeholders either have knowledge of or understand the parallels and lessons than can be drawn out from historical geotourism. The papers in this volume seek to at least partially help redress this deficiency in terms of both the specific and generic Baird, in this paper’s opening quote, was partly reiterating a point made in an article published in 1952 in Scientific Objectives. It was republished in 1970 in the Proceedings of the Geologists’ Association, it which it was noted that geology had two purposes ‘one material, the other spiritual; it provides the raw materials of this kind of civilization, and it satisfies certain spiritual aspirations in its followers. Many more ordinary people could take advantage of its spiritual gifts’ (Read 1970, p. 420); this is perhaps a partial successor to Nicholas Steno’s maxim (see Hansen 2009, p. 6) ‘Beautiful is what we see. More beautiful is what we understand. Far most beautiful is what we are ignorant about’. published in Prooemium or Preface to a Demonstration in Copenhagen Anatomical Theater in the Year 1673 as he contemplated his own religious devotion. Read confessed that he found ‘in geology a certain satisfaction of spiritual needs that is uplifting and cultural’ and suggested that ‘this may be only an exercise of the detective instinct that is in most of us . . . And we should not ignore the actual physical pleasure that arises from hard geological fieldwork. If we can combine the mental exaltation . . . with the strenuous physical exercise that should go with it, then we have the finest life in the world’ (Read 1970, p. 414). However, such spirituality is not confined to formalized religion, and there are many accounts of travellers being emotionally affected by natural attractions and landscapes; the psychology of this has been well reviewed by Kaplan & Kaplan (1989), who noted the concept of the ‘restorative environment’ that was especially found in aesthetically attractive natural environments such as scenery (that is landscapes). As has already been argued in this paper, the engagement in geological fieldwork in the eighteenth and nineteenth centuries by men of considerable social and economic standing, of similar backgrounds to those who enjoyed the classic

19

Grand Tour, indicated a fundamental shift in the prevailing social climate to permit venturing into wild places and engaging on their travels with persons of different social stature. Historical geotourism’s development also benefitted from another fundamental shift; this was in the way in which landscapes or scenery were perceived and then exploited for tourism purposes – especially in the later years of the Grand Tour. The recognition that rugged landscapes were worthy places to visit was essential. Curiosity and aesthetic value were motivators before scientific value for travels into such landscapes. These changes can be traced through the way in which writers described and artists visualized them. Indeed, these changes influence modern geotourism up to the 1970s and beyond. The Romantic legacy left to twentiethand twenty-first-century travellers and tourists was the established preference for participants to sojourn in attractive ‘wild’ or ‘natural’ landscapes rather than the ‘controlled’ agricultural landscapes and the ‘brutal’ locations of heavy industry and mining. Eighteenth- to nineteenth-century improvements in physical and intellectual access to encourage elite tourists to visit previously virtually inaccessible locations were developed in the twentieth century for mass tourists – a worthy field of future study, as modern geotourism (Hose 2011, 2012b), in its own right. For geotourism as a whole, unlike its parent discipline as James Hutton suggested in his ‘Principle of Uniformitarianism’, the past really is the key to the present. The invaluable research support, especially the online access to journals, provided by the University of Bristol is most gratefully acknowledged. The perceptive review by Randell Stephenson, for which grateful thanks are extended, has helped to hone this paper’s text and to improve the clarity of some its figures.

References Aldridge, D. 1975. Guide to Countryside Interpretation, Part One: Principle of Countryside Interpretation and Interpretive Planning. HMSO/Countryside Commission for Scotland/Countryside Commission, London. Baird, D. M. 1968. Geology in the public eye. In: Neale, E. R. W. (ed.) The Earth Sciences in Canada – A Centennial Appraisal and Forecast. Toronto University Press, Toronto, Royal Society of Canada Special Publications, 11, 222– 230. Baird, J. C. 1994. Naked rock and the fear of exposure. In: O’halloran, D., Green, C., Harley, M., Stanley, M. & Knill, J. (eds) Geological and Landscape Conservation. Geological Society, London, 335– 336. Barton, H. A. 1998. Northern Arcadia: Foreign Travelers in Scandinavia, 1765–1815. Southern Illinois University Press, Carbondale.

20

T. A. HOSE

Beazley, E. 1971. The Countryside on View: A Handbook on Countryside Centres, Field Museums, and Historic Buildings Open to the Public: Their Planning & Some Display Techniques. Constable, London. Bedell, R. 2001. The Anatomy of Nature: Geology and American Landscape Painting, 1825–1875. Princeton University Press, Princeton. Black, J. 1993. Italy and the Grand Tour. Yale University Press, New Haven, CT. Black, J. 2003. The British Abroad: The Grand Tour in the Eighteenth Century, 2nd edn. History Press, Stroud. Bristow, C. 2015. The role of Carclaze tin mine in eighteenth and nineteenth century geotourism. In: Hose, T. A. (ed.) Appreciating Physical Landscapes: Three Hundred Years of Geotourism. Geological Society, London, Special Publications, 417. First published online February 11, 2015, http://doi.org/10.1144/ SP417.11 Brown, K. M. 2014. Leave only footprints? How traces of movement shape the appropriation of space. Cultural Geographies, 1 –29. http://doi.org/10.1177/ 1474474014558987 Burek, C. V. & Prosser, C. D. (eds) 2008. The History of Geoconservation. Geological Society, London, Special Publications, 300. Burek, C. V. & Hose, T. A. 2015. The role of local societies in early modern geotourism: a case study of the Chester Society of Natural Science and the Woolhope Naturalists’ Field Club. In: Hose, T. A. (ed.) Appreciating Physical Landscapes: Three Hundred Years of Geotourism. Geological Society, London, Special Publications, 417. First published online February 6, 2015, http://doi.org/10.1144/SP417.6 Buzzard, J. 1993. The Beaten Track: European Tourism, Literature, and the Ways to ‘Culture’, 1800– 1918. Oxford University Press, Oxford. Cayla, N., Gauchon, C. & Hoble´a, F. 2015. From tourism to geotourism: a few historical cases from the French Alpine foreland. In: Hose, T. A. (ed.) Appreciating Physical Landscapes: Three Hundred Years of Geotourism. Geological Society, London, Special Publications, 417. First published online February 11, 2015, http://doi.org/10.1144/SP417.10 Chard, C. 1999. Pleasure and Guilt on the Grand Tour: Travel Writing and Imaginative Geography, 1600– 1830. Manchester University Press, Manchester. Constantine, D. 2001. Fields of Fire: A Life of Sir William Hamilton. Wiedenfeld & Nicholson, London. Cope, M. A. 2014. Three centuries of open access to the caves in Stoney Middleton Dale Site of Special Scientific Interest, Derbyshire. In: Hose, T. A. (ed.) Appreciating Physical Landscapes: Three Hundred Years of Geotourism. Geological Society, London, Special Publications, 417. First published online October 30, 2014, http://doi.org/10.1144/SP417.3 Dann, G. 1999. Writing out the tourist in space and time. Annals of Tourism Research, 26, 159–187. De Bastion, R. D. 1994. The private sector – threat or opportunity? In: O’halloran, D., Green, C., Harley, M., Stanley, M. & Knill, J. (eds) Geological and Landscape Conservation. Geological Society, London, 391– 395. Dolan, B. 2001. Ladies of the Grand Tour. HarperCollins, London.

Dowling, R. K. 2011. Geotourism’s global growth. Geoheritage, 3, 1 –13. http://doi.org/10.1007/s12371010-0024-7 Dowling, R. K. 2013. Global geotourism – an emerging form of sustainable tourism. Czech Journal of Tourism, 2, 59–79. Dowling, R. K. & Newsome, D. (eds) 2008. Geotourism. Elsevier, London. Edwards, J. A. & Llurdes, J. C. 1996. Mines and quarries industrial heritage tourism. Annals of Tourism Research, 23, 341– 363. Ellis, N. V. (ed.) 1996. An Introduction to the Geological Conservation Review. Joint Nature Conservation Commission, Peterborough. Erfurt-Cooper, P. & Cooper, M. 2010. Volcano and Geothermal Tourism: Sustainable Geo-Resources for Leisure and Recreation. Earthscan, London. Espriella, D. M. A. [SOUTHEY, R.] 1814. Letters from England (3 volumes). Longman, Hurst, Rees, Orme, and Brown, London. Faith, N. 1994. The World the Railways Made. Pimlico, London. Fairbrother, N. 1970. New Lives, New Landscapes. Architectural, London. Frey, M.-L., Schafer, K., Buchel, G. & Patzak, M. 2008. Geoparks – a regional European and global policy. In: Dowling, R. K. & Newsome, D. (eds) Geotourism. Elsevier, London, 95–117. Gillett, P. 1990. Art Publics in Late Victorian England. In: Gillet, P. (ed.) The Victorian Painter’s World. Sutton, Gloucester, 192–241. Gilpin, W. 1786. Observations, Relative Chiefly to Picturesque Beauty, Made in the Year 1722, on Several Parts of England: Particularly the Mountains and Lakes of Cumberland and Westmorland (in 2 volumes). London. Gordon, J. E. & Baker, M. 2015. Appreciating geology and the physical landscape in Scotland: from tourism of awe to experiential re-engagement. In: Hose, T. A. (ed.) Appreciating Physical Landscapes: Three Hundred Years of Geotourism. Geological Society, London, Special Publications, 417. First published online January 12, 2015, http://doi.org/10. 1144/SP417.1 Gray, M. 2008. Geodiversity: developing the paradigm. Proceedings of the Geologists’ Association, 119, 287–298. Green, W. 1819. The Tourist’s New Guide, Containing a Description of the Lakes, Mountains, and Scenery in Cumberland, Westmorland, and Lancashire, with Some Account of Their Bordering Towns and Villages. Being the Result of Observations Made During a Residence of Eighteen Years in Ambleside and Keswick (2 volumes). Lough & Co., Kendal. Hall, C. M. & Page, S. J. 2008. Progress in tourism management: from the geography of tourism to geographies of tourism – a review. Tourism Management, 30, 1 –14. Hall, C. M. & Weiler, B. 1992. What’s special about special interest tourism? In: Weiler, B. & Hall, C. M. (eds) Special Interest Tourism. Belhaven, London, 1– 14. Hansen, J. M. 2009. On the origin of natural history: Steno’s modern, but forgotten philosophy of science.

THREE CENTURIES OF APPRECIATING PHYSICAL LANDSCAPES Bulletin of the Geological Society of Denmark, 57, 1–24. Hebron, S. 2006. The Romantics and the British Landscape. British Library, London. Henry, C. J. & Hose, T. A. 2015. The contribution of maps to appreciating physical landscape: examples from Derbyshire’s Peak District. In: Hose, T. A. (ed.) Appreciating Physical Landscapes: Three Hundred Years of Geotourism. Geological Society, London, Special Publications, 417. First published online September 7, 2015, http://doi.org/10.1144/SP417.13 Hibbert, C. 1969. The Grand Tour. Weidenfeld & Nicolson, London. Hindley, G. 1983. Tourists, Travellers and Pilgrims. Hutchinson, London. Hose, T. A. 1995a. Evaluating interpretation at Hunstanton. Earth Heritage, 4, 20. Hose, T. A. 1995b. Selling the Story of Britain’s Stone. Environmental Interpretation, 10, 16– 17. Hose, T. A. 1996. Geotourism, or can tourists become causal rockhounds? In: Bennett, M. R., Doyle, P., Larwood, J. G. & Prosser, C. D. (eds) Geology on Your Doorstep: The Role of Urban Geology in Earth Heritage Conservation. Geological Society, London, 208–228. Hose, T. A. 1997. Geotourism – selling the Earth to Europe. In: Marinos, P. G., Koukis, G. C., Tsiamamaos, G. C. & Stournass, G. C. (eds) Engineering Geology and the Environment. Balkema, Rotterdam, 2955–2960. Hose, T. A. 1998. Towards a history of Geotourism: definitions, antecedents and the future. In: Burek, C. V. & Prosser, C. D. (eds) The History of Geoconservation. Geological Society, London, Special Publications, 30, 37–60. http://doi.org/10.1144/ SP300.5 Hose, T. A. 2000. European geotourism – geological interpretation and geoconservation promotion for tourists. In: Barretino, D., Wimbledon, W. P. & Gallego, E. (eds) Geological Heritage: Its Conservation and Management. Instituto Tecnologico Geominero de Espana, Madrid, 127–146. Hose, T. A. 2003. Geotourism in England: a two-region case study analysis. PhD thesis, University of Birmingham. Hose, T. A. 2005. Geotourism appreciating the deep time of landscapes. In: Novelli, M. (ed.) Niche Tourism: Contemporary Issues, Trends and Cases. Elsevier, Oxford, 27–37. Hose, T. A. 2006. Leading the field: a contextual analysis of the filed excursion and the filed-guide in England. In: Wickens, E., Hose, T. A. & Humberstone, B. (eds) Critical Issues in Leisure and Tourism Education: Current Trends and Developments in Pedagogy and Research. Leisure and Tourism Education Research Centre, Buckinghamshire Chilterns University College, High Wycombe, 115– 132. Hose, T. A. 2008. Towards a history of geotourism: definitions, antecedents and the future. In: Burek, C. V. & Prosser, C. (eds) The History of Geoconservation. Geological Society, London, Special Publications, 300, 37–60, http://doi.org/10.1144/SP300.5 Hose, T. A. 2010a. Volcanic geotourism in West Coast Scotland. In: Erfurt-Cooper, P. & Cooper, M. (eds) Volcano and Geothermal Tourism: Sustainable

21

Geo-Resources for Leisure and Recreation. Earthscan, London, 259–271. Hose, T. A. 2010b. The significance of aesthetic landscape appreciation to modern geotourism provision. In: Newsome, D. & Dowling, R. K. (eds) Geotourism: The Tourism of Geology and Landscapes. Goodfellow, Oxford, 13– 25. Hose, T. A. 2011. The English origins of geotourism (as a vehicle for geoconservation) and their relevance to current studies. Acta geographica Slovenica, 51, 343– 360. Hose, T. A. 2012a. Appreciating Physical Landscapes: Geotourism 1670– 1970: Abstracts Book. History of Geology Group/Geological Society, London. Hose, T. A. 2012b. 3G’s for modern geotourism. Geoheritage, 4, 7– 24. Hose, T. A. (ed.) In press. Geoheritage and Geotourism: A European Perspective. Boydell & Brewer, Woodbridge. Hose, T. A. & Vasiljevic´, D. A. 2012. Defining the nature and purpose of modern geotourism with particular reference to the United Kingdom and South-East Europe. Geoheritage, 4, 25– 43. Hudson, B. J. 2015. Waterfalls and the Romantic traveller. In: Hose, T. A. (ed.) Appreciating Physical Landscapes: Three Hundred Years of Geotourism. Geological Society, London, Special Publications, 417. First published online January 12, 2015, http:// doi.org/10.1144/SP417.9 Hutchinson, W. 1774. An Excursion to the Lakes, in Westmoreland and Cumberland, August 1773. J. Wilkie and W. Goldsmith, London. Kaplan, R. & Kaplan, S. 1989. The Experience of Nature: A Psychological Perspective. Cambridge University Press, Cambridge. Knudson, D. M., Cable, T. T. & Beck, L. 1995. Interpretation of Cultural and Natural Resources. Venture, State College, PA. Larwood, J. G. 2014. Geotourism: an early photographic insight through the lens of the Geologists’ Association. In: Hose, T. A. (ed.) Appreciating Physical Landscapes: Three Hundred Years of Geotourism. Geological Society, London, Special Publications, 417. First published online December 18, 2014, http://doi.org/ 10.1144/SP417.7 Larwood, J. & Prosser, C. 1998. Geotourism, conservation and tourism. Geologica Balcanica, 28, 97–100. Leafe, R. 1998. Conserving our coastal heritage– a conflict resolved. In: Hooke, J. (ed.) Coastal Defence and Earth Science Conservation. Geological Society, London, 10–19. Nabholtz, J. R. 1964. Wordsworth’s ‘Guide to the Lakes’ and the picturesque. Modern Philology, 61, 288– 297. Marples, M. 1959. Shank’s Pony. Dent, London. Martini, G. 1994. The protection of geological heritage and economic development: the saga of the Digne ammonite slab in Japan. In: O’halloran, D., Green, C., Harley, M., Stanley, M. & Knill, S. (eds) Geological and Landscape Conservation. Geological Society, London, 383–386. Mather, J. D. 2014. Geology and landscape in SW England in the late eighteenth century, as recorded

22

T. A. HOSE

in the travel journals of William George Maton (1774– 1840). In: Hose, T. A. (ed.) Appreciating Physical Landscapes: Three Hundred Years of Geotourism. Geological Society, London, Special Publications, 417. First published online November 13, 2014, http://doi.org/10.1144/SP417.8 Migon´, P. 2014. Rediscovering geoheritage, reinventing geotourism: 200 years of experience from the Sudetes, Central Europe. In: Hose, T. A. (ed.) Appreciating Physical Landscapes: Three Hundred Years of Geotourism. Geological Society, London, Special Publications, 417. First published online November 13, 2014, http://doi.org/10.1144/SP417.2 Morris, C. (ed.) 1949. The Journeys of Celia Fiennes. Cresset, London. Morritt, J. B. S. 1985. A Grand Tour: Letters and Journeys 1794–96. Century, London. Newsome, D. & Dowling, R. K. 2010. Geotourism: The Tourism of Geology and Landscape. Goodfellow, Oxford. Novelli, M. (ed.) 2005. Niche Tourism Contemporary Issues, Trends and Cases. Elsevier, Oxford. Nugent, J. 1749. The Grand Tour. Containing an Exact Description Of most of the Cities, Towns, and Remarkable Places of Europe. Together with a Distinct Account of the Post-Roads and Stages, with their respective Distances, Through Holland, Flanders, Germany, Denmark, Sweden, Russia, Poland, Italy, France, Spain, and Portugal. Likewise Directions relating to the Manner and Expence of Travelling from one Place and Country to another. As also Occasional Remarks on the Present State of Trade, as well as of the Liberal Arts and Sciences, in each respective Country. By Mr. Nugent. In Four Volumes. S. Birt, D. Browne, A. Millar, G. Hawkins, London. Nugent, J. 1768. Travels through Germany – With a particular account of the courts of Mecklenburg: In a series of letters to a friend; Embellished with elegant cuts. E. & C. Dilly, London. Oldroyd, D. R. 2002. Earth, Water, Ice and Fire: Two Hundred Years of Geological Research in the English Lake District. Geological Society, London, Memoirs, 25. Page, K. N. 1998. England’s Earth Heritage Resource – an asset for everyone. In: Hooke, J. (ed.) Coastal Defence and Earth Science Conservation. Geological Society, London, 196– 209. Parsons, N. T. 2007. Worth the Detour: A History of the Guidebook. Sutton, Stroud. Patzak, M. & Eder, W. 1998. “UNESCO GEOPARK”. A new programme – a new UNESCO label. Geologica Balcania, 28, 33–35. Plog, S. C. 1974. Why destination areas rise and fall in popularity. Cornell Hotel and Restaurant Administration Quarterly, 15, 55–58. Pointon, M. 1978. Geology and landscape painting in nineteenth century England. In: Jordanova, L. J. & Porter, S. R. (eds) Images of the Earth: Essay in the History of the Environmental Sciences (BSHS Monographs 1). British Society for the History of Science, Chalfont St. Giles, 84– 108. Porter, L. 1992. On Spartan Lines: Early Years of the YHA. Ashbourne Editions, Ashbourne.

Pullin, R. 2014. The artist as geotourist: Eugene von Gue´rard and the seminal sites of early volcanic research in Europe and Australia. In: Hose, T. A. (ed.) Appreciating Physical Landscapes: Three Hundred Years of Geotourism. Geological Society, London, Special Publications, 417. First published online December 15, 2014, http://doi.org/10.1144/SP417.4 Read, H. H. 1970. The geologist as historian. Proceedings of the Geologists’ Association, 81, 409– 420. Read, S. E. 1980. A prime force in the expansion of tourism in the next decade: special interest travel. In: Hawkins, D. E., Shafer, D. E. & Rovelstad, J. M. (eds) Tourism Marketing and Management Issues. George Washington University Press, Washington, DC, 193– 202. Robinson, E. 1998. Tourism in geological landscapes. Geology, 14, 151–153. http://doi.org/10.1046/j. 1365-2451 Ruff, A. R. & Mellors, O. 1993. The mountain bike – the dream machine? Landscape Research, 18, 104–109. Schnapp, A. 1996. The Discovery of the Past: The Origins of Archaeology. British Museum Press, London. Sillitoe, A. 1995. Leading the Blind: A Century of Guidebook Travel 1815–1911. Macmillan, London. Slomka, T. & Kicinska-Swiderska, A. 2004. Geotourism–the basic concepts. Geoturystyka, 1, 5– 7. Spiteri, A. 1994. Malta: a model for the conservation of limestone regions. In: O’halloran, D., Green, C., Harley, M., Stanley, M. & Knill, J. (eds) Geological and Landscape Conservation. Geological Society, London, 205–208. Stoye, J. 1989. English Travellers Abroad 1604–1667, rev. edn. Yale University Press, New Haven, CT. Stueve, A. M., Cook, S. D. & Drew, D. 2002. The Geotourism Study: Phase 1 Executive Summary. National Geographic, Washington, DC. Sweet, R. 2012. Cities and the Grand Tour: The British in Italy, c. 1690–1820. Cambridge University Press, Cambridge. Taylor, J. 1888. Early Prose and Poetical Works of John Taylor, the Water Poet (1580–1653). Hamilton, Adams & Co. & Thomas D. Moeison, London & Glasgow. Towner, J. 1984. The Grand Tour Sources and a methodology for an historical study of tourism. Tourism Management, 5, 215– 222. Towner, J. 1985. The grand tour a key phase in the history of tourism. Annals of Tourism Research, 12, 297–333. Towner, J. 1996. An Historical Geography of Recreation and Tourism in the Western World 1540–1940. Wiley, London. Toynbee, P. & Whibley, L. (eds) 1971. Correspondence of Thomas Gray: Volume III, 1766– 1771. Revised 1935 edition with corrections by H. W. Starr. Clarendon, Oxford. UNESCO 2000. UNESCO Geoparks Programme Feasibility Study. UNESCO, Paris. van den Ancker, J. A. M. & Jungerius, P. D. In press. Landscape and Geotourism on the Dutch coast in the seventeenth century as depicted by landscape artists. In: Hose, T. A. (ed.) Appreciating Physical Landscapes: Three Hundred Years of Geotourism.

THREE CENTURIES OF APPRECIATING PHYSICAL LANDSCAPES Geological Society, London, Special Publications, 417, http://doi.org/10.1144/SP417.14 Vasiljevic´, D. A., Markovic´, S. B. & Vujicˇic´, M. D. 2014. Appreciating loess landscapes through history: the basis of modern loess geotourism in the Vojvodina region of North Serbia. In: Hose, T. A. (ed.) Appreciating Physical Landscapes: Three Hundred Years of Geotourism. Geological Society, London, Special Publications, 417. First published online November 3, 2014, http://doi.org/10.1144/ SP417.5 Vaughn, J. 1974. The English Guidebook c. 1780–1870: An Illustrated History. David & Charles, Newton Abbot. VICTORIA & ALBERT MUSEUM 1984. The Discovery of the Lake District: A Northern Arcadia and Its Uses. Victoria and Albert Museum, London. Watkins-Pitchford, D. 1959. The Autumn Road to the Isles. Nicholas Kaye, London. Weiler, B. & Hall, C. M. (eds) 1992. Special Interest Tourism. Belhaven, London. West, T. 1778. A Guide to the Lakes: Dedicated to the Lovers of Landscape Studies, and to All Who Have Visited, or Intend to Visit the Lakes in Cumberland, Westmorland and Lancashire. B. Law, etc, London. Whalley, W. B. & Parkinson, A. F. In press. Visitors to the Northern Playgrounds: tourists and exploratory science in north Norway. In: Hose, T. A. (ed.) Appreciating Physical Landscapes: Three Hundred Years

23

of Geotourism. Geological Society, London, Special Publications, 417. http://doi.org/10.1144/SP417.12 Wilson, C. (ed.) 1994. Earth Heritage Conservation. Geological Society/Open University, London. Wilton, A. & Bignamini, I. 1996. Grand Tour: Lure of Italy in the Eighteenth Century. Tate, London. Withey, L. 1998. Grand Tours and Cook’s Tours: A History of Leisure Travel 1750– 1916. Aurum, London. WORLD TOURISM ORGANIZATION 1997. What Tourism Managers Need to Know. A Practical Guide for the Development and Application of Indicators of Sustainable Tourism. Retrieved from http://www.worldtourism. org/cgi-bin/infoshop.storefront/EN/product/ 1020-1 Wordsworth, W. 1835. The River Duddon, A Series of Sonnets; Vaudracour & Julia: and Other Poems. To which is annexed, A Topographical Description Of the Country of the Lakes, In the North of England. Hudson and Nicholson, Kendal. Wyatt, J. 1995. Wordsworth and the Geologists. Cambridge University Press, Cambridge. Wyse Jackson, P. N. 2007. Four Centuries of Geological Travel: The Search for Knowledge on Foot, Bicycle, Sledge and Camel. Geological Society, London, Special Publication, 287. Young, A. 1770. A six months tour through the north of England: Containing, An Account of the prefent State of Agriculture, Manufactures and Population, in feveral Counties of this Kingdom (3 volumes). W. Strahan, London.

Appreciating geology and the physical landscape in Scotland: from tourism of awe to experiential re-engagement JOHN E. GORDON1* & MATT BAKER2 1

School of Geography and Geosciences, University of St Andrews, St Andrews, Fife KY16 9AL, Scotland, UK 2

The Solway Centre for Environment and Culture, University of Glasgow, Crichton University Campus, Dumfries DG1 4ZL, Scotland, UK *Corresponding author (e-mail: [email protected])

Abstract: This chapter explores people’s experience of the physical landscape in Scotland from the perspective of parallel developments in geological science, landscape aesthetics and tourism since the middle of the eighteenth century. It begins with tourism of awe, inspired by the Romantic movement and the excitement of discovering natural wonders promoted through contemporary literature and art during the development of modern geological science in the late eighteenth–early nineteenth centuries. Popular interest and engagement in geology declined with increasing scientific specialization, although the physical landscape continued to draw many visitors and provide creative inspiration for poets and artists. In the 1940s, the beginnings of statutory geoconservation were accompanied by renewed interest in raising public awareness of geology and the physical landscape mainly through didactic methods. More recently, exploration of the cultural links between geology and landscape is providing new opportunities for experiential re-engagement, a shift that recognizes the close links between people and the physical landscape, and one promoted through voluntary sector activity in geoconservation and the development of Geoparks. Rediscovering a sense of wonder and reconnecting with the landscape offer a means of reconciling the natural and cultural worlds and enabling wider public appreciation of geodiversity.

The physical landscape of Scotland has long been a source of wonder, inspiration and creativity founded on its remarkable geodiversity (Geikie 1905; Holloway & Errington 1978; Hunter 1995; Gairn 2008; Gordon 2012a). The strong connections between geology, landscape and culture have also formed a major underpinning of tourist appreciation of the Scottish landscape since the latter part of the eighteenth century, although in Gaelic culture a deep reverence for the natural world extends much further back in time (Hunter 1995). In this chapter we explore selected aspects of these connections viewed through different forms of cultural representation and interpretation. Our starting point is the visit in 1788 by James Hutton, John Playfair and Sir James Hall to Siccar Point on the Berwickshire coast (Fig. 1). Hutton’s revelation of deep time from his examination of the rocks there was not only a pivotal moment of enlightenment in geological understanding, but also represented the beginning of a fundamental shift in the cultural experience of geology and landscape: a shift away 1

from the appreciation of physical features in terms of mythology towards understanding influenced by Enlightenment science.1 As revealed in the record of the rocks, the Earth was in constant flux over vast timescales; ‘we find no vestige of a beginning, – no prospect of an end’ (Hutton 1788, p. 304). To our modern minds there is a poetic quality in Hutton’s words, but when one considers the reality of what was lost in that moment of revelation, there is also a certain terror in Playfair’s later confession that ‘[t]he mind seemed to grow giddy by looking so far into the abyss of time’ (Playfair 1805, p. 73). What Hutton found at Siccar Point was compelling evidence of an unconformity that clearly demonstrated that the surface of the Earth had been formed progressively by many different rock cycles separated widely in time. What was lost in that moment of revelation of a hidden order to the world recorded in the rocks was a long-held mythological understanding of the relationship between human beings and their physical surroundings; what was gained was the opening of a new

The Enlightenment, an intellectual movement that flourished during the seventeenth and eighteenth centuries, progressed the advancement of knowledge through scientific method and reason rather than tradition or religion. James Hutton was one of the key figures in the Scottish Enlightenment of the second half of the eighteenth century, alongside Adam Smith, David Hume, Joseph Black and others.

From: Hose, T. A. (ed.) 2016. Appreciating Physical Landscapes: Three Hundred Years of Geotourism. Geological Society, London, Special Publications, 417, 25– 40. First published online January 12, 2015, http://dx.doi.org/10.1144/SP417.1 # The Geological Society of London 2016. Publishing disclaimer: www.geolsoc.org.uk/pub_ethics

26

J. E. GORDON & M. BAKER

Fig. 1. Locations of sites mentioned in the text (Base map by Uwe Dedering from http://commons.wikimedia.org/ wiki/File%3AScotland_relief_location_map.jpg [CC-BY-SA-3.0 Unported], via Wikimedia Commons).

world of potential discovery and insight into the way the Earth works and the evolution of life. As the science of geology set about developing an

empirical understanding of the ‘abyss of time’ and reconstructing Earth history, artists, writers and poets wrestled with a different problem: building

APPRECIATING THE PHYSICAL LANDSCAPE IN SCOTLAND

new mythologies that situated humankind’s ‘sense of self’ in relation to the geological timescale. If we take a moment to acknowledge this threshold in the history of cultural relationships to landscape, we can begin to see that this enlightenment – while bringing huge new freedoms and possibilities – also left a void to fill, a ‘creative mythology of self’ in relation to the rocks (Baker & Gordon 2012). Against this background, we examine how appreciation of the physical landscape in Scotland since the middle of the eighteenth century is interwoven with aspects of cultural representation in geological science, landscape aesthetics, tourism, literature, art and geoconservation. In particular, we consider three overlapping phases involving tourism of awe, utilitarian exploitation and experiential re-engagement.

Tourism of awe The origins of tourism in Scotland in the latter part of the eighteenth century are closely linked to a fundamental shift in the appreciation of the Scottish landscape by outside visitors, arising from the development of landscape aesthetics and changing perceptions of nature and geology and how these were conveyed in contemporary literature, art and the journals of early exploratory travellers (Smout 1982; Butler 1985; Andrews 1987; Gold & Gold 1995; Glendening 1997; Durie 2003, 2012; Grenier 2005; Hose 2010a; Gordon 2012b). As travel became safer in post-Culloden times, and roads and knowledge of geography improved through the military campaigns and surveys of General George Wade and Major William Caulfield, the social and literary elite both in Scotland and in England (and later beyond) began to discover Scotland for scenic pleasure and recreation. PreRomantic responses of aversion and horror at the bleakness and sterility of the landscape (Defoe 1726; Johnson 1775), reinforced by the difficulties of travel and the limited and poor accommodation, were superseded by the appreciation of the picturesque and sublime qualities of the landscape influenced by the landscape aesthetics of William Gilpin, Edmund Burke, Richard Payne Knight, Uvedale Price and others (Heringman 2004; Dean 2007). The Romantic movement and the aesthetics of the sublime, associated with the expression of emotions of wonder, awe and pleasurable terror, 2

27

helped to change perceptions of the physical landscape and were essentially founded in its underlying geological and geomorphological history and characteristics (Dean 2007). Visitors were both fascinated by the apparent ruins of the Earth and ‘repelled by the horror of its chaos’ (MacLeod 2012, p. 140). For example, writing in 1766 about Glen Croe in Argyll west of Loch Lomond, Mrs Elizabeth Montagu, the London socialite, literary critic, social reformer and leading figure in the ‘bluestocking’ movement, described in a letter to her sister-in-law how: Mountain rose above mountain . . . . sometimes the rocks hung threatening over our heads, in some places the torrents had washd away the earth, & the bare skeleton of some mountain appeard, which like that of a Giant, look’d more terrible in unmitigated strength & unsoftend form than when compleat & perfect. The cheeks of the mountains were furrowd by the falling streams, & the grey moss of age grew upon them; here & there a firr tree shatterd by a thousand storms, & huge rocks that had roll’d half way down, & waited, another earthquake to compleat its journey, formd the terrible sublime’ (quoted in Ross (1965) from the original in the Huntington Library).

Romantic fascination with the Highlands was also inspired by the publication of James Macpherson’s Ossianic poetry2 in the 1760s (Macpherson 1760, 1765; Gaskill 1996), which had an appeal far beyond Scotland (Stafford 2005). Visitors came in increasing numbers in the late eighteenth and early nineteenth centuries in search of sublime landscapes and places associated with the poetry of Ossian and Fingalian legend (Smout 1982; Andrews 1987; Gold & Gold 1995). Particularly influential were the descriptions and illustrations of natural landscape wonders and their wild and romantic character portrayed for the reading public in the journals and guidebooks of early travellers, including Richard Pococke (1887), Thomas Pennant (1771, 1776), Samuel Johnson (1775), James Boswell (1785), William Gilpin (1789), Sarah Murray (1799, 1805), Dorothy and William Wordsworth (Wordsworth 1874) and geologist John Macculloch (1819a, b, 1824), who produced the first large-scale geological map of Scotland (published posthumously in 1836). Natural wonders such as caves, waterfalls and scenic views were appreciated through the aesthetics of the sublime (Pointon 1979). Many, including Staffa, the Falls of Clyde, the Falls of Foyers (Fig. 2), the Falls of Braan, Loch Lomond, The

In the 1760s, the Scottish poet James Macpherson published in an English translation what he claimed were epic Gaelic poems handed down from ancient sources and narrated and authored by Ossian, a third century Scottish bard and supposedly the son of Fingal, king of a legendary race of warrior giants. The poems, however, were largely written by Macpherson himself, based on material collected throughout the Highlands. His work was internationally translated and highly influential in the Romantic movement in Europe.

28

J. E. GORDON & M. BAKER

Fig. 2. ‘Lower Fall of Fyers’, an engraving by J. Storer from a picture by Alexander Nasmyth (from Storer & Greig 1805). The Falls of Foyers have been a popular tourist destination since the late eighteenth century. They were visited by Thomas Pennant, Dr Johnson, Robert Burns and the Wordsworths, among many others (By permission of University of Glasgow Library, Special Collections).

Trossachs and Loch Coruisk on the Island of Skye (Fig. 3), became essential Romantic tourist destinations recommended in the numerous contemporary guidebooks and journal accounts (Hose 2010a; Gordon 2012a). Their attraction, too, was increased by emotive writing. For example, the ‘majesty and strength of the water’ at the Falls of Clyde struck Dorothy Wordsworth ‘with astonishment’ (Wordsworth 1874), Sarah Murray (1805) experienced ‘sublime heaven-like sensations’ at Fingal’s Cave on Staffa (Fig. 4) and geologist Robert Jameson wrote

of the granite mountains of Arran that ‘[n]ature exhibits to the astonished eye the most terrific and sublime scenery’ (Jameson 1813, p. 17). Other publications also catered for the scientific tourist, including details of geology, geomorphology and mineralogy (Walford 1818). Visitors flocked to the most popular sites: Macculloch (1819a) complained on Staffa that ‘visitors . . . crowd this far-famed spot’, a view echoed by Keats and Wordsworth. He also bemoaned the number of visitors at Loch Katrine (Macculloch 1824), where John Knox’s

APPRECIATING THE PHYSICAL LANDSCAPE IN SCOTLAND

29

Fig. 3. ‘Loch Coruisq near Loch Scavig’ (1819), aquatint by William Daniell (1820). ‘The approach to this point, towards which the rock and perpendicular shores appear to converge, is awfully magnificent and sublime’ (Daniell 1820, p. 36) (By permission of University of Glasgow Library, Special Collections).

Fig. 4. ‘View of Staffa from the South West’ by John Macculloch (1819b), engraving by L. Byrne. This depiction by Macculloch and aquatints by Daniell (1818) represented more realistic interpretations of geological and landscape detail than earlier stylized and exaggerated drawings designed to excite awe at the natural wonder of Fingal’s Cave (see Klonk 1997) (By permission of University of Glasgow Library, Special Collections).

30

J. E. GORDON & M. BAKER

Landscape with Tourists at Loch Katrine (oil on canvas, c. 1820) illustrates the popularity of The Trossachs following the publication of Scott’s The Lady of the Lake (1810). Many of these natural features and landscapes also acquired historical and literary associations through the works of James Macpherson, Robert Burns and Sir Walter Scott. The romantic images portrayed in the poetry and novels of Scott, in particular, were highly influential in presenting the attractions of the Scottish landscape to a wider audience through popular literature in the early nineteenth century. They also inspired the Scottish ‘romantic tour’ (Andrews 1987; Gold & Gold 1995; Glendening 1997), which was effectively an early form of geotourism defined in a broad sense but linked with literary and historical associations. Scott successfully interweaved literature, art and history with the physical landscape. His publications stimulated a great expansion in visitor numbers, facilitated through enhanced accessibility by road and steamship. Scott’s engagement with the landscape was informed by an awareness of geology (Gordon 2012b). As a contemporary of Hutton, he was clearly aware of the latter’s ideas of deep time (cf. McPhee 1981) and the recycling of continents; for example in describing Ben Venue in The Trossachs in The Lady of the Lake (Scott 1810), he refers to topographic features representing the ‘fragments of an earlier world’. In parallel with developments in geological thinking, contemporary pictorial representations of particular landscape features in the accounts of early travellers enhanced their romantic appeal (Hose 2010a; Gordon 2012a; MacLeod 2012). With varying degrees of geological accuracy, these reflected changing perceptions from sublime creations of nature to the more picturesque. For example, images of Staffa and Fingal’s Cave (Klonk 1997) ranged from stylized architectural engravings of a cathedral-like cave and regimented columnar jointing to Turner’s atmospheric expression in Staffa, Fingal’s Cave (oil on canvas, 1832), and the more realistic geological interpretations by Daniell (1818) and Macculloch (1819b) (Fig. 4). In particular, the links between geology and landscape art were explicitly recognized by Macculloch: [i]t is thus that geology . . .. . .. illustrates even the pursuits of the artist. As far as landscape depends on forms, it will be found that it is very often essentially regulated, as to its beauty or deformity as well as its character, by the nature of the rocks of which a country consists. And this is often true, even where the rocks are not visible; as the character of the surface, the outlines of the hills, the forms of the shores, and many other circumstances, depend on the geological nature and disposition of the rocks beneath (Macculloch 1824, Vol. 2, p. 274).

Landscape or topographic sketching was also widely used as a documentary technique to illustrate early geological publications and formed part of the development of the visual language of geology evolving from descriptive accompaniment to being integral to theoretical interpretation as the science progressed (Rudwick 1976; Klonk 1996). The technically accurate sectional drawings by John Clerk of Eldin intended for the third volume of Hutton’s Theory of the Earth (Craig 1978) are particularly notable in introducing the time dimension, but Clerk also produced a much wider body of ‘picturesque’ landscape drawings and etchings engraved for various publications (Bertram 2012). At the same time, artists, poets and writers rediscovered a new awe in the ‘works’ of nature which they contrasted with the works of man. This ability to speculate about unknown but unquestionably awesome creative forces in nature presented a challenge to artists to redefine the scale of landscape in relation to mankind as something ‘other’, something that was literally terrifying to contemplate. Literature, art and poetry thus opened new worlds in landscape appreciation for a much wider audience. Invoking a time accessible only through the romantic imagination, poets and artists, as well as geologists, encouraged a sense of wonder about the natural forces that fashioned the landscape (MacLeod 2012). Thus many contemporary landscape painters, notably Paul Sandby, Jacob More, Alexander Nasmyth, Hugh William Williams, John Knox, John Thomson of Duddingston and Horatio McCulloch evolved a style that merged the classical influences of Claude and Poussin with the reality of the Scottish landscape (Macdonald 2000) and depicted spectacular Highland landscapes and natural features that appealed to the aesthetics of the picturesque and sublime (Holloway & Errington 1978; Campbell 1993; Macmillan 2000; Morrison 2003). Indeed, for geological content, the influential critic John Ruskin (1844) admired Turner’s Loch Coruisk, Skye (watercolour, 1831–34) more than contemporary geological drawings, although Miller (1892) was not convinced about the reality of representation of geology in Turner’s work, which he described as ‘a kind of transcendental or transfigured geology’ (p. 150). Turner’s imaginative response to the physical landscape is nevertheless grounded in a degree of truthfulness to the geology, but far transcends simple topographical depiction. Wider appreciation of the Scottish landscape was also enabled through the publication of collections of landscape engravings and prints, including The Virtuosi’s Museum (Sandby 1778–1781), Views in North Britain Illustrative of the Works of Robert Burns (Storer & Greig 1805), Scenery of the Grampian Mountains (Robson 1814) and the Scottish

APPRECIATING THE PHYSICAL LANDSCAPE IN SCOTLAND

volumes of A Voyage Round Great Britain (Daniell 1818–1822).

Utilitarian exploitation Following the Industrial Revolution and into the Victorian era, geology became one of the ultimate contexts against which human achievements and technological ingenuity in overcoming the environment through scientific discovery and civil engineering were measured. Against a background of advances in understanding geological processes, the immensity of geological time and the evolution of past life from the fossil record, a romantic age of Victorian science offered a human myth of conquering the Earth by overcoming geology, exemplified by the development of science fiction such as Jules Verne’s Journey to the Centre of the Earth (first French edition 1864). The inconceivably long timescales and intractable materials of the Earth became the obstacles for humanity to overcome. This was reflected not only in great advances in the use of geological materials in industry and for energy generation, but also in great progress in transport through the construction of roads, canals and railways across challenging physical landscapes and the invention of steamships. These developments greatly facilitated mobility at an affordable cost for greater numbers of people, and were accompanied by a huge change in the nature and scale of tourism and appreciation of the potential use of the landscape and its natural features for new forms of recreation and enjoyment. Tourism expanded through the rise of commercial companies, including railway and steamship operators offering tours utilizing the new forms of transport. Prominent among these were David MacBrayne’s summer tours on the Royal Mail Steamers and Thomas Cook’s tours to Scotland that began in 1846. Typical tours included Loch Lomond, The Trossachs and Staffa, involving itineraries essentially founded on the attractions of the physical landscape and locations associated with Scott’s novels and poetry and following in the footsteps of earlier travellers, literary figures and artists. The proliferation of popular guidebooks (e.g. Lumsden & Son 1839; Black 1847; Cook 1866; Murray 1867; MacBrayne 1886) also helped to promote the appeal of the Scottish landscape, as did the publication of further illustrated books on the Scottish landscape (e.g. Beattie 1838; Hill et al. 1840) facilitated by advances in printing methods (MacLeod 2012). Notable among the former is George and Peter Anderson’s Guide to the Highlands and Islands of Scotland which included accounts of local geology and geomorphology (Anderson & Anderson 1851). The development of photography and promotion

31

of the scenic view through the early landscape photography of John Muir Wood, James Valentine and George Washington Wilson and later that of Robert Adam, together with the postcard industry, also did much to promote appreciation of the physical landscape. Queen Victoria’s tours of the Highlands and Prince Albert’s acquisition of Balmoral Castle near Braemar on Deeside in 1852 gave additional prestige to the region’s attractions. Romanticized Victorian stereotypes of ‘wilderness’ were epitomized in landscape and wildlife paintings by Edwin Landseer, Carl Haag, James Giles and others, overlooking the fact that the Highland clearances enabled the creation of sporting estates for the privileged pleasure of the rich and influential (Smout 1982). Landseer stayed in Glen Feshie in the Cairngorm Mountains in the company of the Duchess of Bedford, where he depicted the romantic moods of the landscape but also ventured deeper into the mountains. His painting Loch Avon and the Cairngorm Mountains (oil on panel, c. 1833), looking along the loch towards the great bulk of the Shelter Stone Crag with gathering storm clouds behind, is not only a ‘sublime’ expression of Highland scenery but captures the geomorphologically classic landscape of selective glacial erosion (see Sugden 1968). In this respect, it met a prerequisite of Ruskin who, in Modern Painters (Ruskin 1844), emphasized the importance of geological truth in landscape art. This required an understanding of the underlying rocks and geological structure of the landscape. Miller (1892), too, emphasized the need for geological awareness in landscape art. The Highlands offered an experience of wild and sublime landscapes not available in the ‘domesticated’ landscapes of England (with the exception of the central Lake District). However, the mechanistic explanations of geological science of the forces that had shaped the landscape became a form of ‘intellectual mastery’ that devalued the mystery and myths perpetuated in the literary texts, guidebooks and visual representations that had all helped to define the landscape for visitors in terms of the feelings they were expected to experience (Grenier 2005). Industrial development also began to intrude, and in the late nineteenth century tourists complained that the spectacle of the Falls of Foyers was ruined by the abstraction of water for the aluminium works on the shore of Loch Ness. Similarly, the Falls of Clyde were much reduced following the construction in 1927 of the first large-scale hydroelectric power station in Britain, leading to a drop in its popularity (Gordon 2012a). The development of new forms of recreation and the sporting attractions of shooting, fishing, mountaineering, golf and the seaside and health spas (Durie 2003; Grenier 2005) were all based on the assets provided by Scotland’s geodiversity and

32

J. E. GORDON & M. BAKER

involved a more utilitarian approach to the landscape, although most were only accessible to a social elite until after the First World War. Members of the Scottish Mountaineering Club founded in 1889 included eminent geologists who wrote articles in the Club’s publications on the links between geology and Scottish mountain landscapes (Geikie 1896; Harker 1900; Peach & Horne 1921). Archibald Geikie’s popular Scenery of Scotland (3rd edition 1901) included what may be the first detailed geotourism itinerary in Scotland. It was extensively illustrated by accurate landscape sketches, as was Andrew Ramsay’s influential text The Physical Geology and Geography of Great Britain (6th edition 1894). Landscape sketching was an essential geological skill, with art and geological appreciation of the landscape merging in the watercolours of Ben Peach and Henry Cadell. In Peach’s many sketches and paintings in his field notebooks and on the back of his field maps, ‘aesthetic sense is combined with geological insight’ (Anderson 1980). Ramsay, too, in his earlier account of the geology of Arran appreciated the aesthetic beauty of the landscape: Having reached the highest point of Goatfell, the eye of the geologist suddenly rests on a scene, which, if he be a true lover of nature, cannot fail to inspire him with astonishment and delight. The jagged and spiry peaks of the surrounding mountains, – the dark hollows and deep shady corries into which the rays of the sun scarce ever penetrate, – the open swelling hills beyond, – the winding shores of Loch-fyne, and the broad Firth of Clyde studded with its peaceful and fertile islands, – the rugged mountains of Argyllshire, and the gentle curves of the hills of the Western Isles, their outlines softened in the distance, form a scene of most surpassing grandeur and loveliness (Ramsay 1841, p. 7).

At the same time, infrastructure developments offered many new opportunities for the study of geological sections through canal, railway and dry dock cuttings, as did the pits and quarries opened to provide sources of aggregate and building stone. Widely reported in the Transactions of the Edinburgh and Glasgow Geological Societies and elsewhere (e.g. Craig 1874; Crosskey 1874; Braithwaite 2011), many of these studies represent important historical accounts of temporary exposures and were sometimes published by amateurs, their curiosity and interest inspired by the publications of Charles Lyell, Hugh Miller, Andrew Ramsay and Archibald and James Geikie. The utilitarian era also saw the first developments in geoconservation in Scotland, with moves to protect Salisbury Crags in Edinburgh (1845), Fossil Grove in Glasgow (1887) and Agassiz Rock (1908) in Edinburgh (Thomas & Warren 2008), all of which involved an element of raising wider

interest in the landscape and its geological and geomorphological features. In the first systematic geoconservation audit and assessment in 1871, the Royal Society of Edinburgh established the Boulder Committee to survey glacial erratics in Scotland that merited conservation (Milne Home 1872, 1884).

Experiential re-engagement: search for new unconformities Interest in Romantic landscape aesthetics and early tourism occurred at a time of major developments in the science of geology, with many interactions and influences. Hutton’s insights, as interpreted by Playfair (1802) and developed by Lyell (1830– 1833) in his widely read Principles of Geology, were fundamental in the growth of modern geoscience, but discoveries in geology held wider appeal and seized the public imagination. For example, the announcement of the existence of the Ice Age in Scotland in The Scotsman newspaper in 1840 was described by Ascherson (2002, p. 31) as scooping ‘the biggest story on earth’. In eighteenth and nineteenth century Britain, poets and artists also acknowledged the ‘new’ geology to be a powerful source of inspiration (Wyatt 1995; Nicolson 1997; Heringman 2004; Dean 2007; O’Connor 2007; Hose 2008). Geology offered a different way of looking at the landscape and allowed imaginative insights into past worlds in deep time (Rudwick 1992, 2008; Dean 2007; O’Connor 2007), while literature and poetry revealed the ‘inner meaning’ of the landscape (Geikie 1905). Science writing was an integral part of eighteenth and nineteenth century literature, and authors of works on popular geology (e.g. Charles Lyell, Archibald Geikie and Hugh Miller) made effective use of literary techniques (O’Connor 2007; Buckland 2013). In particular, Miller was an exceptional communicator through literary geology (Knell & Taylor 2006; Taylor 2007). Developing Hutton’s vision of geology as a subject of wider public interest (Hutton 1785), he captured the public imagination with tales of his encounters with rocks and fossils through a style of popular writing that also included accounts of local folklore and scenery (e.g. Miller 1858). His son later described the study of geology ‘as an exercise for the imagination’ (Miller 1892, p. 152). When viewed through a cultural filter (Hose 2010b), people engaged with geology and geomorphology through appreciation of the landscape, literature, poetry, art and tourism in an experiential way during the late eighteenth and nineteenth centuries. Natural features, places and past events inspired a sense of wonder conveyed through these media. However, the development

APPRECIATING THE PHYSICAL LANDSCAPE IN SCOTLAND

over time of an increasingly scientific and professional discipline (Heringman 2004; Rudwick 2005, 2008), initiated by the formation of the Geological Society of London in 1807 and its adoption of an empirical approach as a reaction against the conjectural theories about the Earth prevalent at the time, ultimately diminished the Romantic view of geology in the public imagination after the end of the Victorian period. In the twentieth century, re-engaging with the public has been closely allied to the development of modern geoconservation since the 1940s following the passing of the National Parks and Access to the Countryside Act (1949) (Prosser 2013) and successor legislation that has led to the designation of networks of Sites of Special Scientific Interest (SSSIs), latterly underpinned by the Geological Conservation Review (Ellis 2011). Alongside site protection and management, education and interpretation have become pillars of modern geoconservation as part of a more holistic approach that recognizes the integrity of nature, landscape and people (Scottish Geodiversity Forum 2013). Founded on reading the landscape (Gordon et al. 2004), the conventional approach to interpretation has essentially been a didactic one: geologists telling their stories and presenting their way of understanding the landscape (Gordon & Kirkbride 2009). Conversely, artists, poets and writers continued to draw powerful creative inspiration from the physical landscape and geological metaphors (Gordon 2012b). New, modernist approaches in landscape art included the experimental work of the Colourists who returned frequently to the island of Iona, the analytical approach of D. Y. Cameron, the abstract watercolours of W. G. Gillies and William Johnstone, the seascapes of Joan Eardley and the landscapes of James Morrison (Macmillan 2000). In poetry, a modernist, radical approach appeared in the work of Hugh MacDiarmid, Sorley McLean, Norman MacCaig and Ian Crichton Smith, for example in MacDiarmid’s On a Raised Beach (1934), MacLean’s An Cuilithionn (1939), MacCaig’s A Man in Assynt (1969) and Crichton Smith’s The White Air of March (1972), among other works. A strong connection with landscape and place is also present in the writings of Neil Gunn, George Mackay Brown, Lewis Grassic Gibbon and Nan Shepherd, and in the geopoetics of Kenneth White (Gairn 2008). In the last decade however, there has been convergence through the development of more experiential approaches in geointerpretation embracing: the cultural dimension of geodiversity; better understanding and application of best-practice principles in interpretation (Tilden 1977; Ham 2007); new developments in landscape art; and an awareness of the need for greater reconnection

33

with nature and the land through direct experience rather than remote scientific study (McIntosh 2001; White 2003; Macfarlane 2013). Geotourism has also emerged as an important driver for innovation in geointerpretation, both in a narrow sense (Hose 2012) as well as in a broader sense through recognition of geoheritage as a significant asset for tourism, particularly in Geoparks with their agenda to engage with a wide and varied nonspecialist audience (McKeever et al. 2010). Conservation has also broadened from a narrow science-based activity to a more inclusive exercise with much greater voluntary sector involvement (Whiteley & Browne 2013), recognizing the need to gain wider public and political support as well as acknowledging the tourism benefits from geoconservation. More experiential approaches to public engagement offer a means of provoking interest in geoheritage and rediscovering a sense of wonder about the landscape and its history (Dixon et al. 2012; Gordon 2012a, b). This is based on the premise that if people understand and have deeper connections with geoheritage, they are more likely to value it and help to manage it sustainably (Tilden 1977). Developing such connections forms part of the Geoparks’ rationale to promote the conservation of geoheritage, as well as sustainable economic and social development, through understanding and experience linking geology, natural heritage and cultural heritage. For example, Geopark Shetland has developed innovative interpretation integrating the islands’ geological and cultural history based around trails and on-site panels, digital tools and novel exhibits and installations designed to stimulate people’s interest rather than simply present information (http://www.shetlandamenity.org/ geopark-shetland) (Fig. 5), along with creative experiential engagement with local schools. Such ‘multi-sensuous engagement’ (Dixon et al. 2012) through exploring the cultural dimension can: † provide creative opportunities for people to connect with and appreciate geodiversity through different cultural experiences and a renewed sense of wonder; † link geodiversity to cultural roots and sense of place, allowing exploration of deep connections between people and the natural world; and † enable a more holistic view of the world. Curiosity about the built landscape provides opportunities for innovative interpretation such as ‘naming stones’ and the geological wall at the Scottish Parliament (Lothian & Borders Geoconservation 2011; Gordon 2012b). From a more practical perspective of living in a dynamic world, appreciation of the physical landscape and an awareness and understanding of geodiversity and natural

34

J. E. GORDON & M. BAKER

Fig. 5. At the Old Haa museum in Burravoe on the island of Yell, Geopark Shetland have installed a series of stone figures built from different types of rock in the style of Inuit ‘inuksuit’. In the same way that the Inuit use inuksuit as waymarkers and navigation aids, the Shetland figures help visitors to ‘navigate’ the island’s geology. The Inuit theme was chosen because most of Yell’s rocks were formed about 900 million years ago when Yell was part of what is now North America. Audio commentaries about the rock types represented were written and recorded by the pupils of Burravoe School. Further museum displays highlight Yell’s cultural links with North America through whaling and the Hudson’s Bay Company. The photograph shows the cover of the accompanying booklet. (Photograph: Shetland Amenity Trust.)

processes are fundamental to public debate about the risk from geohazards and the difficult decisions that will need to be made concerning managing adaptation to climate change (Prosser et al. 2010; Lane et al. 2011). In the debatable era of the Anthropocene, where for the first time there is discussion of geology as

something that humans are really influencing (Crutzen 2002; Zalasiewicz et al. 2010), the creative response to humanity’s ‘conquest’ is more nuanced than the celebration of achievement of an earlier age. One of the strategies employed by artists in this modern era is to creatively re-appropriate the scientific processes of geology as a new

APPRECIATING THE PHYSICAL LANDSCAPE IN SCOTLAND

mythology, inviting contemporary humans to re-place themselves in the context of immeasurable time signatures (i.e. to question the popular belief that human technology has the capacity to prolong itself into a geological timescale). These new ‘unconformities’ placed deliberately and unceremoniously within the public landscape reinterpret the language of science to reassert our understanding of ourselves as part of a living system (cf. Hutton 1788). This is reflected in the work of environmental artists such as Andy Goldsworthy, Charles Jencks and Matt Baker. Examples include Goldsworthy’s ‘Striding Arches’ at Cairnhead in Dumfries and Galloway, Jencks’ ‘Landform’ at the Scottish National Gallery of Modern Art in Edinburgh and Baker’s installations at Cairnsmore of Fleet National Nature Reserve in Dumfries and Galloway and St Mary’s Loch in the Scottish Borders (Fig. 6). These artworks are placed to be part of the landscape and to site the viewer in an enhanced relationship to the environment, rather than as objects in a landscape. The artistic conception of

35

these works is to be wholly perceived as a means of physical experiencing, interacting and understanding landscape rather than through their intellectual connection to an abstracted art history.

Conclusion The geological history of Scotland forms the foundation of the scenery that visitors have actively sought out for different reasons from the early exploratory travellers in the late eighteenth and early nineteenth centuries, to the hordes of Victorian tourists seeking to be awed and inspired by the mountains, glens and lochs and, more recently, those seeking various forms of outdoor recreation and geotourism experiences. A remarkable number of artists and writers have also found inspiration from the physical landscape and helped to raise awareness and promote the popularity of the Scottish landscape. Geotourism in its broadest sense therefore has a long tradition in Scotland.

Fig. 6. ‘Shinglehook’, 2006, by Matt Baker, is a ‘permanent’ installation in the landscape that is part time-based experiment and part romantic speculation in futility. Located at St Mary’s Loch in the Southern Uplands, a place of inspiration to Romantic poets such as Wordsworth, Scott, Hogg and Burns, ‘Shinglehook’ is conceived as a ‘place to wait for geology’. Four bronze castings were made from a mould taken from the furthest tip of a prograding alluvial fan on the north shore of the loch that will eventually form a land bridge dividing the loch. These metal casts float in a line projecting out from the south shore; the casts can move laterally but are anchored to the shore via cabling secured by two oak drag anchors. The viewer is invited to speculate that ‘one day’ the bronze casts will be enveloped by the arrival of the land bridge across the loch, and the mobile structure of ‘Shinglehook’ will become static. (Photograph: Matt Baker.)

36

J. E. GORDON & M. BAKER

The Enlightenment was the time of creation of disciplines and the bracketing of certain areas of knowledge and enquiry. Geology (and specifically the interpretation of geology) could be viewed as a progressive linear sequence of acquiring empirical knowledge that can be relayed to the viewer as the way of understanding the landscape in front of them. Weaving around, over and through this empirical line is a second story however: the story of how the creative mind has wrestled with geology and continually reinvented a human mythology around stones and landscape. Can we rediscover that former sense of wonder about the physical landscape that has diminished over time as geology has become increasingly specialized and detached from people’s everyday experience? This is a serious issue since geoscience is fundamental to society in many different ways expressed, for example, in the current concept of ecosystem services (Gordon et al. 2012; Gordon & Barron 2013; Gray et al. 2013; Prosser et al. 2013). Geodiversity also has many strong cultural resonances (Ellsworth & Kruse 2012; Gordon 2012b), and human activity is now having such a widespread impact on the planet that a new geological period has been proposed: the Anthropocene (Crutzen 2002; Zalasiewicz et al. 2010, 2011). Yet despite its relevance, the role of geoscience is frequently overlooked at all levels from public awareness to policy formation. We propose the idea of a human identity in relation to a geological timescale as a connecting theme for the building of geological mythologies, a theme that has the idea of unconformity at its heart. Experiential re-engagement through the ‘search for new unconformity’ offers a means of provoking interest (Baker & Gordon 2012). This requires thinking about best-practice interpretation sensu stricto that, for the great majority of visitors, extends beyond the presentation of information. However, the latter is also needed for appropriate parts of the continuum (or ‘opportunity spectrum’) of education-interpretation needs and messages for different audiences, ranging from those actively seeking geoeducation as a primary focus to the general public, policy makers and those with no desire for geoeducation and simply interested in ‘being there’ (Watson 2010). It requires thinking about the process of engagement with people and how to add value to their experiences and provide involvement that evokes a sense of wonder about, as well as interpretation of and education about, the physical landscape. This means moving from appreciating landscape stories through geologists’ eyes to encouraging the rediscovery of a sense of wonder through alternative narratives about the landscape (Lanza & Negrete 2007; Dixon et al. 2012; Gordon 2012a; Stewart & Nield 2013;

Walliss & Kok 2014). Such an approach can help to reconcile the natural and cultural worlds and enable wider public appreciation of geodiversity through reconnecting with the landscape in a holistic way. We thank R. Barton for providing Figure 5 and information about Geopark Shetland and T. Hose for helpful comments that improved the text.

References Anderson, A. 1980. Ben Peach’s Scotland. Institute of Geological Sciences, Edinburgh. Anderson, G. & Anderson, P. 1851. Guide to the Highlands and Islands of Scotland Including Orkney and Zetland Descriptive of Their Scenery, Statistics, Antiquaries, and Natural History. Containing also Directions for Visiting the Lowlands of Scotland with Descriptive Notices, and Maps, Views, Tables of Distances, Notices of Inns, etc. 3rd edn. Adam & Charles Black, Edinburgh. Andrews, M. 1987. The Search for the Picturesque: Landscape, Aesthetics and Tourism in Britain, 1760–1800. Scolar Press, Aldershot. Ascherson, N. 2002. Stone Voices. Granta, London. Baker, M. & Gordon, J. E. 2012. Unconformities, schisms and sutures – geology and mythology in Scotland. In: Ellsworth, E. & Kruse, J. (eds) Making the Geologic Now. Punctum Books, New York, 163–169. Beattie, W. 1838. Scotland Illustrated in a Series of Views Taken Expressly for this Work by Messrs T. Allom, W. H. Bartlett and H. M’Culloch. George Virtue, London, 2 vols. Bertram, G. 2012. The Etchings of John Clerk of Eldin. Enterprise Editions, Thurloxton. Black, A. & C. 1847. Black’s Picturesque Tourist of Scotland. 5th edn. A & C Black, Edinburgh. Boswell, J. 1785. The Journal of a Tour to the Hebrides, with Samuel Johnson, LL.D. Charles Dilly, London. Braithwaite, C. J. R. 2011. Transactions and neglected data. Scottish Journal of Geology, 47, 179–188. Buckland, A. 2013. Novel Science. Fiction and the Invention of Nineteenth-Century Geology. University of Chicago Press, Chicago & London. Butler, R. W. 1985. Evolution of tourism in the Scottish Highlands. Annals of Tourism Research, 12, 371– 391. Campbell, M. 1993. The Line of Tradition. Watercolours, Drawings & Prints by Scottish Artists 1700–1990. National Galleries of Scotland, Edinburgh. Cook, T. 1866. Cook’s Scottish Tourist Practical Directory: A Guide to the Principle Tourist Routes, Conveyances, and Special Ticket Arrangements, Sanctioned by Railway, Steamboat and Coach Companies, Commanding the Highland Excursion Traffic. 3rd edn. Thomas Cook, London. Craig, G. Y. (ed.) 1978. James Hutton’s Theory of the Earth: The Lost Drawings. Scottish Academic Press, Edinburgh. Craig, R. 1874. On the section on the Crofthead and Kilmarnock railway in Cowden Glen, Neilston, Renfrewshire, with remarks on the upper boulder clay.

APPRECIATING THE PHYSICAL LANDSCAPE IN SCOTLAND Transactions of the Geological Society of Glasgow, 4, 17– 32. Crosskey, H. W. 1874. Post Tertiary fossiliferous beds of Scotland. VII. Garvel Park New Dock, Greenock. Transactions of the Geological Society of Glasgow, 4, 32– 45. Crutzen, P. J. 2002. Geology of mankind. Nature, 415, 23. Daniell, W. 1818. Illustrations of the Island of Staffa, in a Series of Views, Accompanied by Topographical and Geological Descriptions. Longman, Hurst, Rees, Orme, and Brown, London. Daniell, W. 1818–1822. A Voyage Round Great Britain, Undertaken in the Summer of the Year 1813, and Commencing from the Land’s-End, Cornwall. With a Series of Views, Illustrative of the Character and Prominent Features of the Coast, by William Daniell, ARA. Longman, Hurst, Rees, Orme, and Brown, London, 3 –6. Daniell, W. 1820. A Voyage Round Great Britain, Undertaken in the Summer of the Year 1813, and Commencing from the Land’s-End Cornwall. With a Series of Views, Illustrative of the Character and Prominent Features of the Coast, by William Daniell, ARA. Longman, Hurst, Rees, Orme, and Brown, London, 4. Dean, D. R. 2007. Romantic Landscapes. Geology and its Cultural Influence in Britain, 1765–1835. Scholars’ Facsimiles & Reprints, Ann Arbor. Defoe, D. 1726. A Tour Thro’ the Whole Island of Great Britain, Divided into Circuits or Journies. London, 3. Dixon, D. P., Hawkins, H. & Straughan, E. R. 2012. Wonder-full geomorphology: sublime aesthetics and the place of art. Progress in Physical Geography, 37, 227–247. Durie, A. J. 2003. Scotland for the Holidays. A History of Tourism in Scotland, 1780–1939. Tuckwell Press, East Linton. Durie, A. J. 2012. Travels in Scotland, 1788–1881. A Selection from Contemporary Tourist Journals. Boydell & Brewer, Woodbridge, Suffolk. Ellis, N. 2011. The Geological Conservation Review (GCR) in Great Britain – rationale and methods. Proceedings of the Geologists’ Association, 122, 353–362. Ellsworth, E. & Kruse, J. (eds) 2012. Making the Geologic Now. Punctum Books, New York. Gairn, L. 2008. Ecology and Modern Scottish Literature. Edinburgh University Press, Edinburgh. Gaskill, H. (ed.) 1996. The Poems of Ossian and Related Works. Edinburgh University Press, Edinburgh. Geikie, A. 1896. Scottish mountains. Scottish Mountaineering Club Journal, 4, 113– 125. Geikie, A. 1901. The Scenery of Scotland Viewed in Connection with its Physical Geology. 3rd edn. Macmillan, London. Geikie, A. 1905. Landscape in History and Other Essays. Macmillan, London. Gilpin, W. 1789. Observations, Relative Chiefly to Picturesque Beauty, Made in the Year 1776, on Several Parts of Great Britain; Particularly the High-Lands of Scotland. R. Blamire, London, 2 vols. Glendening, J. 1997. The High Road: Romantic Tourism, Scotland and Literature, 1720– 1820. Macmillan, Basingstoke.

37

Gold, J. R. & Gold, M. M. 1995. Imagining Scotland. Tradition, Representation and Promotion in Scottish Tourism since 1750. Scolar Press, Aldershot. Gordon, J. E. 2012a. Rediscovering a sense of wonder: geoheritage, geotourism and cultural landscape experiences. Geoheritage, 4, 65–77. Gordon, J. E. 2012b. Engaging with geodiversity: ‘stone voices’, creativity and ecosystem cultural services in Scotland. Scottish Geographical Journal, 128, 240– 265. Gordon, J. E. & Barron, H. F. 2013. Geodiversity and ecosystem services in Scotland. Scottish Journal of Geology, 49, 41–58. Gordon, J. E. & Kirkbride, V. 2009. Reading the landscape: unveiling Scotland’s Earth stories. Proceedings, Vital Spark Conference, Aviemore, October 2007. Available at: http://www.ahi.org.uk/include/pdf/TVS papers/Gordon_J_and_Kirkbride_V.pdf. Gordon, J. E., Brazier, V. & Macfadyen, C. C. J. 2004. Reading the landscapes of Scotland: raising earth heritage awareness and enjoyment. In: Parkes, M. (ed.) Natural and Cultural Landscapes – the Geological Foundation. Royal Irish Academy, Dublin, 227– 234. Gordon, J. E., Barron, H. F., Hansom, J. D. & Thomas, M. F. 2012. Engaging with geodiversity – why it matters. Proceedings of the Geologists’ Association, 123, 1– 6. Gray, M., Gordon, J. E. & Brown, E. J. 2013. Geodiversity and the ecosystem approach: the contribution of geoscience in delivering integrated environmental management. Proceedings of the Geologists’ Association, 124, 659–673. Grenier, K. H. 2005. Tourism and Identity in Scotland, 1770– 1914: Creating Caledonia. Ashgate Publishing Ltd, Aldershot. Ham, S. 2007. From interpretation to protection. Interpretation Journal, 12(3), 20–23. Harker, A. 1900. Notes, geological and topographical, on the Cuillin hills, Skye. Scottish Mountaineering Club Journal, 6, 1– 13. Heringman, N. 2004. Romantic Rocks, Aesthetic Geology. Cornell University Press, Ithaca & London. Hill, D. O., Wilson, J. & Chambers, R. 1840. The Land of Burns, A Series of Landscapes & Portraits Illustrative of the Life and Writings of the Scottish Poet. Blackie & Son, Glasgow. Holloway, J. & Errington, L. 1978. The Discovery of Scotland. The Appreciation of Scottish Scenery Through Two Centuries of Scottish Painting. National Gallery of Scotland, Edinburgh. Hose, T. A. 2008. Towards a history of geotourism: definitions, antecedents and the future. In: Burek, C. V. & Prosser, C. D. (eds) The History of Geoconservation. Geological Society, London, Special Publications, 300, 37–60. Hose, T. A. 2010a. Volcanic geotourism in West Coast Scotland. In: Erfurt-Cooper, P. & Cooper, M. (eds) Volcano and Geothermal Tourism: Sustainable Geo-resources for Leisure and Recreation. Earthscan, London, 259–271. Hose, T. A. 2010b. The significance of aesthetic landscape appreciation to modern geotourism provision. In: Newsome, D. & Dowling, R. K. (eds)

38

J. E. GORDON & M. BAKER

Geotourism: The Tourism of Geology and Landscape. Goodfellow Publishers, Oxford, 13– 26. Hose, T. A. 2012. 3G’s for modern geotourism. Geoheritage, 4, 7 –24. Hunter, J. 1995. On the Other Side of Sorrow. Nature and People in the Scottish Highlands. Mainstream Publishing, Edinburgh & London. Hutton, J. 1785. Abstract of a Dissertation Read in the Royal Society of Edinburgh, upon the Seventh of March, and Fourth of April, MDCCLXXXV, Concerning the System of the Earth, its Duration, and Stability. Edinburgh. Hutton, J. 1788. Theory of the Earth; or an investigation of the laws observable in the composition, dissolution, and restoration of land upon the globe. Transactions of the Royal Society of Edinburgh, 1, 209–304. Jameson, R. 1813. Mineralogical Travels Through the Hebrides, Orkney and Shetland Islands, and Mainland of Scotland, with Dissertations upon Peat and Kelp. Archibald, Constable & Co, Edinburgh, 1. Johnson, S. 1775. A Journey to the Western Islands of Scotland. W. Strahan & T. Cadell, London. Klonk, C. 1996. Science and the Perception of Nature: British Landscape Art in the Late Eighteenth and Early Nineteenth Centuries. Yale University Press, New Haven. Klonk, C. 1997. From picturesque travel to scientific observations: artists’ and geologists’ voyages to Staffa. In: Rosenthal, M., Payne, C. & Wilcox, S. (eds) Prospects for the Nation: Recent Essays in British Landscape, 1750–1880. Yale University Press, New Haven & London, 205 –229. Knell, S. J. & Taylor, M. A. 2006. Hugh Miller: fossils, landscape and literary geology. Proceedings of the Geologists’ Association, 117, 85–98. Lane, S. N., Odoni, N., Landstro¨m, C., Whatmore, S. J., Ward, N. & Bradley, S. 2011. Doing flood risk science differently: an experiment in radical scientific method. Transactions of the Institute of British Geographers, 36, 15–36. Lanza, T. & Negrete, A. 2007. From myth to earth education and science communication. In: Piccardi, L. & Masse, W. B. (eds) Myth and Geology. Geological Society, London, Special Publications, 273, 61–66. Lothian and Borders Geoconservation 2011. Canongate Wall. Art, Science and Politics. (Local Geodiversity Site leaflet, Scottish Parliament Building). Lothian and Borders GeoConservation, Edinburgh. Lumsden J. & Son 1839. Lumsden and Son’s Steam-boat Companion; or Stranger’s Guide to the Western Isles and Highlands of Scotland, including the Voyages from London to Leith, and from Glasgow to Liverpool and Belfast; with a Land Tour to the Giant’s Causeway, etc. Also, a Short Tour to the Lakes of Cumberland. Oliver & Boyd, W. & A.K. Johnston, Edinburgh. Lyell, C. 1830– 1833. Principles of Geology, Being an Attempt to Explain the Former Changes of the Earth’s Surface, by References to Causes Now in Operation. John Murray, London. MacBrayne, D. 1886. Summer Tours in Scotland. Glasgow to the Highlands. ‘Royal Route.’ (Via Crinan and Caledonian Canals.) With Time Tables and List

of Fares, by David MacBrayne’s Royal Mail Steamers, ‘Columba’, ‘Iona’, &c. Official Guide. New edn. David MacBrayne, Glasgow. Macculloch, J. 1819a. A Description of the Western Islands of Scotland, Including the Isle of Man: Comprising an Account of their Geological Structure; with Remarks on their Agriculture, Scenery and Antiquities. Archibald Constable & Co, Edinburgh; Hurst, Robinson & Co, London, 2. Macculloch, J. 1819b. A Description of the Western Islands of Scotland, Including the Isle of Man: Comprising an Account of their Geological Structure; with Remarks on their Agriculture, Scenery and Antiquities. Archibald Constable & Co, Edinburgh; Hurst, Robinson & Co, London, 3. Macculloch, J. 1824. The Highlands and Western Isles of Scotland, Containing Descriptions of their Scenery and Antiquities, with an Account of the Political History and Ancient Manners, and of the Origin, Language, Agriculture, Economy, Music, Present Condition of the People, &c &c &c Founded on a Series of Annual Journeys Between the Years 1811 and 1821, and forming an Universal Guide to that Country, in Letters to Sir Walter Scott, Bart. Longman, Hurst, Rees, Orme, Brown & Green, London, 4 Vols. Macculloch, J. 1836. A Geological Map of Scotland by Dr. Macculloch, F.R.S., etc. Published by Order of the Lords of the Treasury by S. Arrowsmith, Hydrographer to the King, London. Macdonald, M. 2000. Scottish Art. Thames & Hudson, New York. Macfarlane, R. 2013. New words on the wild. Nature, 498, 166 –167. MacLeod, A. 2012. From an Antique Land. Visual Representations of the Highlands and Islands 1700– 1880. John Donald, Edinburgh. Macmillan, D. 2000. Scottish Art: 1460– 2000. Mainstream Publishing, Edinburgh & London. Macpherson, J. 1760. Fragments of Ancient Poetry, Collected in the Highlands of Scotland, and Translated from the Gaelic or Erse Language. G. Hamilton & J. Balfour, Edinburgh. Macpherson, J. 1765. Works of Ossian, the Son of Fingal, in Two Volumes. Translated from the Gaelic Language by James Macpherson. 3rd edn. T. Becket & P. A. Dehondt, London. McIntosh, A. 2001. Soil and Soul. People Versus Corporate Power. Aurum Press, London. McKeever, P. J., Zouros, N. & Patzak, M. 2010. The UNESCO Global Geoparks Network. European Geoparks Magazine, 7, 10–13. Available at: http:// www.scribd.com/doc/80356479/EGN-MagazineIssue-7 McPhee, J. 1981. Basin and Range. Farrar, Straus and Giroux, New York. Miller, H. 1858. The Cruise of the Betsey; or, A Summer Ramble among the Fossiliferous Deposits of the Hebrides. With Rambles of a Geologist; or, Ten Thousand Miles over the Fossiliferous Deposits of Scotland. Thomas Constable & Co, Edinburgh. Miller, H. 1892. Landscape geology: a plea for the study of geology by landscape-painters. Transactions of the Edinburgh Geological Society, 6, 129–154.

APPRECIATING THE PHYSICAL LANDSCAPE IN SCOTLAND Milne Home, D. 1872. Scheme for the conservation of remarkable boulders in Scotland, and for the indication of their position on maps. Proceedings of the Royal Society of Edinburgh, 7, 475– 488. Milne Home, D. 1884. Tenth and final report of the Boulder Committee; with appendix containing an abstract of the information in the nine annual reports of the Committee; and a summary of the principal points apparently established by the information so received. Proceedings of the Royal Society of Edinburgh, 12, 765– 926. Morrison, J. 2003. Painting the Nation. Identity and Nationalism in Scottish Painting, 1800–1920. Edinburgh University Press, Edinburgh. Murray, J. 1867. Handbook for Travellers in Scotland. John Murray, London. Murray, S. 1799. A Companion, and Useful Guide to the Beauties of Scotland, to the Lakes of Westmoreland, Cumberland, and Lancashire; and to the Curiosities in the District of Craven, in the West Riding of Yorkshire. To which is Added a More Particular Description of Scotland, Especially that Part of it, Called the Highlands. Published by the author, London. Murray, S. 1805. A Companion and Useful Guide to the Beauties in the Western Highlands of Scotland, and in the Hebrides. To Which is Added, a Description of Part of the Main Land of Scotland, and of the Isles of Mull, Ulva, Staffa, I-Columbkill, Tirii, Coll, Eigg, Skye, Raza, and Scalpa. 2nd edn. Printed by W. Bulmer & Co. for the author, London, 2. Nicolson, M. H. 1997. Mountain Gloom and Mountain Glory. The Development of the Aesthetics of the Infinite. University of Washington Press, Seattle & London. O’Connor, R. 2007. The Earth on Show. Fossils and the Poetics of Popular Science, 1802– 1856. University of Chicago Press, Chicago & London. Peach, B. N. & Horne, J. 1921. Geological features of Scottish mountains. In: Young, J. R. (ed.) General Guide-Book. Geology: Meteorology: Botany: Bird Life: Equipment: Maps, &c.: Rock & Snow Craft: Photography. Sir Hugh Munro’s Classified Tables of the 3000 ft. Mountains of Scotland. Scottish Mountaineering Club, Edinburgh. Pennant, T. 1771. A Tour in Scotland 1769. John Monk, Chester. Pennant, T. 1776. A Tour in Scotland and Voyage to the Hebrides, 1772. Benjamin White, London. Playfair, J. 1802. Illustrations of the Huttonian Theory of the Earth. William Creech, Edinburgh. Playfair, J. 1805. Biographical account of the late Dr James Hutton, F.R.S. Edin. Transactions of the Royal Society of Edinburgh, 5, 39–99. Pococke, R. 1887. Tours in Scotland 1747, 1750, 1760. (Edited with a Biographical Sketch of the Author by Daniel William Kemp.) Scottish History Society, Edinburgh. Pointon, M. 1979. Geology and landscape painting in nineteenth-century England. In: Jordanova, L. J. & Porter, R. S. (eds) Images of the Earth: Essays in the History of the Environmental Sciences. The British Society for the History of Science, Chalfont St. Giles, 84– 108.

39

Prosser, C. D. 2013. Planning for geoconservation in the 1940s: an exploration of the aspirations that shaped the first national geoconservation legislation. Proceedings of the Geologists’ Association, 124, 536– 546. Prosser, C. D., Burek, C. V., Evans, D. H., Gordon, J. E., Kirkbride, V. B., Rennie, A. F. & Walmsley, C. A. 2010. Conserving geodiversity sites in a changing climate: management challenges and responses. Geoheritage, 2, 123–136. Prosser, C. D., Brown, E. J., Larwood, J. G. & Bridgland, D. R. 2013. Geoconservation for science and society – an agenda for the future. Proceedings of the Geologists’ Association, 124, 561– 567. Ramsay, A. C. 1841. The Geology of the Island of Arran from Original Survey. R. Griffin, Glasgow. Ramsay, A. C. 1894. The Physical Geology and Geography of Great Britain: A Manual of British Geology. 6th edn. Edward Stanford, London. Robson, G. F. 1814. Scenery of the Grampian Mountains; Illustrated by Forty Etchings in the Soft Ground. G. F. Robson, London. Ross, I. 1965. A bluestocking over the border: Mrs. Elizabeth Montagu’s aesthetic adventures in Scotland, 1776. Huntington Library Quarterly, 28, 213–233. Rudwick, M. J. S. 1976. The emergence of a visual language for geological science 1760–1840. History of Science, 14, 149–195. Rudwick, M. J. S. 1992. Scenes From Deep Time. Early Pictorial Representations of the Prehistoric World. University of Chicago Press, Chicago & London. Rudwick, M. J. S. 2005. Bursting the Limits of Time. The Reconstruction of Geohistory in the Age of Revolution. University of Chicago Press, Chicago & London. Rudwick, M. J. S. 2008. Worlds Before Adam. The Reconstruction of Geohistory in the Age of Reform. University of Chicago Press, Chicago & London. Ruskin, J. 1844. Modern Painters: Their Superiority in the Art of Landscape Painting to All the Ancient Masters Proved by Examples of the True, the Beautiful, and the Intellectual, from the Works of Modern Artists, Especially from Those of J. M. W. Turner, Esq., R.A. 2nd edn. Smith, Elder & Co, London. Sandby, P. 1778– 1781. The Virtuosi’s Museum. G. Kearsly, London. Scott, W. 1810. The Lady of the Lake. A Poem. 8th edn. John Ballantyne & Co, Edinburgh. Scottish Geodiversity Forum 2013. Scotland’s Geodiversity Charter. Available at: http://scottishgeodiversi tyforum.org/charter/. Smout, T. C. 1982. Tours in the Scottish Highlands from the eighteenth to the twentieth centuries. Northern Scotland, 5, 99– 121. Stafford, F. 2005. Scottish Romanticism and Scotland in Romanticism. In: Ferber, M. (ed.) A Companion to European Romanticism. Blackwell Publishing, Oxford, 49– 66. Stewart, I. S. & Nield, T. 2013. Earth stories: context and narrative in the communication of popular geosciences. Proceedings of the Geologists’ Association, 124, 699–712. Storer, J. & Greig, J. 1805. Views in North Britain Illustrative of the Works of Robert Burns. Vernor and Hood, London.

40

J. E. GORDON & M. BAKER

Sugden, D. E. 1968. The selectivity of glacial erosion in the Cairngorm Mountains, Scotland. Transactions of the Institute of British Geographers, 45, 79– 92. Taylor, M. A. 2007. Hugh Miller. Stonemason, Geologist, Writer. NMS Enterprises Limited – Publishing, Edinburgh. Thomas, B. A. & Warren, L. M. 2008. Geological conservation in the nineteenth and early twentieth centuries. In: Burek, C. V. & Prosser, C. D. (eds) The History of Geoconservation. Geological Society, London, Special Publications, 300, 17– 30. Tilden, F. 1977. Interpreting Our Heritage. 3rd edn. University of North Carolina Press, Chapel Hill. Verne, J. 1864. Voyage au Centre de la Terre. Hetzel, Paris. Walford, T. 1818. The Scientific Tourist Through England, Wales, & Scotland: by Which the Traveller is Directed to the Principal Objects of Antiquity, Art, Science, and the Picturesque, Including the Minerals, Fossils, Rare Plants, and Other Subjects of Natural History: Arranged by Counties. To Which is Added an Introduction to the Study of Antiquities, and the Elements of Statistics, Geology, Mineralogy, and Botany. John Booth, London, 2 Vols. Walliss, J. & Kok, K. 2014. New interpretative strategies for geotourism: an exploration of two

Australian mining sites. Journal of Tourism and Cultural Change, 12, 33– 49. Watson, J. 2010. Simply ‘being there’: a legitimate point on the geotourism opportunity spectrum. Australasian Cave & Karst Management Association (ACKMA) Journal, 80, 35–38. White, K. 2003. Geopoetics: Place, Culture, World. Alba Editions, Glasgow. Whiteley, M. J. & Browne, M. A. E. 2013. Local geoconservation groups – past achievements and future challenges. Proceedings of the Geologists’ Association, 124, 674– 680. Wordsworth, D. 1874. Recollections of a Tour Made in Scotland A.D. 1803, (ed.) Shairp, J. C. Edmonston and Douglas, Edinburgh. Wyatt, J. 1995. Wordsworth and the Geologists. Cambridge University Press, Cambridge. Zalasiewicz, J., Williams, M., Steffen, W. & Crutzen, P. J. 2010. The new world of the Anthropocene. Environmental Science and Technology, 44, 2228– 2231. Zalasiewicz, J., Williams, M., Haywood, A. & Ellis, M. 2011. The Anthropocene: a new epoch of geological time? Philosophical Transactions of the Royal Society A, 369(1938), 835– 841.

Waterfalls and the Romantic traveller BRIAN J. HUDSON School of Civil Engineering and Built Environment, Queensland University of Technology, 2 George Street, GPO Box 2434, Brisbane, QLD 4001, Australia (e-mail: [email protected]) Abstract: Waterfalls have long attracted the attention of travellers, some of whom were writers and artists who have left us a cultural legacy of their observations and interpretations. Likewise, geologists have studied and recorded these landscape features since the infancy of their science. An examination of travellers’ experiences of waterfalls since the emergence of Romanticism in eighteenth-century Europe reveals a variety of responses, both utilitarian and aesthetic. Seen as valuable sources of renewable energy, impediments to navigation, beautiful, sublime or picturesque natural wonders and resources for tourism, waterfalls continue to appeal to the Romantic traveller and the pleasure-seeking tourist. Increasingly, waterfalls are being threatened by schemes to exploit them, especially for power generation or intensive tourism development. In many parts of the world, this presents a serious challenge to those responsible for the management of this often spectacular aspect of geodiversity. This paper explores these various themes which are contextualized within the historical and cultural framework of Romanticism.

This paper examines the Romantic traveller’s experience of waterfalls, landscape features that have remained popular with tourists since Romanticism emerged in eighteenth-century Europe. The ‘Romantic era’ is loosely defined, for the purposes of the following discussion, as the period from the late eighteenth to the middle of the nineteenth centuries. This was when the Age of Enlightenment gave way to Romanticism, an aesthetic movement that put less emphasis on reason and greater value on emotion, imagination, individualism and selfexpression. The term ‘romantic’, however, is used more broadly in the following discussion which recognizes that Romanticism continued to resonate long after the middle of the nineteenth century, and acknowledges the use of the word in association with human love. While the term ‘Romanticism’ is normally associated with the arts and humanities, scholars have recognized ‘the potential interconnections in the Romantic period between literature, aesthetic theory and the emergence of Earth sciences’ (Furniss 2010, p. 305). In his study of James Hutton’s influential book Theory of the Earth, Furniss argues that the author, widely regarded as the founder of modern geology, developed ideas ‘that drew on and reconfigured what Heringman calls the ‘geological sublime’ – a geo-aesthetic category that was common to the aesthetic geology and geological aesthetics of the period’ (Heringman 2004; Furniss 2010, p. 307). The sublime is an aesthetic quality that characterized Romantic taste. In his famous treatise of 1757, Edmund Burke made a clear distinction between the beautiful and the sublime (Burke 1757). The former is founded on pleasure, the latter

on pain. Characteristics of the beautiful include smallness, smoothness, gradual variation, lightness and delicacy, while the sublime is associated with vastness, ruggedness, massiveness, darkness and gloom. Beauty evokes a sense of pleasure; in the presence of the sublime we experience a ‘delightful horror’ (Burke 1757, p. 52), a different kind of delight felt ‘when we have an idea of pain and danger, without being actually in such circumstances’ (Burke 1757, p. 32).The beautiful and the sublime are qualities that can be found alone or together in a picturesque landscape. The concept of the picturesque was popularized when William Gilpin published his account of a tour of the Wye Valley and parts of South Wales made in 1770 (Gilpin 1782). A picturesque scene is simply a view that meets the eye as a satisfactory picture, one comparable with the landscape paintings that became fashionable in seventeenth and eighteenth century Europe. The term ‘picturesque’ refers mainly to the arrangement, texture and form of landscape features such as hills, valleys, trees and grass, rocks and water, but it is also often associated with the presence of scenic curiosities, for example old ruins, grottoes, rustic bridges, mills and quaint cottages. Early geologists investigating natural landforms and the rocks which composed them often sought to understand these phenomena in terms of the Bible. Later, when some investigators began to entertain doubts about the literal interpretation of the Scriptures, they became filled with awe at the immensity of geological time and the amazing results of natural processes operating over unimaginable ages. In the words of philosopher Eugene Hargrove (2008, p. 35), ‘While the earlier

From: Hose, T. A. (ed.) 2016. Appreciating Physical Landscapes: Three Hundred Years of Geotourism. Geological Society, London, Special Publications, 417, 41– 57. First published online January 12, 2015, http://dx.doi.org/10.1144/SP417.9 # 2016 The Author(s). Published by The Geological Society of London. All rights reserved. For permissions: http://www.geolsoc.org.uk/permissions. Publishing disclaimer: www.geolsoc.org.uk/pub_ethics

42

B. J. HUDSON

WATERFALLS AND THE ROMANTIC TRAVELLER

conception of time in terms of human history was picturesque, geologic time was sublime’. The publication of accounts of geological investigations and discoveries helped to popularize the science and, in the late nineteenth and early twentieth centuries, tourist guidebooks commonly included details about the local geology in their descriptions of the attractions found in the areas they described. Indeed, ‘geologically based tourism’ or ‘geotourism’ (Hose 2012), as well as tourism associated with the picturesque beauty of landscape, has a history that extends back more than two centuries (Hose 2010; Gordon 2012). The ‘romantic gaze’, by which observers derive aesthetic pleasure from the contemplation of beautiful and sublime landscapes, was the product of European society at a particular time (Urry 1990). This is not to say that other societies in other places did not share similar responses to landscape features such as mountains, rivers and waterfalls. In discussing this subject however it is appropriate to recognize that, at other times and in other places, people responded differently to natural scenery such as wild rivers, rapids, cascades and cataracts. Waterfalls and rapids have long attracted the attention of writers and artists, and geologists have studied them since the infancy of their science. A pioneer in this field of study, as in many others, was Leonardo da Vinci. Leonardo’s accurate geological observations are well illustrated in his earliest dated work, a drawing entitled The Hills of Tuscany or Landscape with River (1473) which features a waterfall dropping over a cliff that exposes clearly defined horizontal strata. Herodotus and Theocritus were classical writers who referred to waterfalls and rapids while, in China, artists and poets have been inspired by waterfalls for over a thousand years. John Leland (c. 1503–1552), William Camden (1551–1623), John Speed (1552?– 1629), Daniel Defoe (c. 1660–1731) and Celia Fiennes (1662–1741) were English travellers who occasionally mentioned waterfalls in their writings, saying little about their appearance (Camden 1610; Defoe 1727; Fiennes 1982; Leland [1710–1712] 1910). In Europe, the first paintings featuring waterfalls appeared around the beginning of the seventeenth century, becoming particularly popular among Romantic painters and poets after the mideighteenth century (Hudson 2012, 2013a; Fig. 1). Writers sometimes noted the utility of falls, for example as sources of energy for operating mills or as good places for fishing. As early as the

43

sixteenth century, antiquarian traveller William Camden noted the prolific salmon fishery on the River Teifi at Cilgarron, Wales, ‘where the river from on high falleth downright’ (Camden 1610, p. 654). Describing his visit to High Force, a waterfall on the River Tees, the eighteenth century traveller John Byng included the observation that ‘the basin at the bottom I should suppose a fine place for fishing, and here the salmon must stop’ (Byng [1792] 1936, p. 70; Fig. 2). By this time, the Industrial Revolution was well advanced, and in Britain and overseas many waterfalls were being harnessed to power machinery of all kinds. This was something appreciated by Jamaican sugar planter Edward Long, who had an eye for both the practical and aesthetic values of cascading rivers. In his History of Jamaica, Long observed ‘The precipitate current of most of the rivers . . . is attended with very happy effects [including] . . . the conveniency which the height of their fall admits for the better taking up and conducting their water to mechanical uses’ (Long 1774, p. 357). Falls and rapids were also regarded as nuisances, notably as impediments to navigation. Among the best-known examples in Europe were the Rhine Falls in Switzerland and Trollhatten Falls in Sweden. As European explorers penetrated more deeply into the lands that they ‘discovered’ overseas, they often reported that their progress was hindered by waterfalls and rapids. The perceived economic and social consequences of these natural barriers in Africa were expressed in typical European supremacist fashion by British geographer George Chisholm (1850–1930) who wrote, ‘The most signal example of the effect of waterfalls and rapids in retarding the development of civilisation is undoubtedly presented by the continent of Africa, the ‘darkness’ of which is almost entirely due to this cause’ (Chisholm 1885, p. 420).

Waterfalls, landscape aesthetics, travel and tourism The aesthetic appreciation of waterfalls has a long recorded history (Hudson 2000). For over a thousand years, waterfalls have featured prominently in Chinese painting and poetry. Explorers, as well as artists, were also moved by the beauty of cascading streams. Sailing through the Caribbean in 1493, Dr Chanca, one of the men accompanying Columbus on his second voyage to America, was impressed

Fig. 1. A high waterfall (1822), by Thomas Rowlandson (1756– 1827). Watercolour with pen and brown, grey and black ink over graphite (30.5 cm × 40.2 cm). Yale Center for British Art, Paul Mellon Collection. Best known for his caricatures and depictions of lowlife, this English artist also included landscapes among his wide range of subjects. The plunging waterfall, rugged cliffs, tumbled rocks and tangled vegetation seen in this picture are features typically found in landscapes of the Romantic period.

44

B. J. HUDSON

WATERFALLS AND THE ROMANTIC TRAVELLER

by the beauty of a waterfall that he saw as they approached the island of Guadeloupe (Chanca 1969, p. 133). Scottish explorer and missionary David Livingstone enthused over the beauty of Victoria Falls (or Mosi-oa-Tunya, meaning ‘the Smoke that Thunders’), famously suggesting that ‘scenes so lovely must have been gazed upon by angels in their flight’ (Livingstone 1858, p. 558). The seventeenth century Dutch painter and engraver Allaertvan Everdingen (1621–1675) was fascinated by the waterfalls he saw when he travelled to Norway and Sweden in 1644; many of his landscape paintings depict cascading rivers in wild settings, a motif that was enthusiastically adopted by fellow Dutch artist Jacob van Ruisdael (1629–1682). Wild scenery with torrential streams and waterfalls did not appeal to all travellers, however. The former English county of Westmorland (now subsumed into the modern county of Cumbria), whose mountains, lakes and waterfalls were beloved by the Romantic poets and painters of the late eighteenth and early nineteenth centuries, was unpleasing to Daniel Defoe; he described it as ‘a Country eminent only for being the wildest, most barren and frightful of any that I have passed over in England, or even Wales it self’ (Defoe 1727, p. 223). The cascading Lakeland streams that so appealed to the Romantic travellers of later years appear not to have impressed Defoe who makes very little mention of waterfalls in his famous Tour Thro’ the Whole Island of Great Britain, published in three volumes (1724– 1727). References to waterfalls in Defoe’s fictional A New Voyage Round the World reflect a mainly negative attitude to these landscape features. Importantly, they are seen as impediments to navigation. In Defoe’s narrative, explorers travelling by boat on a South American river encountered ‘a mighty cataract or water-fall’, a ‘frightful’ sight, made worse by the fact that it was but the first of several that they discovered on this stretch of river. ‘These cataracts made the river perfectly useless to them for above twenty miles’ (Defoe 1810, pp. 201–203). To the gold-seeking explorers, one good thing about the falls was their erosive power which had excavated a deep hollow in the rock and deposited alluvial gold on a nearby shoal (Defoe 1810, p. 203). Elsewhere, as they crossed the Andes, Defoe’s explorers were not favourably impressed by the landscape where there was ‘little to be seen but steep precipices, inhospitable rocks, and impassable mountains, immuring us on every side, innumerable rills

45

and brooks of water falling from the clifts [sic], making a barbarous and unpleasant sound’ (Defoe 1810, p. 79). A famous traveller who was very favourably impressed by the waterfalls he saw on his South American journeys was the German polymath Alexander von Humboldt (1769– 1859), whose lifetime spanned the Romantic period (Hudson 2013a). Already widely travelled in Europe, Humboldt went on an expedition to Latin America with companion Aime´ Bonpland from 1799 to 1804. His investigations took him from the Amazon Basin and the Andes to the Caribbean island of Cuba, from the subterranean depths of the Gua´charo Cave in present-day Venezuela, to the snowy peak of Chimborazo. Humboldt’s meticulously recorded scientific and other observations ranged from meteorology, geology and biology to anthropology, geography, history, economics and politics. Having studied at the Freiberg University of Mining and Technology in 1791–1792, Humboldt was well equipped with the theoretical and practical knowledge that he was able to apply on his explorations. He made detailed observations on waterfalls and rapids, discussing them in terms of both science and aesthetics (Humboldt 1821; Fig. 3). In his account of Tequendama Falls, in what is now Colombia, Humboldt reflects on the aesthetic qualities of waterfalls in general: ‘The impression they leave on the mind of the observer depends on the concurrence of a variety of circumstances. The volume of the water must be proportioned to the height of the fall, and the scenery around must wear a wild and romantic aspect’ (Humboldt 1814, p. 76). Humboldt’s published writings were highly influential, and among those whom he inspired to follow in his footsteps were English scientist Charles Darwin and American artist Frederic Edwin Church. Books such as Darwin’s Voyage of the Beagle (1839) and paintings such as Church’s Tequendama Falls, Near Bogota, New Grenada (1854) brought the wonders of South America to the attention of a wider public. However, it is with British travellers of the Romantic period that the rest of this paper is concerned. Waterfalls were among the sights that distinguished man of letters Dr Samuel Johnson and James Boswell, his biographer, hoped to see when they planned their tour of Scotland in 1773 (Birkbeck Hall 1964, p. 112). The main purpose of their journey was ‘to contemplate a system of life almost totally different from what we had been accustomed

Fig. 2. High Force or Fall of Tees, after J. M. W. Turner (1775–1851), line engraved by J. Landseer (1769– 1862).This is an illustration from An History of Richmondshire by Thomas D. Whitaker, published in 1823, with many illustrations based on Turner’s paintings. Two fishermen can be seen in the bottom-left corner of the picture. The towering cliffs clearly display their geological structure, with the Whin Sill overlying the horizontally bedded Carboniferous Limestone.

46

B. J. HUDSON

Fig. 3. Cascade de Tequendama (1810), after Alexander von Humboldt (1769– 1859). Line engraved book illustration from Humboldt’s Vues des Cordille`res (Paris, 1810). Humboldt made a sketch of this waterfall when he visited it in 1801.

WATERFALLS AND THE ROMANTIC TRAVELLER

to see; and, to find simplicity and wildness, and all circumstances of remote time or place’ (Birkbeck Hall 1964, p. 13). Not usually regarded as writers of the Romantic period, both Johnson and Boswell were, nevertheless, familiar with the concept of Romanticism. Boswell described as ‘romantic’ some of the landscapes he saw in Switzerland and his native Scotland. The ‘beautiful wild valley’ of the River Reuse, ‘surrounded by immense mountains, some covered with frowning rocks, others with clustering pines, and others with glittering snow’, created a ‘romantic prospect’ (Boswell [1764] 1953, p. 215). Describing a Scottish landscape with a ruined castle in its rocky, wooded setting beside a river, Boswell remarks, ‘I cannot figure a more romantick [sic] scene’ (Birkbeck Hall 1964, p. 379). On his tour of North Wales in 1774, Johnson saw a wood which he describes as ‘diversified and romantic’ (Johnson [1774] 1816, p. 23) while, in his opinion, Beaumaris Castle, with its lofty tower, deep dungeon and private passage, ‘corresponds with all the representations of romancing narratives’ (Johnson [1774] 1816, pp. 101 –102). In their travel writings, both men express a fascination with waterfalls. Indeed, in Boswell’s words, ‘Dr. Johnson . . . always said, that he was not come to Scotland to see fine places, of which there were enough in England; but wild objects, – mountains, – waterfalls . . .; in short, things he had not seen before’ (Birkbeck Hall 1964, p. 112). Johnson was a friend of Edmund Burke, author of the influential Philosophical Enquiry Into the Origin of Our Ideas of the Sublime and Beautiful, first published in 1756. While, according to Boswell, he and his companion Dr Johnson ‘at no time had much taste for rural beauties’ (Birkbeck Hall 1964, p. 112), both men display an appreciation of the Sublime and the Beautiful in their travel writings. In particular, Johnson makes full use of the vocabulary of the Sublime when describing waterfalls (Hudson 2013b). The enjoyment of landscape was not the primary reason for the journeys undertaken by Johnson and Boswell; they were more interested in the lives of the people who lived in the places they visited. Nevertheless, the published accounts of their travels include geological observations of some landscape features that they saw (Hose 2010, p. 265). The following extract from Johnson’s account of his Scottish tour, published in 1775, reveals a remarkable appreciation of geomorphological processes: We passed many rivers and rivulets, which commonly ran with a clear shallow stream over a hard pebbly bottom. These channels, which seem so much wider than the water that they convey would naturally require, are caused by the violence of wintry floods, produced by the accumulation of innumerable streams

47

that fall in rainy weather from the hills, and bursting away with restless impetuosity, make themselves a passage proportionate to their mass (Johnson [1775] 1944, p. 34).

There were eighteenth-century travellers who journeyed mainly for aesthetic pleasure, however. On his summer holidays, William Gilpin (1724–1804) enjoyed walking tours in the British countryside, recording his observations with sketches and commentaries. In this activity he was deliberately: searching after effects. This is the general intention of picturesque travel. We mean not to bring it into competition with any of the more useful ends of travelling. But as many travel without any end at all, amusing themselves without being able to give a reason why they are amused, we offer an end, which may possibly engage some vacant minds; and may indeed afford a rational amusement to such as travel for more important purposes (Gilpin 1792, p. 41).

As Gilpin explains, a picturesque scene contains elements of the Beautiful or the Sublime or both, forming a visual composition that meets the eye as a satisfying picture. For example, Lodore Falls in the Lake District was described as presenting ‘a singularly harmonious assembledge of the sublime and the beautiful’ (Robinson 1819, p. 127; Fig. 4). Waterfalls are among the landscape features Gilpin describes and illustrates in his travel writings and drawings. In one of his books, Gilpin devotes several pages to the discussion of the picturesque beauties of ‘the cascade, which . . . may be divided into the broken and the regular fall’ (Gilpin 1808, p. 117). Gilpin’s example was followed by others, such as Sir Uvedale Price, author of the influential Essay on the Picturesque, and poet Thomas Gray, both of whom wrote about their journeys and the aesthetic enjoyment of landscape (Price 1794, 1801; Gray 1820). Publications of this kind further popularized travel to places of scenic beauty. This was the beginning of the remarkable growth in tourism which had its origin in The Grand Tour, a European peregrination that became fashionable in eighteenth century British society (Fig. 5). With the development of tourism came the tourist guidebook written for the use of visitors seeking to experience beautiful, sublime and picturesque scenery. It was in Britain and Germany that these travellers’ tools were invented (Sillitoe 1995), with geological field guides following soon after (Hose 2008). While Burke, Gilpin and others explained what to look for in the landscape, guidebook writers such as Thomas West and William Wordsworth told readers where and how to find the best views. In Britain, the vogue for landscape tourism was greatly encouraged by the publication of Gilpin’s Observations on the River Wye, and Several Parts of South Wales, etc. relative chiefly to Picturesque Beauty; made in the Summer of the

48

B. J. HUDSON

WATERFALLS AND THE ROMANTIC TRAVELLER

Year 1770 (Andrews 1989, p. 86; Gilpin 1782). Thereafter, tourist guidebook writers began to produce popular volumes for the use of visitors to scenically attractive regions such as North Wales, the Peak District, the Lake District and the Scottish Highlands. Four years after Gilpin’s Welsh tour, Henry Wyndham spent six weeks travelling through Wales, an account of which was published in 1775. In the Preface, Wyndham informs his readers that: The Author of the following concise Tour, has no other view in the publication of it, than a desire of inducing his countrymen to consider Wales, as an object worthy attention. The romantic beauties of nature are so singular and extravagant in the principality, particularly in the counties of Merioneth and Caernarvon, that they are scarcely to be conceived, by those, who have confined their curiosity to other parts of Great Britain (Wyndham 1775, p. i).

The author notes that, ‘while the English roads are crouded [sic] with travelling parties of pleasure, the Welsh are . . . rarely visited’ (Wyndham 1775, p. ii). This situation was soon to change and, like the Lake District, Wales (particularly rugged Snowdonia) began to attract crowds of tourists. The tourism boom in Britain spawned many popular guidebooks; among the best known is Thomas West’s Guide to The Lakes in Cumberland, Westmorland and Lancashire, first published in 1778. In this book, the author identifies what he calls ‘stations’, viewpoints from which visitors could fully appreciate the aesthetic qualities of the selected scenes. As the author explains, the absence of views from mountain summits, and his focus on the region’s lakes, is a reflection of his own personal landscape preferences for these water features (West 1784, p. 12). References to waterfalls abound in West’s book, and the author recommends the ‘sixteen miles of excellent mountain road’ between Ambleside and Keswick for ‘all the possible variety of cascades, water-falls, and cataracts, [which] are seen in this ride’ (West 1784, p. 77). West informs his readers that this is an experience best enjoyed when the Lakeland streams are swollen by rain, elsewhere warning that the celebrated Lodore Falls beside Derwentwater, ‘has the misfortune . . . to fail entirely in the dry season’ (West 1784, p. 91). Advice of this kind about waterfalls and rainfall is commonly found in guidebooks published ever since (Hudson 2002).

49

Romantic poet William Wordsworth, whose Guide to the Lakes was first published anonymously in 1810, is critical of this view. He wrote, ‘It is generally supposed that waterfalls are scarcely worth being looked at except after much rain, and that, the more swoln the stream, the more fortunate the spectator; but this, however, is true only of large cataracts with sublime accompaniments, and not even these without drawbacks’ (Wordsworth [1810] 1951, p. 135). On a visit to Aysgarth Falls in Yorkshire after heavy rain, the poet’s sister Dorothy felt that, ‘There was too much water for the beauty of the falls’ which lose their characteristic terraced form when in spate (Wordsworth [1897] 1941, p. 181). She was also critical of some of the amenities, such as pavilions and inappropriate planting, introduced at waterfall sites by landowners for the benefit of visitors. Like other tourists, however, she enjoyed the benefits of improved access which ‘pleasure paths’ provided (Hudson 2001). Travellers with serious purposes other than the enjoyment of beautiful scenery, such as agricultural improvement and enquiry into social conditions, were also appreciative of the aesthetics of landscape. Gilpin’s remarks about ‘rational amusement’ from ‘picturesque travel’ while engaged in travel for more important purposes have been noted earlier, and the writings of authors such as Thomas Pennant (1726– 1798) and Arthur Young (1741–1820) well illustrate this approach. Pennant’s journeys in Britain and on the Continent were mainly for the purpose of his scientific and antiquarian research, while Young’s main interests on his travels were agriculture, economics and social statistics. In their travel writings, both men discuss at length matters relating to their primary scholarly and professional concerns, but also remark on the beauty of the countryside through which they travelled. Waterfalls gave them particular pleasure. Among those seen by Pennant on his travels in Scotland are the Falls of Clyde and the Falls of Foyers. The latter are described by Pennant as ‘a vast cataract, in a darksome glen of stupendous depth [where] the foam, like a great cloud of smoke, rises and fills the air’ (Pennant 1771, p. 170). His description was probably the inspiration for the excursion made by Johnson and Boswell to see the celebrated waterfall which, much depleted by drought, was a disappointment to the two men. One of the highlights of Young’s tour through northern England was his visit to High Force. Describing the road to this famous waterfall

Fig. 4. Lodore Rocks – fall and cottage distance (no date) by Joseph Farington (1747–1821). Pen and ink with watercolour over graphite. Yale Center for British Art, Paul Mellon Collection. In this distant view of Lodore Falls in the English Lake District, the artist greatly exaggerates the appearance of the celebrated cascade, which often disappoints visitors with its usually modest flow of water.

50

B. J. HUDSON

WATERFALLS AND THE ROMANTIC TRAVELLER

on the River Tees, Young writes, ‘From the hills around . . . innumerable cascades pour down the rocky clefts, and render every spot elegantly romantic’ (Young 1770a, p. 197). On arrival at High Force, Young is deeply impressed by the sights and sounds of ‘this noble fall’ which he describes in detail; he found ‘The whole scene . . . gloriously romantic’ and ‘truly sublime’ (Young 1770a, pp. 198 –199). It was the sound the falls made rather than their visual aspects that seems to have struck some writers most. Celia Fiennes noted that the ‘greate falls’ on the River Wear at Durham made ‘a murmuring sound acceptable to the people passing’ (Fiennes [1888] 1982, p. 180), while ‘a little fall’ on the Derwent at Chatsworth ‘makes a pretty murmurring [sic] noise’ (Fiennes [1888] 1982, p. 105). William Camden, John Speed and Celia Fiennesall noted two waterfalls that, in those quieter times, could be heard from the Westmorland town of Kendal. Local people used the varying audibility of the falls to foretell future weather (Camden 1610; Speed 1646; Fiennes [1888] 1982). Recalling his tour of Scotland, Samuel Johnson wrote of a stormy night in the Highlands when ‘The wind was loud, the rain was heavy, and the whistling of the blast, the fall of the shower, the rush of the cataracts and the roar of the torrent, made a nobler chorus of rough musick of nature than it had ever been my chance to hear before’ (Johnson [1775] 1944, pp. 143– 144). Johnson’s companion, James Boswell, also noticed the sounds of waterfalls, including one close to a bed chamber in Dunvegan Castle where the old Laird was lulled to sleep by the tumbling stream (Birkbeck Hall 1964, pp. 207 –208). Visitors to the Lake District were enchanted by the sounds as well as the sights they experienced there, waterfalls contributing much to the pleasures of travel in this popular region. Describing his visit to Derwentwater, Young refers to the ‘noise of distant waterfalls heard most gloriously’ (Young 1770b, p. 145), while West quotes another writer who recommends experiencing the lake by night: ‘a walk by still moon-light (at which time the distant waterfalls are heard in all the variety of sound) among these enchanting dales, opens a scene of such delicate beauty, repose and solemnity, as exceeds all description’ (West 1784, p. 113). Later, West describes ‘an Alpine ride’ from Haweswater to Kendal along which ‘The tumult of cataracts and

51

water-falls on all sides, adds much to the solemnity of these tremendous scenes’ (West 1784, p. 160). Approaching a large waterfall, the sound of the tumbling river announces its presence even when rocks and trees at first prevent it from being seen, ‘but the roar is prodigious’ (Young 1770a, p. 198). Words such as ‘tumult’, ‘solemnity’, ‘tremendous’ and ‘prodigious roar’, which were used to describe waterfalls, indicate how strongly waterfalls appealed to a Sublime sensibility. The dangers that threatened visitors to such rugged, wild places also contributed to these powerful emotions. This was Young’s experience at High Force, the approach to which was fraught with difficulty in those days before improved access was provided for tourists. Today’s wheelchair-accessible path and the steps down to the plunge pool at the foot of the falls are the culmination of improvements made to the site over the past two centuries.

The pleasures of hardship Hardship and danger were commonly encountered by travellers in search of the Sublime. Waterfalls, typically found in rugged ravines where tumbled boulders and tangled woodland often hindered progress, were especially challenging for the Romantic traveller. Tourist John Byng (later Viscount Torrington), who visited High Force in 1792, recorded similar experiences; he and his party followed their guide thro’ many hilly, boggy fields, a mile of walk, till we enter’d a little birch wood; when, being anxious to stand beneath the fall, we endured a most fatiguing descent, and a very dangerous crawl at the river’s edge, over great stones, and sometimes up to our knees in water, till we arrived at the very bottom of the fall. – The sweat running from my brow, and a flap of my coat, my only coat, nearly torn off by the bushes (Byng [1792] 1936, p. 70).

Recording his visit to Yorkshire’s Gordale Scar with its lofty cliffs and cascading stream, Thomas Gray wrote of the threat of ‘instant destruction’ which he felt as he stood below the overhanging rocks in order to view the waterfall. It was this which formed ‘the principal horror of the place’ (Gray 1820, p. 374). Samuel Taylor Coleridge described the hardships and dangers he encountered on his waterfalls expeditions in some of his correspondence. The following is an extract from a letter he

Fig. 5. Upper Falls of the Reichenbach: Rainbow (1810) by J. M. W. Turner. Watercolour, graphite, heightening with white and scratching out (27.9 × 39.4 cm). Yale Center for British Art, Paul Mellon Collection. Turner’s paintings of waterfalls were greatly admired by John Ruskin. The Reichenbach Falls are among the scenic attractions that made the Swiss Alpine town of Meiringen a popular tourist resort. Writer Arthur Conan Doyle chose this location for the climactic struggle between Sherlock Holmes and Professor Moriarty.

52

B. J. HUDSON

WATERFALLS AND THE ROMANTIC TRAVELLER

wrote in 1802 to a woman friend about an excursion he made in the Lake District: I . . . climbed up by the waterfall as near as I could, to the very top of the Fell – but it was so craggy – the Crags covered with spongy soaky Moss, and when bare so jagged as to wound one’s hands fearfully – and the Gusts came so sudden & strong, that the going up was slow, & difficult & earnest – & the coming down, not only all that, but likewise extremely dangerous. However, I have always found this stretched and anxious state of mind favourable to depth of pleasurable Impression, in the resting Places and lownding Coves (Coleridge [1802] 1951, pp. 240–241).

Clearly, the difficulties and dangers encountered by Coleridge on his waterfall expeditions heightened his enjoyment of the experience. This and previous quotations reflect the romance attached to the notion of suffering in travel, a topic explored at length by Carl Thompson (2007) in his book The Suffering Traveller and the Romantic Imagination. Thompson discusses how Romantic travellers, such as the poets Lord Byron and William Wordsworth, sought to differentiate themselves from other contemporary tourists, particularly those who engaged in the conventional Grand Tour or merely went in search of the picturesque. For travellers like Byron, dangerous exploits and misadventures such as perilous mountain ascents and dramatic shipwrecks were romantic in the extreme, and many readers shared in their adventures surreptitiously through travel literature and poetry. The connection with Burke’s concept of the Sublime in terms of horror and pain seems obvious. This is well exemplified in Byron’s poem The Falls of Terni, the first verse of which is quoted below: THE ROAR of waters! – from the headlong height Velino cleaves the wave-worn precipice: The fall of waters! rapid as the light The flashing mass foams shaking the abyss; The hell of waters! where they howl and hiss, And boil in endless torture; while the sweat Of their great agony, wrung out from this Their Phlegethon, curls round the rocks of jet That gird the gulf around, in pitiless horror set.

Phrases such as ‘endless torture’ and ‘pitiless horror’ convey the imagery of pain that re-emerges strongly in the last verse which refers to ‘the torture of the scene’. ‘Charming the eye with dread’, the spectacle is summarized as ‘Horribly beautiful!’ (Fig. 6).

53

Waterfall ‘improvements’ and the expansion of tourism Not all Romantic travellers were prepared to undergo hardship in order to enjoy the pleasures of landscape, many seeking to experience the ‘delightful horror’ (Burke 1757, p. 129) or picturesque qualities of waterfalls in relative safety and comfort. This was recognized by Thomas West who informed his readers, ‘What may now be mentioned as another inducement to visit these natural beauties, is the goodness of the roads, which are much improved since Mr. Gray made his tour in 1765, and Mr. Pennant his, in 1772’ (West 1784, p. 2). Later, West advises ‘Whoever would enjoy, with ease and safety, Alpine views, and pastoral scenes in the sublime stile, may have them in this morning ride’ (West 1784, pp. 129– 130). The advantages of travel in England rather than abroad are emphasized by West who observes, ‘If the roads in some places be narrow and difficult, they are at least safe. No villainous banditti haunt the mountains; innocent people live in the dells. Every cottager is narrative of all he knows; and mountain virtue, and pastoral hospitality are found at every farm. This constitutes a pleasing difference betwixt travelling here and on the continent, where every inn-keeper is an extortioner, and every voiturin an imposing rogue’ (West 1784, p. 133). Even the intrepid Arthur Young, who endured danger and discomfort in order to visit High Force, advocated the provision of safer and easier access to places of scenic beauty, including waterfalls. Referring to the Lake District, he wrote of ‘wild romantic spots which command the most delicious scenes, but which cannot be reached without the most perilous difficulty: To such points of view, winding paths should be cut in the rock, and resting places made for the weary traveller: . . . At the bottoms of the rocks also, something of the same nature should be executed for the better viewing of the romantic cascades’ (Young 1770b, p. 156). In response to the increasing demands of tourism, many landowners constructed ‘pleasure paths’ to and through places of scenic beauty, often providing viewing places, seats and other amenities for the visitor. At some falls guides offered their services in expectation of payment, and owners of ‘beauty spots’ began to charge for entry (Hudson 2001). Among the many waterfalls

Fig. 6. Cascade of Terni (c. 1820) after J. M. W. Turner, from a finished sketch by James Hakewill (1778– 1843). Lithograph engraved by J. Landseer. Yale Centre for British Art, Paul Mellon Collection. Alternatively known as Marmore Falls, this waterfall in Umbria, Italy, is an artificial creation resulting from the diversion of the Velino River by Roman engineers in 271 BC. Now harnessed for hydroelectric power generation, the water from the Velino is discharged over the cliff at scheduled times for the benefit of today’s tourists.

54

B. J. HUDSON

commercially exploited in this way were Lodore Falls in the Lake District, Hardraw Force in the Yorkshire Dales and Swallow Falls in North Wales. Today, visitors still have to pay to see these falls that have been popular tourist attractions since the Romantic era. With improvements in transport, especially the advent of the passenger railway, the number of visitors to places of scenic beauty increased rapidly, and tourism boomed in places that had previously been difficult to reach. At this time, ‘Ruskin was a major figure in setting the agenda and directing the gaze, not only for the upper- and middle-class British tourists who played such an important part in the consolidation of international tourism in the age of the railway, but also for the emergent working-class explorers of hills and countryside within Britain in the late nineteenth and early twentieth centuries’ (Hanley & Walton 2010, p. 1). These developments are exemplified by the Yorkshire village of Goathland, the scenic attractions of which were publicized as part of a promotion scheme for the Whitby and Pickering Railway, completed in 1836 (Belcher 1836). Near Goathland are several waterfalls which were prominently featured in the tourist guidebooks for Whitby and district in the nineteenth and twentieth centuries. In a guidebook published in 1860, the author observes ‘Before the times of railroads, these glens were but little known and visited’ (Robinson 1860, p. 234). The Pennine village of Ingleton was similarly affected by the coming of the railway in 1849. A few decades earlier the waterfalls of this district were little known, but one of them was brought to the attention of geologists by John Playfair who wrote ‘This was a cascade, in the river Greta, called Thornton Force . . . the Greta here precipitates itself from a horizontal rock of limestone; and, after a fall of about eighteen to twenty feet, is received in a bason [sic] which it has worked out in primary schistus. This schistus is in beds almost perpendicular; . . . ’ (Playfair 1822, p. 225). Playfair described this unconformity as evidence for the theory that geological processes of the past were the same as those operating today, and that the physical landscape evolved gradually over vast periods of time – that is, James Hutton’s theory of uniformitarianism. Thornton Force has since become a popular tourist attraction in the Yorkshire Dales National Park, the waterfall now famous as one of the geological sites which illustrates the theory of uniformitarianism. By 1861, two rival railway companies had begun to provide passenger services to this place near the confluence of two rivers, each with a series of waterfalls. Access to the falls remained difficult and dangerous, several visitors suffering fatal accidents.

One Victorian guidebook writer remembered visiting the Ingleton waterfalls when ‘In some places it was necessary to swing from tree to tree, and spring with the utmost caution on to projecting bosses of rock, lest a false step should have launched him into some yawning watery gulf, deep below’ (Speight 1892, p. 224). Some visitors had the misfortune to be seriously injured while attempting to negotiate these obstacles, and the Ingleton Improvement Committee was formed in 1884– 1885 to redress the problem. The solution was provided in the form of footpaths and bridges by which ‘the two glens are accessible to even infirm pedestrians . . . and the scenery of them both, which involves a walk of four or five miles, viewed with ease and composure in the course of a summer afternoon’ (Speight 1892, p. 224). Today, the Ingleton Falls Trail remains a popular tourist attraction, the car park close to the Trail entrance occupying the site of a former railway station. Well into the twentieth century and beyond, waterfalls remained popular attractions among tourists who viewed beautiful landscapes with a ‘romantic gaze’, and guidebooks reflected this taste. Echoes of Romanticism are also found in much of the topographical literature of the period, notably in the books of journalist H. V. Morton (1892–1979) (Taylor 2013; Morton, http://www.hvmorton.co. uk/). Another prolific topographical writer was Tom Stephenson (1893–1987), journalist and champion of walkers’ rights in the countryside. Stephenson edited a book entitled Romantic Britain, published in the late 1930s. In the Introduction, Stephenson wrote, ‘Never has there been so much general admiration of our peaceful rural scenes, our mountain grandeur, our rivers, lakes and forests (Stephenson, n.d. p. 7). On the same page there is a single illustration: a photograph of Cauldron Snout, a waterfall on the River Tees. Waterfalls are well represented in this illustrated volume which was republished in 1946. Since then mass tourism has grown enormously, with an impact on the landscape that would have horrified Ruskin. The cars, tourist coaches and crowds that converge on popular waterfall attractions today often detract from the aesthetic pleasures sought by the sensitive lover of landscape beauty. It is a problem that dates as far back as the Romantic age when William Beckford wrote, ‘There are few lovers of nature who do not wish to retire from the crowded terrace to the lonely glade, from the burstings of the horn to the warbling of the thrush, from the rumblings of the carriage to the murmurs of the torrent or the precipitation of the cascade’ (Beckford 1790, p. 208). Nuisances associated with large-scale tourism were not experienced in eighteenth-century Jamaica where Beckford owned extensive sugar plantations. Here the wealthy planter found ‘The

WATERFALLS AND THE ROMANTIC TRAVELLER

cascades, the torrents, the rivers, and the rills . . . enchantingly picturesque’, some of which ‘might eclipse the boasted waters of Schaffhausen, the brilliancy of Pivaˆche, and the gloom of Terni’ (Beckford 1790, p. 12). Circumstances have changed greatly since Beckford’s time, and many of Jamaica’s waterfalls have been heavily exploited as tourism attractions to the extent that these scenic resources have lost much of their appeal for visitors with a love of unspoiled landscape beauty (Hudson 1998, 1999). Like many other countries, Jamaica now exploits its waterfalls for romantic tourism of a different kind. Romantic love is a phenomenon which may appear to have little connection with Romanticism as an artistic, literary and intellectual movement, but ‘romantic’ settings have long appealed to lovers. Waterfall locations are among those which are favoured by couples seeking places in which to share their love (Hudson 2012). This fact is widely exploited by the tourist industry. A 1990s Jamaica Tourist Board pamphlet, Jamaica Honeymoon, contains a section headed ‘Especially for Lovers’ which contains an invitation to ‘play together in the cascading falls that plunge into seven freshwater pools’ on one of the island’s rivers (Jamaica Tourist Board c. 1996). From the Caribbean to the Pacific, waterfall weddings are widely advertised for couples wishing to make their marriage vows in a romantic setting with a cascading stream in the background. Niagara Falls has long been associated with romantic love (Dubinsky 1999). There is even a waterfall guidebook entitled Romance of Waterfalls: Northwest Oregon and Southwest Washington which suggests good locations for kissing (Bloom & Cohen 1998).

Conclusion While amorous dalliance may have been on the minds of some travellers in the age of Romanticism, the sentiments expressed by travel writers and authors of tourist guidebooks when describing waterfalls mainly reflect the cultivated sightseers’ emotional response to the sublime, beautiful and picturesque qualities of the landscape. This newly emerged attitude to scenery owed much to the scientific studies that revealed how natural processes sculpted the Earth’s surface over unimaginable periods of time, and to the published writings of travellers and thinkers of the Enlightenment. In eighteenth-century Europe, the emergence of Romanticism led to a growing appreciation among the social elite of landscape beauty, a taste which later extended to people in the lower ranks of society. The proliferation of waterfall guidebooks around the world over the past three decades suggests that

55

now, more than ever before, tumbling rivers and streams are popular features of the landscape. Whether seen as beautiful, sublime or picturesque, waterfalls became attractions for the romantic traveller prepared to suffer hardship and danger in order to experience the wonders of nature, and for those who sought to enjoy their delights in relative safety and comfort. For romantic travellers of today, those who take aesthetic pleasure in the sights and sounds of nature like their predecessors of the Romantic era, many waterfalls have lost much of their attraction due to excessive development and other problems associated with mass tourism. The challenge for those responsible for the management of landscapes today is to provide access to places of scenic beauty, such as waterfalls, in a way that minimizes development which detracts from the aesthetic experience. For the modern romantic traveller, it may be preferable to avoid waterfalls that are easily accessible and attract large crowds of people who prefer fun and frolic at the falls. Instead, it may be preferable to suffer some discomfort, hardship and even danger in order to reach a remote cascade where quiet contemplation of nature is possible. The valuable advice of editor T. Hose and the helpful and constructive comments of two anonymous reviewers in the preparation of this paper are gratefully acknowledged.

References Andrews, M. 1989. The Search for the Picturesque; Landscape Aesthetics and Tourism in Britain, 1760– 1800. Scolar Press, London. Beckford, W. 1790. A Descriptive Account of the Island of Jamaica. T. & J. Egerton, London, 1. Belcher, H. 1836. Illustrations of the Scenery on the Line of the Whitby and Pickering Railway, in the Northeastern Part of Yorkshire.From Drawings by G. Dodgson. With a Short Description of the District and Undertaking, by Henry Belcher. Longman, Rees, Orme, Brown, Green & Longman, London. Bloom, B. & Cohen, G. 1998. Romance of Waterfalls Northwest Oregon and Southwest Washington. Outdoor Romance Publishing, Portland. Boswell, J. [1764] 1953. In: Frederick, A. Pottle (ed.) Boswell on The Grand Tour: Germany and Switzerland 1764. McGraw Hill Book Co., London. Birkbeck Hall, G. (ed.) 1964. Boswell’s Life of Johnson Together with Boswell’s Tour to The Hebrides and Johnson’s Diary of a Journey to North Wales, in six volumes, Volume V The Tour of The Hebrides and The Journey into North Wales, 2nd edn. Clarendon Press, Oxford. Burke, E. 1757. A Philosophical Enquiry Into the Origin of Our Ideas of the Sublime and Beautiful. R. & J. Dodsley, London. Byng, J. [1792] 1936. In: Andrews, C. B. (ed.) The Torrington Diaries. Eyre & Spottiswoode, London, 3.

56

B. J. HUDSON

Camden, W. 1610. Britain . . . written first in Latine by William Camden . . . translated by Philemon Holland. Georgii Bishop & Ionis Norton, London. Chanca, D. A. 1969. The letter written by Dr. Chanca to the City of Seville. In: Cohen, J. M. (ed.) The Four Voyages of Christopher Columbus. Penguin, Harmondsworth, 129– 157. Chisholm, G. 1885. Rapids and waterfalls. Scottish Geographical Magazine, 1, 401–422. Coleridge, S. T. [1802] 1951. Letter to Sara Hutchinson, Aug/Sep, 1802. In: Coburn, K. (ed.) Inquiring Spirit: a New Presentation of Coleridge from His Published and Unpublished Prose. Routledge & Kegan Paul, London, 240– 242. Defoe, D. 1727. A Tour Thro’ the Whole Island of Great Britain. G. Strahan, W. Mears, & J. Stagg, London, 3. Defoe, D. 1810. The Novels of Daniel De Foe, Volume Eleventh containing A New Voyage Round the World. John Ballantyne & Co. and Brown & Crombie, Edinburgh; Longman, Hurst, Rees & Orme, and John Murray, London. Dubinsky, K. 1999. The Second Greatest Disappointment: Honeymooning and Tourism at Niagara Falls. Rutgers University Press, New Brunswick. Fiennes, C. [1888] 1982. In: Christopher, M. (ed.) The Illustrated Journeys of Celia Fiennes 1685– c.1712. Macdonald & Co, London & Sydney; Webb & Bower: Exeter. Furniss, T. 2010. A Romantic geology: James Hutton’s 1788 ‘Theory of earth’. Romanticism, 16, 305– 321. Gilpin, W. 1782. Observations on the River Wye, and Several Parts of South Wales, &c. Relative Chiefly to Picturesque Beauty. R. Blamire, London. Gilpin, W. 1792. Three Essays; on Picturesque Beauty; on Picturesque Travel; and on Sketching Landscape: To Which is Added a Poem on Landscape Painting. R. Blamire, London. Gilpin, W. 1808. Observations on Several Parts of England, Particularly the Mountains and Lakes of Cumberland and Westmoreland, Relative Chiefly to Picturesque Beauty, Made in 1772, in two volumes. 3rd ed. T. Cadell & W. Davies, London, 1. Gordon, J. E. 2012. Rediscovering a sense of wonder: geoheritage, geotourism and cultural landscape experiences. Geoheritage, 4, 65–77. Gray, T. 1820. The Poems and Letters of Thomas Gray. With Memoirs of His Life and Writings by William Mason, M. A. R. Priestley & W. Clarke, London. Hanley, K. & Walton, J. 2010. Constructing CulturalTourism: John Ruskin and the Tourist Gaze. Channel View Publications, Bristol. Hargrove, E. C. 2008. Foundations of American environmental attitudes. In: Carlson, A. & Lintott, S. (eds) Nature, Aesthetics and Environmentalism: From Beauty to Duty. Columbia University Press, New York & Chichester, 29–48. Heringman, N. 2004. Romantic Rocks: Aesthetic Geology. Cornell University Press, Ithaca & New York. Hose, T. A. 2008. Towards a history of geotourism: definitions, antecedents and the future. In: Burek, C. V. & Prosser, C. (eds) The History of Geoconservations. Geological Society, London, Special Publications, 300, 37–70.

Hose, T. A. 2010. Volcanic tourism in West Coast Scotland. In: Erfurt-Cooper, P. & Cooper, M. (eds) Volcano and Geo-Thermal Tourism; Sustainable Geo-resources for leisure and Recreation. Earthscan, Abingdon & New York, 259 –271. Hose, T. A. 2012. 3G’s for Modern Geotourism. Geoheritage, 4, 7– 24. Hudson, B. J. 1998. Waterfalls: resources for tourism. Annals of Tourism Research, 25, 958 –973. Hudson, B. J. 1999. Fall of Beauty: the story of a Jamaican waterfall – a tragedy in three acts. Tourism Geographies, 1, 343–357. Hudson, B. J. 2000. The experience of waterfalls. Australian Geographical Studies, 38, 71– 84. Hudson, B. J. 2001. Wild Ways and Paths of Pleasure: access to British Waterfalls, 1500– 2000. Landscape Research, 26, 285– 303. Hudson, B. J. 2002. Best after rain: waterfall discharge and the tourist experience. Tourism Geographies, 4, 440–456. Hudson, B. J. 2012. Waterfall: Nature and Culture. Reaktion Books, London. Hudson, B. J. 2013a. Waterfalls, science and aesthetics. Journal of Cultural Geography, 30, 356– 379. Hudson, B. J. 2013b. ‘Lofty hills streaming with waterfalls’: the Tourist Landscape of Johnson and Boswell. In: Rakic, T. & Lester, J. (eds) Travel. Tourism and Art. Ashgate Publishing, Farnham, 35– 46. Humboldt, A. 1814. Researches Concerning the Institutions & Monuments of the Ancient Inhabitants of America, with Descriptions & Views of Some of the Most Striking Scenes in the Cordillieras: Written in French by Alexander de Humboldt, & Translated into English by Helen Maria Williams, Volume 1. Longman, Hurst, Rees, Orme & Brown, J. Murray & H. Colburn, London. Humboldt, A. 1821. Personal Narrative of Travels to the Equinoctial Regions of the New Continent During the years 1799– 1804 by Alexander de Humboldt and AimeBonpland; Written in French by Alexander deHumboldt and Translated into English by Helen Maria Williams, Volume 5. Longman, Hurst, Rees, Orme & Brown, London. Jamaica Tourist Board c. 1996. Jamaica Honeymoon, pamphlet published by the Jamaica Tourist Board, Kingston in the 1990s. Johnson, S. [1774] 1816. In: Duppa, R. (ed: with illustrative notes) A Diary of a Journey Into North Wales in the Year 1774, Robert Jennings, London. Johnson, S. [1775] 1944. In: Chapman, R. W. (ed.) Johnson’s Journey to the Western Islands of Scotland and Boswell’s Journal of a tour to the Hebrides with Samuel Johnson, L.L.D. Oxford University Press, London, New York, Toronto. Leland, J. [1710– 1712] 1910. In: Smith, L. T. (ed.) The Itinerary of John Leland in or about the Years 1553– 1543. George Bell and Sons, London, 5. Livingstone, D. 1858. Missionary Travels and Researches in South Africa: Including a Sketch of Sixteen Years’ Residence in the Interior of Africa, and a Journey From the Cape of Good Hope to Loanda on the West Coast; Thence Across the Continent, Down the River Zambesi, to the Eastern Ocean. Harper & Brothers Publishers, New York.

WATERFALLS AND THE ROMANTIC TRAVELLER Long, E. 1774. The History of Jamaica. T. Lowndes, London, 1. Pennant, T. 1771. A Tour in Scotland; 1769. John Monk, Chester. Playfair, J. 1822. The Works of John Playfair, Esq. with a Memoir of the Author. Archibald Constable & Co., Edinburgh; Hurst Robinson & Co., London, 1. Price, U. 1794. An Essay on the Picturesque, as Compared with the Sublime and the Beautiful. J. Robinson, London. Price, U. 1801. A Dialogue on the Distinct Characters of The Picturesque and The Beautiful. In Answer to the Objections of Mr. Knight. Prefaced by An Introductory Essay on Beauty; With Remarks on The Ideas of Sir Joshua Reynolds and Mr. Burke, Upon That Subject. J. Robson, London. Robinson, J. 1819. A Guide to The Lakes in Cumberland, Westmorland and Lancashire. Lackington, Hughes, Harding, Mayor & Jones, London. Robinson, F. K. 1860. Whitby: its abbey and principal parts of its neighbourhood; or a sketch of the place in its former history and present state, with topography and antiquities of the surrounding country. S. Reed, Whitby. Sillitoe, A. 1995. Leading the Blind: A Century of Guidebook Travel. Macmillan, London. Speed, J. 1646. England, Wales, Scotland and Ireland Described and Abridged with Ye Historic Relation of things worthy Memory from a Farr Larger Volume Done by John Speed. G. Humble, London.

57

Speight, H. 1892. The Craven and North-West Yorkshire Highlands. Elliot Stock, London. Stephenson, T. (ed.) n.d. Romantic Britain. Odhams Press, London. Taylor, N. 2013. H.V. Morton. http://www.hvmorton.co. uk/ (accessed 11 November 2014). Thompson, C. 2007. The Suffering Traveller and The Romantic Imagination. Oxford University Press, Oxford. Urry, J. 1990. The Tourist Gaze; Leisure and Travel in Contemporary Societies. Sage Publications, London. West, T. 1784. A Guide to the Lakes in Cumberland, Westmorland and Lancashire. 3rd edn. B. Law; Richardson & Urquart, London; W. Pennington: London. Wordsworth, D. [1897] 1941. In: De Selincourt, E. (ed.) Journals of Dorothy Wordsworth. Macmillan & Co. Ltd., London, 1. Wordsworth, W. [1810] 1951. A Guide Through the District of The Lakes. Rupert Hart-Davis, London. Wyndham, H. 1775. A Gentleman’s Tour Through Monmouthshire and Wales in the Months of June and July, 1774. T. Evans, London. Young, A. 1770a. A Six Months Tour Through the North of England. W. Strahan, W. Nicoll, London; B. Collins, Salisbury; J. Balfour, Edinburgh, 2. Young, A. 1770b. A Six Months Tour Through the North of England. W. Strahan, W. Nicoll, London; B. Collins, Salisbury; J. Balfour, Edinburgh, 3.

The artist as geotourist: Eugene von Gue´rard and the seminal sites of early volcanic research in Europe and Australia RUTH PULLIN (e-mail: [email protected]) Abstract: The career of the Austrian-born landscape painter Eugene von Gue´rard (1811– 1901) was defined by his travels, which took him to Italy and Germany in the 1830s and 1840s and to Australia and New Zealand between 1852 and 1882. Today he is recognized as one of Australia’s greatest nineteenth-century landscape painters. His formative years coincided with the emergence of geology as an independent scientific discipline and a growing awareness in the wider community of the role played by volcanic activity and other geological processes in the formation of the Earth’s geomorphology. This new understanding was particularly pertinent to landscape painters, whose very subject was the form of the land; in Germany, where von Gue´rard trained and worked between 1838 and 1852, its relevance for landscape painters was emphasized by the influential natural scientist Alexander von Humboldt and the scientist, landscape painter and art theorist Carl Gustav Carus. They argued that the artist should paint from a geologically informed perspective. Von Gue´rard’s interest in volcanic geology was sparked by his experiences in southern Italy, consolidated in Germany on expeditions through the Harz and Eifel regions and then fully realized in response to the landscapes of southeastern Australia. Through his informed portrayal of sites of geological significance in each hemisphere and through the cultural value invested in them as a consequence of his depiction of them, von Gue´rard epitomized that recently conceived construct: the geotourist.

The landscape painter Eugene von Gue´rard (1811– 1901) was an inveterate and adventurous traveller. Throughout his long career he travelled to some of the most significant and spectacular geological sites to be found in Europe and the Antipodes, some already famous and on the tourist trail, and others in remote and inaccessible regions which were still little known. His desire to explore the natural world, fuelled by an apparently inexhaustible intellectual curiosity, far outweighed the dangers and hardships associated with nineteenth-century travel: von Gue´rard was a ‘dedicated’, as distinct from a ‘casual’, geotourist (Hose 2008, pp. 39, 55). He was committed to the idea, advanced by some of the most influential thinkers in Germany in the first half of the nineteenth century, that landscape painters should approach their subject from a scientifically informed perspective. The specifically geological focus of his practice reflected not only the dramatic developments that took place in geological thinking in the early nineteenth century but also their relevance for the wider non-scientific community. On all his travels von Gue´rard carried small pocket-sized sketchbooks, 33 of which are now held by the State Library of New South Wales, Sydney, in which, using a finely sharpened pencil, he made meticulously detailed and accurate sketches of the landscape, its vegetation and its geological features. His subjects included the volcanic landscapes of southern Italy, Germany’s 1

Eifel region and the southwestern plains of the Australian colony of Victoria. He was the exemplary ‘geotourist’. Von Gue´rard’s career as a landscape painter coincided with the emergence of geology as an independent scientific discipline. The educated nonscientific German-speaking community had been introduced to the new geological thinking through the essays of influential figures such as the poet, novelist and dramatist Johann Wolfgang von Goethe (1749– 1832) and the widely read works of the natural scientist Alexander von Humboldt (1769– 1859) (Scott Baldridge 1984, pp. 163–167). In 1847 Humboldt urged artists to travel to the New World to record its landscapes, vegetation and geology in the interests of both art and science (Humboldt [1847] 1849, p. 452). He saw the parallels between his own empirical methodology and that of the artist whose practice was based on the close observation of natural phenomena; in addition, as a scientist, he recognized the potential inherent in an accurate visual portrayal of a subject to convey a wealth of information with enviable directness and immediacy. Von Gue´rard’s journey to Australia in 1852 in search of an ‘unexplored field for study’ was inspired by Humboldt, as were the journeys of a number of contemporary German artists and scientists, notably the eminent botanist Ferdinand von Mueller (1825– 1896) and the geophysicist Georg von Neumayer (1826–1909) (Smith in Nicholls 1877, p. 75).1

[J. Smith], ‘Gue´rard (Von), Jean Eugene’, in Nicholls (1877), p. 75.

From: Hose, T. A. (ed.) 2016. Appreciating Physical Landscapes: Three Hundred Years of Geotourism. Geological Society, London, Special Publications, 417, 59– 69. First published online December 15, 2014, http://dx.doi.org/10.1144/SP417.4 # The Geological Society of London 2016. Publishing disclaimer: www.geolsoc.org.uk/pub_ethics

60

R. PULLIN

Like Humboldt, the German scientist, physician, landscape painter, art theorist and ‘geognosist’ Carl Gustav Carus (1789–1869) argued, in his Nine Letters on Landscape Painting, that art and science were complementary practices, each capable of illuminating the other. For him the very future of landscape painting lay in its connection with science generally, and with geology in particular (Carus [1831] 2002, p. 138). The focus on an informed portrayal of rock types and geological processes in landscape painting was, of course, generated by the radical advances that were made in the field of geological research itself in the early nineteenth century. In Germany the debate between the Vulcanists and the Neptunists over the role of volcanic activity as an agent in the formation of the Earth’s landforms was still heated and it reverberated well beyond scientific circles. This was partly because Vulcanism challenged Christian belief in a way that Neptunism – in part because the latter’s premise of a primeval ocean was reconcilable with the biblical account of the Flood – did not. Similarly, the hitherto unimaginable concept of geological time, the unavoidable corollary of the recognition that volcanic activity and the processes of weathering and uplift were gradual and ongoing, was of profound and widespread relevance. As a landscape painter and theorist, Carus recognized that through the accurate and detailed depiction of geological subjects the geological processes in operation could be inferred and, in this way, the geological history of a site could be suggested. Landscape paintings could take on the significance traditionally associated with history paintings: they could tell the history of the Earth (Busch 1994, pp. 494, 495). Von Gue´rard’s commitment to a scientific understanding of his subject was demonstrated throughout his career. It underpinned his 1843 sketching expedition in the German Eifel region and, on his arrival in Australia, it was evident in his consultation of Major Thomas Mitchell’s (1792–1855) Three Expeditions into the Interior of Eastern Australia. In his early years in the colony of Victoria it was expressed through his membership of the Royal Society of Victoria and his friendships with eminent geologists and scientists of his day, notably Ferdinand von Hochstetter (1829–1884), the NewZealand-based Julius von Haast (1822– 1887),

2

Georg von Neumayer (1826–1909) and Alfred Howitt (1830–1908).2

Von Gue´rard and Vesuvius Von Gue´rard’s career as an artist began in 1826 when, at the age of 15, he and his artist-father, Bernard von Gue´rard (1771– 1836) left Vienna for Italy. They reached Naples in 1832 and, with Vesuvius as a constant and restive presence and on excursions to the nearby Phlegraean Fields and to the islands of Ischia and Sicily, the foundations of von Gue´rard’s lifelong interest in volcanic geology were laid. The place of Naples on the tourist itinerary was already well established when the von Gue´rards arrived. An ever-increasing number of visitors – grand tourists, scientists and artists from northern Europe – had been drawn to Naples throughout the later eighteenth and early nineteenth centuries (Dean 2007, p. 99). The excavation of the ruins of Pompeii and Herculaneum had generated intense excitement and, in the 1770s, publications by Sir William Hamilton (1730–1803) brought Etna, Vesuvius and the Phlegraean Fields to the attention of the English-speaking world.3 German speakers were introduced to the subject through the work of Leopold von Buch (1774–1853) and Alexander von Humboldt, both of whom witnessed the 1822 eruption of Vesuvius and published on the subject.4 Successive generations of artists responded to different aspects of the city and its volcano: eighteenthcentury artists, such as the Englishman Joseph Wright of Derby (1734–1797) and the Frenchman Pierre Jacques Volaire (1729–c. 1802) were inspired by the power of the spectacle of the erupting volcano to evoke the feelings of awe and terror associated with the Sublime. Early nineteenth-century artists such as the German Franz Catel (1778–1856) and the English painter J. M. W. Turner (1775–1851) sought to capture the brilliant light and colour of Naples on paper and canvas. For them, the volcano became an arresting form in their sweeping views of the bay. Some artists, notably the Dane Johan Christian Dahl (1788–1857) and the young Eugene von Gue´rard, began to take a more focused interest in the geology of Vesuvius.5 Two double page drawings of Vesuvius in von Gue´rard’s earliest surviving sketchbook record his

Von Gue´rard accompanied A.W. Howitt on expeditions to Mount Baw Baw (1858) and the Gippsland Alps (1860–61) and Georg von Neumayer to Cape Otway (1862) and Mount Kosciuszko (1862). 3 Sir William Hamilton (1772; 1776). 4 Alexander von Humboldt’s 1828 Tableaux de la Nature, Gide fils, Paris, discussed in Dean, 2007, p. 101. 5 Dahl visited Naples in 1820– 21 and, as the guest of the Danish Crown Prince Christian Frederik, he met experts in many fields, including natural scientists.

THE ARTIST AS GEOTOURIST: VON GUE´RARD

61

Fig. 1. Eugene von Gue´rard Vesuvius during an eruption (Cratere del Vesuvio durante ein eruzione) 1834, fol. 5, Sketchbook III, 1834, Naples and Sicily, pen and ink, coloured wash, pencil, 12.5 × 37.4 cm. Private collection, Coates, UK. Reproduced with the kind permission of the owner.

experience of the volcano, ‘during an eruption’, in 1834. His ascent of the mountain was probably not very different to that described by J. W. Goethe in 1786, in his Italian Journey: ‘At the foot of the steep slope we were met by two guides, one elderly, one youngish, but both competent men. The first took me in charge, the second Tischbein, and they hauled us up the mountain. I say “hauled,” because each guide wears a stout leather thong around his waist: the traveller grabs on to this and is hauled up, at the same time guiding his own feet with the help of a stick. In this manner we reached the flat base from which the cone rises. Facing us in the north was the debris of the Somma. One glance westward over the landscape was like a refreshing bath and the physical pains and fatigue of our climb were forgotten. We then walked round the cone, which was still smoking and ejecting stones and ashes. So long as there was space enough to remain at a safe distance, it was a grand, uplifting spectacle.’ (Goethe [1786] 1962, pp. 183, 184)

Goethe was able to examine the lavas with the knowledgeable older guide who, we are told, ‘could pick them out and give the exact year of each’ (Goethe [1786] 1962, p. 185). Von Gue´rard’s ascent to the Atrio del Cavallo, a semicircular area between the Somma and Vesuvius, was made on a mule or horseback; for the second half of the climb he would have been 6

attached by rope to a guide in order to negotiate the perilous terrain. In the two drawings produced on site, one pen and ink over pencil and the second with the addition of colour, the smoking cone of Vesuvius is viewed from a rugged platform of large and irregular rocks on the Somma. In the second drawing, inscribed Cratere del Vesuvio durante ein Eruzione 1834 (Fig. 1), von Gue´rard’s fellow travellers can be seen moving in pairs and using picks to clamber over giant tephrite-basaltic boulders.6 The sight of the eruption, experienced at such close range, had a profound effect on the artist. The bold, aggressive hatching of the dark foreground rocks conveys a sense of the heightened emotions he experienced as he witnessed (and heard) the roar of the fiery red smoke, ash and rocks spewing from the cone of the volcano. The red wash over the foreground rocks hints at the heat they retained and the word ‘Lava’, inscribed over the rocks, emphatically registers their volcanic origin. The experience was seared into the artist’s imagination and it ignited his sustained interest in volcanic geology.

Von Gue´rard and the Vulkaneifel Von Gue´rard left Naples in May 1838 following the death of his father in the cholera epidemic that swept the city in 1836. He travelled to Du¨sseldorf, the

It is unlikely that von Gue´rard’s visit took place during the major 1834 eruption of 22 and 29 August, an event that caused the destruction of 180 houses and covered 202 hectares (500 acres) of the surrounding land in lava flow. His sketch either predates his departure for Sicily (17 April) or post-dates his return on 14 July. The sequence of drawings in the sketchbooks suggests a February or early March date. If, on the other hand, it dates from late July it would record the May to July period of volcanic activity reported to the Royal Society of London by Charles Daubeny (1835).

62

R. PULLIN

Fig. 2. Eugene von Gue´rard, Old Crater, Mosenberg, near Manderscheid (Alter Krater Mosenberg bei Manderscheid), 1843, fol. 41, Sketchbook XV, 1843, Germany, 1843, 1845, 1850–51, pen and wash, pencil, 12.3 × 22.5 cm. Dixson Galleries, State Library of New South Wales, Sydney, DGB14, vol. 5. Reproduced with the kind permission of the State Library of New South Wales. Von Gue´rard made this closely observed drawing of the Mosenberg Crater while on a two-month sketching tour of Germany’s volcanic Eifel region in September 1843. His expedition, made with a fellow student Georg Saal (1817–70), followed in the tradition of the geologically focused sketching tours that became an essential part of a landscape painter’s training at the Du¨sseldorf Academy in the 1830s and 1840s.

home of his father’s family and also the home of one of the most progressive art academies in Europe. There he studied landscape painting under Johann Wilhelm Schirmer (1807–1863). In 1843 he and a fellow student, Georg Saal (1817–1870), set out on a two-month sketching expedition to the Eifel, an extinct volcanic region located to the south of Du¨sseldorf, to the west of the Rhine and between the Moselle and Ahr rivers. This landscape, with its 74 maar lakes, numerous volcanic cones, exposed basalt cliffs and dolomite outcrops dating back 600 000 years, had become an important site for the fieldwork of German scientists including Leopold von Buch and Alexander von Humboldt; Humboldt visited the Eifel in 1794 and again in 1845. The painters of the Du¨sseldorf School followed in the footsteps of the scientists. In the 1830s Schirmer began leading groups of students on sketching expeditions into the Eifel; his colleague Carl Friedrich Lessing (1808– 1880) made seven expeditions to the area.7 The Eifel landscape became so important to the Du¨sseldorf painters that students in the Academy were formally assessed on their ability to capture its character.8 The two small sketchbooks that von Gue´rard carried on his expedition to the Eifel were the repositories for his penetrating studies of rocks and the 7

spectacular geological features of the region.9 As a result of his systematic practice of dating and documenting his drawings, they also function as travel diaries. Their trip began with a week at Maria Laach, a Romanesque abbey set on the edge of the lake-filled crater, the Laacher See.10 From there they travelled to Koblenz where they boarded a steamer to travel along the Moselle River, from which von Gue´rard sketched the picturesque ruined castles and towers which punctuate the skyline at every bend in the river. They visited the Roman city of Trier and then headed north to their main destination, the small town of Gerolstein (now the centre of the Vulkaneifel European Geopark) in the heart of the Eifel and their base for three weeks of dedicated geological sketching and (geo)tourism. The crater lakes and rock formations sketched by von Gue´rard are today key features of the Geopark. Although tourism on the Rhine was already flourishing in the 1840s and references to some of the volcanic features in the Gerolstein area can be found in Karl Simrock’s 1838 travel guide, Der Rhein, interest in the Eifel as a tourist destination was virtually non-existent (Simrock 1838, pp. 421, 422). In 1843 von Gue´rard and Saal were among the few visitors to this rugged and remote region.

Lessing’s first expedition in 1827 was followed by six subsequent excursions; Schirmer led expeditions to the Eifel in 1831, 1844 and 1845. 8 In his 1843–44 report Schirmer praised von Gue´rard’s classmate Georg Saal for his ability to handle the character of the Eifel. Schu¨lerlisten der Du¨sseldorfer Kunst Akademie, 1843/44, Du¨sseldorf Nordrhein Westfa¨lisches Haupstaatsarchiv; Stefan Horsthemke (1998), p. 168. 9 Eugene von Gue´rard, Sketchbook XV, 1843, Dixson Galleries, State Library of New South Wales, DGB14, vol. 5; Eugene von Gue´rard, Sketchbook XVI, 1843, Dixson Galleries, State Library of New South Wales, DGB14, vol. 6. 10 Maria Laach was already at that time a tourist destination and it featured in contemporary guidebooks.

THE ARTIST AS GEOTOURIST: VON GUE´RARD

In his Nine Letters Carus argued that the precise and accurate study of rocks was as essential to the landscape painter as anatomical studies were to the figure or history painter (Carus [1831] 2002, p. 138). The commitment to geological accuracy that the Du¨sseldorf painters brought to their portrayal of the Eifel landscape was particularly evident in the practice of Carl Friedrich Lessing. He reportedly carried Jakob No¨ggerath’s geological text, Das Gebirge in Rheinland-Westphalen nach mineralogischem und chemischem Bezuge (The mountain range in Rhineland-Westphalia in reference to its mineralogical and chemical basis) (1822–1826) on his 1832 Eifel expedition.11 Schirmer later recalled that Lessing ‘had the structure of the slate impressed on his memory with such geological accuracy, that each piece that he later drew or painted . . . [in the studio] was like a study taken from nature’.12 A similar sensibility informed von Gue´rard’s closely observed pencil and pen and ink studies of the cliffs, rock formations and crater lakes of the Eifel. In the double-page study of the Mosenberg crater lake (Fig. 2), sketched on 23 September 1843, his understanding of the volcanic geology of his subject is evident in his portrayal of the circular lake ringed by a raised tuff embankment which rises to a steeper cone-shaped slag hill on the right, with another rising behind it. The random disposition of basalt fragments and lapilli alludes to the cataclysmic activity involved in the creation of this landscape. Von Gue´rard’s study of the Eifel region placed him in a unique position to respond to the volcanic landscape of Victoria’s southwestern plains in Australia (Pullin 2009, pp. 6–33).

Von Gue´rard in Australia In 1852, motivated by his desire to paint the landscapes of the Australian New World (though the exigencies of the art market in Du¨sseldorf and the lure of the Victorian goldfields were also factors) von Gue´rard made the momentous decision to travel to Australia. Over the next 28 years he travelled into some of ‘the wildest and least-known portions of Australian territory’ (Smith 1877, p. 75). Alexander von Humboldt had urged landscape painters to ‘pass the narrow limits of the Mediterranean’ in order to portray the landscapes of the New World with both ‘scientific accuracy’ and ‘the vivifying breath

11

63

of imagination’ (Humboldt [1845] 1849, p. vii). Whereas most of the artists who responded to Humboldt’s vision for landscape painting travelled to South America, as the scientist himself had done, von Gue´rard became the Humboldtian Reiseku¨nstler (travelling artist) of the Australian New World. In 1855, following 13 adventurous months on the Ballarat gold fields and a year spent in and around Melbourne, von Gue´rard set out on the first of the many intrepid expeditions that characterize his Australian career. Sketches made on day trips from the goldfields reveal that von Gue´rard was aware of the volcanic character of the landscape that stretches across southwestern Victoria to southeastern South Australia from as early as 1853 (Fig. 3). Even so, nothing could have prepared him for the sight of Tower Hill, a spectacular lake-filled maar volcano located to the northwest of Warrnambool in Victoria’s Western District, on 8 August 1855. Tower Hill is one of 40 maar volcanoes, many of which are lake filled, in a young volcanic region that encompasses 15 000 km2 (3706 acres) and 400 individual eruption points, including scoria cones and shield volcanoes. The most recent dating of Mount Gambier has yielded an age of 5000 years for what is probably the youngest volcanic activity in the region; most of the region was built up from lavas that erupted between 2 and 4.5 Ma. In 1836 the explorer Thomas Mitchell recognized the volcanic character of the region he described as ‘Australia Felix’ (Mitchell 1838, p. 333; Joyce 2007, p. 35, 36). James Bonwick, the author of Western District Western Victoria. Its Geography, Geology and Social Condition published in 1858, was also in no doubt as to the volcanic nature of the region. On his visit to Tower Hill in 1857 he compared the volcanic rocks found at the site with the lava types and ash found at Stromboli and Pompeii (Bonwick [1858] 1970, p. 73). By the mid-1850s the Geological Survey of Victoria, under the directorship of Alfred Selwyn (1852– 1868), had not yet reached the Western District and no sustained geological survey of the area had been undertaken (Darragh 2001, p. 387).13 For von Gue´rard the sight of the shallow 3-km-wide crater lake, a nested caldera with an island of scoria cones, must have brought memories of the crater lakes he had studied in the Eifel 12 years earlier flooding back. He immediately understood this landscape. Leopold von Buch’s observation that

Uechtritz, F. von, Blicke in das Du¨sseldorfer Kunst-und Ku¨nstlerleben, 2 vols, Du¨sseldorf, 1839/40, in Baur, O. & Bierende, E. (2000), pp. 113–122; No¨ggerath, Jakob (ed.) (1822–1826). 12 Schirmer, J.W. in Sitt 2000, p. 114. 13 Selwyn published two essays on the geology of Victoria, the first in 1861 and the second in 1866.

64

R. PULLIN

Fig. 3. Major geological features of Kanawinka Geopark. Lewis (2010). Reproduced with the kind permission of Ian Lewis.

‘There is nothing in the world like the Eifel; it will become a guide and a teacher for understanding many other areas and knowledge of it cannot be avoided if one wishes to obtain a clear view of the volcanic phenomena on the continents’14 was, it seems, as relevant to the artist as to the geologist. James Dawson, whose property Kangatong was not far from Tower Hill, recognized the uniqueness of the site; he campaigned for its preservation but, in the long term, nothing he did was as effective to this end as his commission for a painting of the subject by von Gue´rard (Fig. 4). In the 1960s von Gue´rard’s accurate and microscopically detailed portrayal of Tower Hill, in particular the diverse vegetation supported by the fertile volcanic ecosystem of the crater, was used as a template for the re-vegetation of the site following the devastation caused by a century of clearing, grazing, mining for scoria and the destructive impact of introduced plant and animal species.15 Although Dawson recognized the value the painting would have ‘for future 14

generations’, he could not have imagined the role it would play in the physical rehabilitation of the site (Dawson in Bonyhady 2000, p. 343). Von Gue´rard’s view of the lake, its islands of scoria cones and the coastal landscape beyond, was taken from a high vantage point on the rim of the crater. Today a large scoria boulder, into which a reproduction of the painting is set, stands at the site of von Gue´rard’s vantage point: tourists are invited to make a comparison between the physical landscape and the artist’s portrayal of it and to bear witness to the role played by von Gue´rard’s painting in the reclamation of this beautiful and environmentally significant site.

Von Gue´rard and Ferdinand von Hochstetter The sketchbooks von Gue´rard used on his expeditions through the volcanic western plains are

Leopold von Buch (1774– 1853) quoted in GeoLife: The Magazine of the Vulkaneifel European Geopark 1, 2001, p. 4. The restoration was led by Max Downes and undertaken by the Fisheries and Wildlife Department, Victoria.

15

THE ARTIST AS GEOTOURIST: VON GUE´RARD

65

Fig. 4. Eugene von Gue´rard, artist, Thomas McLean, printer, View of Koroit or Tower Hill, extinct volcano between Lady Bay & Port Fairy in Australia, c. 1860, chromolithograph on paper, image 39.8 × 70.6 cm, Ballarat Art Gallery. Reproduced with the kind permission of the Ballarat Art Gallery. Von Gue´rard’s lithograph closely follows the composition of his 1855 painting of Tower Hill, a lake-filled maar volcano with a central island of scoria cones in Victoria’s Western District. The artist’s recognition of the geological significance of this then little known volcanic landscape, and the accuracy of his portrayal of it, reflects his study of crater lakes in Germany’s volcanic Eifel region.

filled with drawings of crater lakes, the interiors of craters and the distinctive forms of the scoria cones that rise from the flat basalt plains; these forms served, then as now, as landmarks for the traveller (Fig. 5).16 On these expeditions he also produced a series of larger drawings. In October 1859 the eminent German geologist Ferdinand von Hochstetter (1829– 1884), who had travelled to the southern hemisphere with the Austrian Novara expedition, visited Melbourne on his return journey to Vienna. Time prohibited a trip to the Western District but he had the opportunity to meet von Gue´rard and see his studies and drawings of the region. These ‘splendid drawings’, observed Hochstetter, ‘have given me better information about them [complete tuff and eruption crater forms] than all I could ascertain from the geologists’ (Hochstetter in Darragh 2001, p. 411). On the basis of von Gue´rard’s drawings, and also his discussions with Selwyn and his colleague Georg Ulrich (1830–1900), Hochstetter was able to compile a comprehensive and descriptive list of volcanic 16 17

formations of western Victoria; his reference to the ‘genuine volcanic bomb’ that von Gue´rard had collected on one of his expeditions provides further evidence of the artist’s interest in geology (Hochstetter in Darragh 2001, p. 411).17

‘Euge`ne von Gue´rard’s Australian Landscapes’ Von Gue´rard’s interest in the geology of the southeastern colonies of Australia and New Zealand’s South Island was not restricted to volcanic subjects. He brought the same level of scientific exactitude to his portrayal of subjects such the stratified sandstone cliffs of the Blue Mountains, a glacial cirque on the Kosciuszko massif and the walls of granite, gneiss and diorite of New Zealand’s glacial Milford Sound. The geological range and scientific ambition behind von Gue´rard’s Australian career was succinctly and comprehensively expressed in his album of colour lithographs, Euge`ne von

Mount Elephant is visible from a distance of 60 km. Volcanic ‘bombs’ are rocks formed from molten pieces of magma, which are shaped as they cool during flight after being ejected from the volcano.

66

R. PULLIN

Fig. 5. Eugene von Gue´rard, The Basin Banks, near Camperdown, (also known as Lake Gnotuk), 1857, oil on canvas, 35.2 × 56.7 cm, Ballarat Art Gallery, Victoria. Gift of Lady Currie in memory of her husband, the late Sir Alan Currie, 1949. Reproduced with the kind permission of the Ballarat Art Gallery. Lake Gnotuk, in western Victoria, is one of a pair of crater lakes that was known as the Basin Banks in the nineteenth century. This work is one of a number of paintings, drawings and lithographs of Victoria’s crater lakes that von Gue´rard produced in the years after he painted Tower Hill in 1855. His interest in the volcanic features of the region is evident in his accurate depiction of the distant scoria cones that rise from the flat basalt plains, from the left Mt Meningoort, Mt Hamilton, Mts Koang and Kurtweeton (the Cloven Hills), Mt Elephant and to the far right Mt Myrtoon.

Gue´rard’s Australian Landscapes, published by Hamel & Ferguson in Melbourne between 1866 and 1868 (Fig. 6). The 24 images, selected from the ‘hundreds of drawings suitable for publication’, constituted a distillation of what were, in the artist’s estimation, the most significant aspects of the Australian landscape.18 Fourteen of the lithographs in the album have a specifically geological focus. Through his book of lithographs, von Gue´rard was able to present geological and botanical information about the New World to the Old World. In 1870 von Gue´rard sent two copies of his Australian Landscapes to Hochstetter in Vienna. One was presented to the Austrian Emperor Franz Josef and the other was for Hochstetter himself.19 In an address on von Gue´rard’s Australian Landscapes presented at the Kaiserlich-ko¨niglichen

18

Geographischen Gesellschaft (Royal Geographical Society) in the same year, Hochstetter drew the attention of his audience to the ‘precipitous syenitic rock massif’ identifiable in von Gue´rard’s lithograph of Kosciusko and the ‘sheer rock formations showing the powerful horizontal strata of sandstone’ of the Weatherboard Falls in the Blue Mountains.20 He compared the granite valleys of Launceston’s Cataract Gorge with the Black Forest or the Riesengebirges in Germany and the imposing basalt (dolerite) columns of Tasman’s Island with the Hebridean Island of Staffa. He commented on ‘the small lake filled craters of extinct volcanoes in southern Australia’, represented in von Gue´rard’s album by the crater lakes at Mount Gambier in South Australia and Mount Eccles in Victoria (Builth 2010, p. 80).21 In a letter accompanying the album

Von Gue´rard, letter to Ferdinand von Hochstetter (1870), in Heger, F. von (1884), p. 156. In recognition of his achievement von Gue´rard was awarded the Order of Franz Josef, Knight’s Cross in 1870. 20 The Weatherboard Falls are now known as the Wentworth Falls. Charles Darwin visited this site in 1836. 21 Hochstetter (1870) in Heger (1884), p. 157. 19

THE ARTIST AS GEOTOURIST: VON GUE´RARD

67

Fig. 6. Eugene von Gue´rard Crater of Mount Gambier, South Australia 1867. Plate 11 from Euge`ne von Gue´rard’s Australian Landscapes, published by Hamel and Ferguson, Melbourne (1866–68). Chromolithograph, 48.5 × 67.1 cm (page). National Gallery of Victoria. From his elevated vantage point von Gue´rard was able to see and portray the three largest crater lakes at Mount Gambier– Browne’s Lake in the foreground, Valley Lake in the centre and the famous Blue Lake with its steep embankments in the distance. The letterpress that accompanied his lithograph reads: Of the comparatively limited number of extinct volcanoes in Australia, . . . [Mt. Gambier] is, perhaps, the most remarkable and, to geologists, the most interesting. It is not so much a crater, as a series of craters, about six miles in circumference, filled with fresh water, and walled in by precipitous rocks of black lava, brown ash, and coralline. These walls vary in height from 200 to 300 feet, and are sprinkled with shea-oak, tea-tree and honeysuckle. The Blue Lake crater, which is the most distant from the spectator, is about 240 feet deep, and has obtained its name from the exquisite azure tint of the water, which is as clear as crystal.

sent to Hochstetter von Gue´rard reflected, rather poignantly, on the experience of the travelling artist and pioneering geotourist: A glance at the maps of southeastern Australia will show you how far apart the sites of the given pictures are, and how I have battled to attain this collection. I had to put thousands of miles behind me on horseback, on foot and over the water, defeat troubles of every kind, endure many months of privation in the wilderness, to unite those few sheets in one volume, which now can be leafed through in a few minutes in a drawing room. (Von Gue´rard 1870)22

Informed travellers and artists had already frequented, studied and depicted some of the sites of geological significance sketched and painted by von Gue´rard; others were as yet little known to Europeans. The artist was, on occasion, one of the first 22 23

Europeans to see a particular site, to comprehend its geological significance and the first to portray it. The geological knowledge he had acquired through his experiences in Europe enabled him to see and understand the geology of the ‘new’ landscape of Australia. The most striking example of this was the link he was able to make between the volcanic landscape of the German Eifel and that of Victoria’s western plains. His recognition in the 1850s of the parallels between these two landscapes on opposite sides of the globe was echoed in the twenty-first century with UNESCO accreditation of the Vulkaneifel European Geopark in 2001 and the Kanawinka Global Geopark in 2007 (Fig. 3) (Joyce 2010, p. 310).23 Von Gue´rard’s landscape painting practice was shaped by the profound shifts that took place in

Von Gue´rard, Letter to Hochstetter (1870) in Heger (1884), p. 156. http://www.kanawinkageopark.org.au/; http://www.geopark-vulkaneifel.de/index.php/en/

68

R. PULLIN

geological thinking in the early nineteenth century. His informed and enduringly beautiful portrayal of sites of geological significance in Australia’s southeast has endowed those places with cultural significance, embedding them in the national cultural imagination. Many of the sites that von Gue´rard recognized for their geological significance are now protected within Victoria’s national parks network and some, notably Tower Hill, owe their preservation, at least in part, to the vision of this landscape painter. His legacy lies not only in the paintings that have earned him a reputation as one of Australia’s greatest nineteenth century artists but in the focus his work has brought to the physical landscapes through which he travelled – as an informed and dedicated geotourist.

Postscript 1 The scientific premise that informed von Gue´rard’s response to the landscape and its geology was the central theme of the 2011 touring exhibition of the National Gallery of Victoria, Eugene von Gue´rard: nature revealed, curated by Ruth Pullin (with Michael Varcoe-Cocks). In the accompanying publication (Pullin 2011), von Gue´rard’s interest in Earth sciences is reflected in the contributions of contemporary geologists, geological historians and botanical historians who were invited to write alongside curators and art historians; the aim was to illuminate both von Gue´rard’s achievements as an artist and the physical landscapes that inspired one of Australia’s greatest landscape painters.

Postscript 2 On 1 January 2013, the Kanawinka Global Geopark lost its global listing after UNESCO did not receive approval from the Australian Government for its reaccreditation, to be undertaken in August 2012. The Kanawinka Geopark committee is working to have its global status reinstated. I would like to thank B. (E.B.) Joyce, Professorial Associate and Honorary Principal Fellow, School of Earth Sciences, The University of Melbourne and A. M. Robinson, Managing Partner of Leisure Solutions for their encouragement and support, and I. D. Lewis, Director, Kanawinka Geopark for his generous assistance.

References Baur, O. & Bierende, E. 2000. Lessing als Zeichner der Vulkaneifel. In: Sitt, M. (ed.) Carl Friedrich Lessing: Romantiker und Rebell. Donat Verlag, Bremen, 113–122.

Bonwick, J. [1858] 1970. Western Victoria. Its Geography, Geology and Social Condition. Heinemann Australia, Melbourne. Bonyhady, T. 2000. The Colonial Earth. The Miegunyah Press, Carlton, Victoria. Busch, W. 1994. Der Berg als Gegenstand von Naturwissenschaft und Kunst. Zu Goethes geologischem Begriff. In: Schulze, S. (ed.) Goethe und die Kunst. Gerd Hatje, Stuttgart, 494. Builth, H. 2010. The cultural and environmental landscape of the Mount Eccles lava flow. In: Byrne, L., Edquist, H. & Vaughan, L. (eds) Designing Place. An Archeology of the Western District. Royal Melbourne Institute of Technology, Melbourne Books, Melbourne, 78–91. Carus, C. [1831] 2002. Nine Letters on Landscape Painting, Written in the Years 1815–1824, with a Letter from Goethe by Way of Introduction. Translated by David Britt, Getty Research Institute, Los Angeles. Darragh, T. A. 2001. Ferdinand Hochstetter’s notes of a visit to Australia and a Tour of the Victorian Goldfields in 1859. Historical Records of Australian Science, 13, 383–437. Daubeny, C. 1835. Some Account of the Eruption of Vesuvius, Which Occurred in the month of August 1834, Extracted from the Manuscript Notes of the Cavaliere Monticelli, Foreign Member of the Geological Society, and from Other Sources; Together with a Statement of the Products of the Eruption, and of the Condition of the Volcano Subsequently to It. Philosophical Transactions of the Royal Society of London, 125, 154. Dean, D. R. 2007. J.D. Forbes and Naples. In: Wyse Jackson, P. N. (ed.) Four Centuries of Geological Travel: The Search for Knowledge on Foot, Bicycle, Sledge and Camel. Geological Society, London, Special Publications, 287, 97–107. Goethe, J. W. 1962. Italian Journey 1786–1788. Translated by W.H. Auden and Elizabeth Mayer. Collins, London. Hamilton, W. Sir. 1772. Observations of Mount Etna, and other volcanoes of the two Sicilies, London. Hamilton, W. Sir. 1776. Campi Phlegraei, Naples. Heger, F. Von. 1884. ‘Ferdinand von Hochstetter’, Mitteilungen der Kaiserlich-Ko¨niglichen Geographischen Gesellschaft, 27. Horsthemke, S. 1998. Saal, Georg Eduard Otto. In: Paffrath, H., Sitt, M., Baumga¨rtel, B., Roth, C., Schroyen, S. & Weiss, S. (eds) Lexikon der Du¨sseldorfer Malerschule. 3 vols, Bruckmann KG, Munich, 168. Hose, T. A. 2008. Towards a History of Geotourism: Definitions, Antecedents and the Future. Geological Society, London, Special Publications, 300. Humboldt, A. Von. [1845] 1849. Cosmos. A Sketch of Ta Physical Description of the Universe. Translated by E.C. Otte´, Henry G. Bohn, London, 1. Humboldt, A. Von. [1847] 1849. Cosmos. A Sketch of a Physical Description of the Universe. Translated by E. C. Otte´, Henry G. Bohn, London, 2. Joyce, E. B. 2007. The young volcanic features of Australia: from first recognition to growing understanding between 1853 and 1903. In: Pierson, R. R. (ed.) The History of Geology in the Second Half of the Nineteenth Century: The Story in Australia, and in Victoria, from Selwyn to McCoy—1853 to 1903. Earth Sciences

THE ARTIST AS GEOTOURIST: VON GUE´RARD History Group Conference, Special Publications, Earth Sciences History Group, GSA Inc., Melbourne, Victoria, 1, 35–40. Joyce, E. B. 2010. Volcano tourism in the New Kanawinka Global Geopark of Victoria and SE South Australia. In: Cooper, M. & Erfurt-Cooper, P. J. (eds) Volcano and Geothermal Tourism: Sustainable Geo-Resources for Leisure and Recreation. Taylor & Francis, London, Chapter 20. Lewis, I. D. 2010. Kanawinka, Australia. In: Dowling, R. & Newsome, D. (eds) Global Geotourism Perspectives. Goodfellow Publishers, Oxford, 192– 214. Mitchell, T. L. 1838. Three Expeditions into the Interior of Eastern Australia. T. & W. Boone, London, 2 vols. Nicholls, C. F. 1877. Victorian Men of the Time. McCarron Bird, Melbourne.

69

No¨ggerath, J. (ed.) 1822–1826. Das Gebirge in Rheinland-Westphalen nach mineralogischem und chemischen Bezuge, 4 vols., Bonn. Pullin, R. 2009. The Vulkaneifel and Victoria’s Western District: Eugene von Gue´rard and the Geognostic Landscape. Marshall, D. (ed.) Europe and Australia: Melbourne Art Journal, The Fine Arts Network, The University of Melbourne, Melbourne, Victoria, 11–12, 6– 33. Pullin, R. 2011. Eugene von Gue´rard. Nature Revealed. National Gallery of Victoria, Melbourne. Scott Baldridge, W. 1984. The geological writings of Goethe: Despite his keen powers of observation, Goethe’s ideas on geology reflected the biases of his time. American Scientist, 72, 163 –167. Simrock, K. 1838– 40. Der Rhein. G. Wigand, Leipzig.

Landscape and geotourism on the Dutch coast in the seventeenth century as depicted by landscape artists J. A. M. (HANNEKE) VAN DEN ANCKER1* & PIETER DIRK JUNGERIUS1,2 1

Stichting Geomorfologie & Landschap, Oude Bennekomseweg 31, 6717 LM Ede, The Netherlands

2

University of Amsterdam – Institute for Biodiversity and Ecosystem Dynamics (IBED), PO Box 94248, 1090 GE Amsterdam, The Netherlands *Corresponding author (e-mail: [email protected]) Abstract: The first evidence of tourism on the Dutch coast can be found in drawings and etchings from the end of the sixteenth century. In this period Holland developed into one of the most urbanized regions of Europe. The interest in landscape originated in the towns. The first scenes depicted are those of mass tourism on the beach: sensation-mongers drawn to the beach by whales, sailing cars and departing kings and queens. Somewhat later the dune landscape became a main recreational focus, in which the physical aspects of the landscape were also appreciated. Around the town of Haarlem, individuals and small groups of people started exploring the dune landscape. In the wake of this new interest, landscape painting developed as an artistic genre. It became the most popular genre in the first half of the seventeenth century, and Haarlem developed as a centre for landscape painters. This paper discusses geomorphological features and geotourism engagement as depicted in several of the early etchings and landscape paintings.

The setting During the last quarter of the sixteenth century the Netherlands’ province of Holland developed into the most urbanized region of Europe, with some two-thirds of its inhabitants living in cities. In 1600, the total number of inhabitants in the Netherlands was about 1.5 million. Fifty years later the number of inhabitants of the combined cities of Amsterdam, Haarlem, Leiden, Rotterdam and The Hague (see Fig. 1) had grown to more than 250 000. Amsterdam alone doubled its population in the next 50 years to more than 200 000. Travellers in those days wondered about the many towns in Holland, and reflected whether Holland should not be regarded as just one great city with open spaces in between the population centres, as a building was to be seen almost everywhere in the countryside. More than three centuries later, with the Netherlands’ population at about 17 million and 5.5 million people living in Holland, this question is still under discussion.1 The rapid urban growth of seventeenth-century Holland was a spin-off of the wealth created by the trade on the Baltic Sea, mainly in grain. In its 1

wake a new bourgeoisie developed with a growing interest in the arts and sciences. Also, the awakening interest in landscape and landscape paintings was rooted in city life, perhaps in a search for a lost rural idyll. By the middle of the seventeenth century, paintings of landscapes and rural scenes became the most frequent subject of Dutch art (Leeflang 1998, 2002). Indeed, several authors have pointed out that there probably exists a relationship between the interest in rural landscapes and the agricultural roots of the population in the new towns (Sitt & Biesboer 2002). A similar situation can be seen in nineteenth-century and even late twentieth-century England in ‘neoromanticism’ and ‘ruralism’ (T. A. Hose pers. comm., 2015; Hose 2012).

Accuracy of geomorphological feature depiction When using drawings or paintings for information, there always is the question of the accuracy of the depiction of sites and events. General experience suggests that drawings and prints are

It regularly turns up in newspaper articles and local government debates, one example being the option of making the areas in between The Hague, Haarlem, Amsterdam and Utrecht into one great city separated by green areas.

From: Hose, T. A. (ed.) 2016. Appreciating Physical Landscapes: Three Hundred Years of Geotourism. Geological Society, London, Special Publications, 417, 71– 82. First published online October 14, 2015, http://dx.doi.org/10.1144/SP417.14 # 2016 The Author(s). Published by The Geological Society of London. All rights reserved. For permissions: http://www.geolsoc.org.uk/permissions. Publishing disclaimer: www.geolsoc.org.uk/pub_ethics

72

J. A. M. VAN DEN ANCKER & P. D. JUNGERIUS

Fig. 1. Map showing the main locations mentioned in the text.

more accurate than paintings, because they reflect a first-hand experience (H. Leeflang pers. comm., 2003). They were produced shortly after the events occurred, with the intention of selling prints to a larger public. Successful prints were reprinted for many decades and were often copied by other printmakers, sometimes 50 or more years after the event, sometimes meticulously, but sometimes more freely interpreted. Paintings usually appeared months, or occasionally years, after an event had taken place. They were produced in studios, and most were assignments for which the artist also had to comply with the commissioner’s wishes.

The authors of this paper can see no reason why the artists might invent the landscapes they drew, etched or painted. First of all, because they and their public were interested in the real events and landscapes and because no other means existed in their time to show real-life images. Second, while studying the prints and paintings the authors noticed that the geomorphological features depicted appear in their correct positions and are similar to present-day landscapes and processes. It seems very unlikely that the artists were able to draw imaginary but scientifically correct landscapes. When we order modern artists’ impressions of landscapes

FIRST EVIDENCE OF GEOTOURISM ON THE DUTCH COAST

for educational leaflets explaining geological features, their difficulty in correctly drawing geomorphological features is noteworthy. For example the inclination or curvature of slopes is often wrong and needs to be corrected. These combined observations make a strong case for the landscape artists of the period under consideration attempting to depict the landscape accurately, although of course there are exceptions.2 Similar findings supporting this supposition have been reported from England, where research was undertaken on coastal erosion, as depicted in paintings of the eighteenth and nineteenth century, on the North Sea coast of Norfolk and Suffolk (McInnes & Stubbings 2010). The authors even believe landscape paintings in general might give a better impression of a particular landscape than most photographs. A photograph of an impressive panoramic view commonly tends to show a narrow line in between a huge sky and an unimpressive foreground; it does not in the least correspond with the impressive view experienced by the photographer when the image was captured. Although photography is accepted as a scientifically precise medium, photographic images often do not document the landscape in the way we perceive it. Landscape painters and professional photographers use techniques to overcome this problem. Painters might have moved or added items such as trees to help create depth or painted features that they wanted to draw attention to slightly bigger than their actual proportional size. An illustration of this difference is given by Ossing & Brauer (2006), who compared a seventeenth-century landscape painting by Jacob Ruisdael of Ootmarsum (1628–1682) with modern photographs; they noted that the geomorphology and the sky on the landscape painting are correctly depicted, but that Ruisdael made the buildings slightly larger and altered their positions.

Depictions of geotourism in Holland From the dozens of sixteenth and seventeenth century prints showing the beach and the hundreds of prints and paintings of the dunes collected by the Rijksmuseum, Amsterdam we selected four prints 2

73

and two paintings that present different types of tourists. For each of the prints of the beach extra information is added on how the specific type of tourism shown links to the geological setting. Most prints and paintings show ‘casual’ geotourists, implicitly interested in the landscape, but those walking in the dunes have a real interest in the landscape. Furthermore information is given about the artist and the event and geology depicted.

Mass tourism on the beach The first drawings and etchings of landscapes were made of special events on the beaches. They show sensation-mongers who flocked to the beach to witness, for example the stranding of a whale (see Fig. 2), the experimental sailing car of Simon Stevin (see Fig. 3), the departures of kings and queens (see Fig. 4), shipwrecks, and the aftermath of storms. As these events can only be witnessed on the beach, this makes that geological setting part of their tourism experience. The drawings and etchings are mainly focussed around the towns of Scheveningen and Zandvoort. Their beaches are now the two largest sea resorts of the Netherlands and sites of modern mass tourism. On sunny days these beaches are crowded with bathers who interact with the tidal forces when building their castles of wet sand or use the pools behind the sand banks at low tide as bathing pools for toddlers. Today, as presumably in the past, on every day of the week people walk these beaches to enjoy the view of the sea, the active processes of waves and tides, and the wind blowing sand over the beach during stormy weather, or they go there to watch the cliffs and mass movements in the foredunes that form after a storm.

Whale strandings Whale strandings were depicted by several artists in the sixteenth and seventeenth centuries; for example Hendrik Goltzius’s (1558–1617) Stranded whale at Zandvoort of 1594 and a second drawing he made of a sperm whale beached at Berkhey in 1598 (Jungerius 1997). The drawings and etchings show the public lined up to see and even climb upon the dead whale, as well as the setting – the beach

Apart from real-life scenes, landscape painters also made compositions of scenes of various locations. The making of compositions was promoted, for example, by the famous Dutch Romantic landscape painter Barend Cornelis Koekkoek (1803–1862), founder of the Drawing Academy of Cleves (Germany), near the Dutch border. In the study book he wrote, Herinneringen en Mededeelingen van eenen Landschapsschilder (Recollections and communications of a landscape painter), he advised his students to keep as close as possible to nature. Natura artis magistra. Paintings consisting of landscape compositions can be detected by geomorphologists with knowledge of these landscapes. Furthermore, we have to keep in mind that modern landscape photographers also present us with selections from a landscape.

74

J. A. M. VAN DEN ANCKER & P. D. JUNGERIUS

and foredune. The authors selected a drawing of Goltzius reproduced by Matham (see Fig. 2) in which, besides the whale, the foredune is shown as a cliff – evidence of a past storm event. The foredune also shows a regular indentation. The regular pattern of indentation does not seem to fit a formation due to trampling and erosion along visitors’ footpaths. It might have been formed by storms gradually cutting away the tails of the parabolic dunes. These parabolic dunes, the so-called new dunes, formed along the Dutch coast between the eleventh and fourteenth centuries (Jelgersma et al. 1970).

The sailing car of Simon Stevin The experimental sailing car (or sand yacht) of Simon Stevin (1548–1620) was a vehicle that needed a flat beach, together with a hard and moist sand surface after high tide to support the wheels, to reach its maximum speed of 48 km/h (30 mph). The vehicle was ordered by his Royal Highness Prince Maurits (1567– 1625), who rode it himself. The engraving (see Fig. 3, which is part of the engraving) shows people walking along the beach and halting to witness the car pass, as well as people having especially gone there for the occasion.

Royal arrivals and departures Other events that attracted many visitors to the beach were the arrivals and departures of Europe’s kings and queens. A good example is shown in a print of the departure of the restored monarch King

3

Charles II (1630–1685) of England in 1660 from the beach of Scheveningen (see Fig. 4).3 The royal departure seemed to have attracted an incredible number for the time of more than 50 000 spectators sitting on the dune tops to get a good view (Scott 1907), and an eyewitness recorded ‘Never have more people been seen together in Holland’.4

The sea-way Several seventeenth-century etchings show the more than 1-kilometre-long level drive, the ‘sea-way’ of 1665 connecting The Hague with Scheveningen. Figure 5 shows the gate and toll house and the paved road ending on the beach in Scheveningen. It was designed and promoted by Sir Constantijn Huygens (1596–1687), a diplomat and secretary to the House of Orange and the most famous poet of this period. His friends included Rene´ Descartes and Galileo Galilei. He was father to the scientist Constantijn Huygens, Jr (1628–1697) and the famous astronomer and horologist Christiaan Huygens (1629–1695). To promote the sea-way Sir Constantijn Huygens wrote a long poem describing the sea-way’s objective and construction.5,6 The sea-way was mainly meant for recreational purposes; visitors could from then on comfortably travel to the seaside by carriages or on foot.5,6 The building of the paved street must have been a considerable investment in terms of money and labour. Dunes had to be levelled for its construction. Its creation implies that going to the beach was not something restricted to the special occasions reflected in the previously noted drawings and etchings, but must have been a regular public

King Charles II was returning to England to take up his reign after an exile in France and a stay in The Hague with his sister Mary, who was married to William II of Orange (1626– 1650). 4 Although the Dutch government gave Charles several gifts on the occasion, trying to establish good relationships between the two countries, it did not restrain Charles II from starting two wars against the wealthy Dutch Provinces! It provides an early example of Britain’s several centuries of trade wars with European competitors (T. A. Hose pers. comm., 2015). 5 Sir Constantijn Huygens in the seventeenth century already realized the recreational possibilities of the beach, and he published a detailed plan for the construction of a ‘Zee-straet’ (1653), a straight and paved road connecting The Hague and Scheveningen beach. It took the government almost 10 years to decide about its construction. When the Zee-straet was ready Huygens wrote his poem De Zee-straet van’s Graven-hage op Schevening (1667). In more than 1000 lines he describes the process of its construction and the three reasons for building the road: publicity for The Hague and for the country, convenience and vermaak (in that period the Dutch word for recreation/tourism). In this poem he describes events that could be watched on the beach: the stranding of big fish (whales), ships passing by (strangely enough he does not mention the fishing boats on the beach), shipwrecks and sailing cars; and things to do on the beach: walking (with the girls), collecting shells, playing sports and games and eating fresh fish in one of the many new restaurants along the road. The poem can be downloaded from: http://www.dbnl.org/tekst/huyg001zees01_01/huyg001zees01_01_0002.php, Edition L. Strengholt, 1667 (in Dutch). 6 During an inquiry with the help of students in the 1980s, the authors learned that most modern tourists still prefer to walk to the beach on such straight paved ways. A similar situation is found at other European coastal locations with casual geotourism provision (T. A. Hose pers. comm., 2015).

FIRST EVIDENCE OF GEOTOURISM ON THE DUTCH COAST

outing, which is confirmed by the archives of the town of Scheveningen.

The development of landscape tourism in the dunes around Haarlem Appreciation of the Dutch dune landscape started in the early seventeenth century around the city of Haarlem. Haarlem in those days was considered to have the most beautiful and cleanest surroundings in the Republic of the Seven United Netherlands (1588–1795). Moreover, the Dutch countryside was one of the first in mainland Europe where it was relatively safe to walk, as bandits were under control. Furthermore, painters no longer only travelled to Italy as part of their professional training to see the landscapes depicted by earlier notable painters; they also drew their inspiration from the local Dutch landscapes (Leeflang 1998, 2002). The first dune tourists can be divided into two groups (Leeflang 1998): (1)

Small groups, mostly youngsters, picnicking, drinking wine, reciting poems and singing songs praising the beauties of the surrounding dune landscape. An example of such a poem, describing the new interest for the local dune landscape, was written by Karel van Mander7 (1548–1606): Waar wil ik reysen doch, als ick my wel bedinck, mijn Helikon zij slechts voortaan den Witte Blinck

(2)

7

Individuals or couples seeking inspiration from Nature, who did not go to the dunes for relaxation as most of us do today when going for a walk. They went there for a spiritual experience, to actively observe and think about the wonders of God’s creation. The panoramic views from the dune tops, were especially perceived as inspirational, much

75

as they are today. Summarizing a text from this time ‘These dunes seemed a wasteland but gave the citizens clean drinking water, they triggered the wealth of Haarlem’s breweries, linen and silk industries as well as laundry business. They provided medical plants and beauty’ (after Leeflang 1998). The dune painting (see Fig. 6) by Esaias van der Velde (1587–1630),8 one of the famous painters of that time, is an illustration of these individuals and small groups appreciating the dune landscape around Haarlem. The figures in the painting walk or sit in the dunes, at a considerable distance from Haarlem. In those days this was something special. The two people on top of a dune seem to discuss the landscape view. The painting shows everyone in a contemplative mood in a rather level dune area. The levelling of dune areas around Haarlem was done for defensive purposes, to improve the view from the existing city walls during a siege.

The development of landscape paintings as a new genre In the wake of the growing public and artistic interest in Dutch landscapes, landscape painting developed as a new genre. As a consequence of the high demand, many painters moved to Haarlem, and the city developed as a centre of landscape painting. Thousands of paintings of dune landscapes were produced, often using the cheaper brown and red pigments to make them more affordable. They were created for the new bourgeoisie to decorate the walls of their houses, to enjoy, to show off the owner’s wealth or taste, as an investment and as a source of inspiration for those not able to travel. Later, many of these paintings found their way into museums and art collections all over the world. Of these seventeenth-century painters of dune landscapes Jacob van Ruisdael is the most important. This is not only for artistic reasons but also

Where would I like to travel now, if I consider well and think, my Helicon* from now on, will only be the ‘Witte Blink’** *Helicon in Greek mythology is the mountain where the Muses inspiring artists lived. **A blink is a high dune top with active blowouts, which shines brightly (¼blinken) in the sunlight. Mander was a Flemish artist and writer whose most important work, Het Schilderboeck (1604), chronicled the achievements of painters who had worked in Flanders and the Netherlands from the fifteenth century up to his own times; it was intended as a teaching aid for artists and is still an invaluable source of information for modern art historians. 8 Although Esaias van der Velde painted genre and military subjects, he was principally a landscape painter noted for his naturalistic depictions. He played a pioneering role in painting Dutch landscapes; for example A Winter Landscape of 1623 in the collections of the National Gallery in London.

76

J. A. M. VAN DEN ANCKER & P. D. JUNGERIUS

Fig. 2. Sperm whale stranded at Berkhey in 1598 as depicted in an engraving by Jacob Matham (1571– 1631) after a drawing by Hendrik Goltzius (Rijksmuseum, Amsterdam). It shows the public lined up to see and even climbing upon the dead whale, as well as the landscape setting with the beach and foredune.

because his paintings show geomorphological detail not encountered in the works of any of his contemporaries or in most eighteenth-century dune landscape paintings. A good example is the details of road erosion and the mobile dune (or loopduin)9 near the village of Egmond that can be seen on his paintings View of Egmond-on-the Sea of 1648 in the Currier Gallery of Art in Manchester, New Hampshire, USA (Inv. Nr. 291) and also the slightly different view of this location in the Pushkin Museum in Moscow. To illustrate the geomorphological detail in Ruisdael’s paintings, we selected a Ruisdael painting (see Fig. 7) from the collection of the Rijksmuseum in Amsterdam. It shows a blowout with water erosion, road erosion and a valley with water. Water erosion as a dominant erosional process in the vegetated dunes only

9

became apparent to scientists in the 1980s (Jungerius et al. 1981; Jungerius & Dekker 1990; Rutin 1983). Watercourses had disappeared from the dunes in the early twentieth century due to the extraction of water to improve the quality of the drinking water in the cities of Haarlem and Amsterdam; modern nature management has restored such elements. Ruisdael’s famous painted view from the dunes in the town of Haarlem shows the use of fields adjacent to the dunes for bleaching clothes from the laundries that used the clean dune water. The painting indicates the lack of blown sand from the dunes. In the beginning of the twenty-first century many dune managers were engaged in an ongoing argument that the dunes had become vegetated through human interference and should be destabilized again. Ruisdael’s

Modern degraded versions of such active mobile dunes can be found at De Blinkert near Kraantje Lek in Overveen and in Schoorl, both the Netherlands. Fossilized mobile dunes also are found along the Great Lakes (USA–Canada), where they cover prehistoric settlements.

FIRST EVIDENCE OF GEOTOURISM ON THE DUTCH COAST

77

Fig. 3. Test of the sailing car newly invented in1602 by the engineer Simon Stevin as depicted in an engraving by Willem Isaacsz. van Swanenburg (1580–1612) (Rijksmuseum, Amsterdam). The engraving shows people walking along the beach and halting to witness the car pass, as well as people specially having gone there for the occasion.

paintings and those of other painters, however, show that the dunes were already vegetated in the early seventeenth century and that wind erosion only occurred in blowouts and the foredunes.

The first known etching of a dune landscape in which the human figure is more or less absent was made around 1655 by Claes van Beresteyn (1637– 1684; see Fig. 8). A still earlier example

78

J. A. M. VAN DEN ANCKER & P. D. JUNGERIUS

Fig. 4. Departure of Charles II of England from Scheveningen, 2 June 1660, as depicted in an etching by Pieter Hendricksz. Schut (c. 1618– 1866) (Rijksmuseum, Amsterdam). The royal departure seemed to have attracted the incredible number for the time of more than 50 000 spectators, seen sitting on the dune tops to get a good view (Scott 1907).

is the black chalk and grey wash of Aelbert Cuyp (1620– 1691), Dune Landscape with a Ruin (Rijksmuseum Amsterdam) of c. 1640–45; in it only a ruin is present. The new Dutch interest in landscape painting eventually crossed the North Sea and came to fruition in England in the late eighteenth and the first quarter of the nineteenth centuries; for example in the realistic landscape paintings of John Constable (1776–1837)10 and those of the Norwich school of landscape painters (McInnes & Stubbings 2010).10 Echoing the paintings of the Dutch seventeenth

10

century artists, the paintings of these artists somewhat accidentally show casual geotourists visiting eastern England’s North Sea coast (T. A. Hose pers. comm., 2015) and significant geomorphological features (McInnes & Stubbings 2010). Other eighteenth century English landscape developments are the landscape garden movement and the rise of geotourism in the Peak District and Lake District, where issues around real and imagined visitor damage prompted discussion on the need to protect sites from the effects of mass tourism (Hose 2008, 2010, 2012).

Of the English landscape painters, Constable certainly drew his inspiration from the Dutch landscape painters, especially Ruisdael. However, Thomas Gainsborough (1727– 1788) and Richard Wilson (1714–1782) had earlier laid the foundations of English landscape painting as a genre. Whilst the former had learned his craft (and had been apprenticed to an engraver) in London, the latter had learned his craft on visits to Italy and had been encouraged by his French contemporary Claude Joseph Vernet (1714– 1789) to take up landscape painting. The Norwich school was founded in 1803 as a small group of self-taught working-class artists. They painted the hinterland of Norwich up to the mid-1830s and far from creating mere pastiches of the work of Dutch seventeenth-century artists, its members established a distinctive school of landscape painting.

FIRST EVIDENCE OF GEOTOURISM ON THE DUTCH COAST

79

Fig. 5. The sea-way between The Hague and Scheveningen, with toll house and gate, as depicted in an etching of 1681 by Cornelis Elandts (1641– 1687?) (Rijksmuseum, Amsterdam). In the seventeenth century Sir Constantijn Huygens already realized the importance of the beach for recreational purposes and published a detailed plan for the construction of a ‘Zee-straet’, a straight and paved road connecting The Hague and Scheveningen beach. He promoted healthy forms of vermaak (amusement) such as walking along the beach (with the girls), collecting shells and eating fresh fish.

Fig. 6. Dune landscape near Haarlem, 1629, as depicted in an oil painting by Esaias van der Velde (Rijksmuseum, Amsterdam). The persons in the painting are either walking or sitting in the dunes, which are at a considerable distance from Haarlem. Two people on top of a dune even seem to be discussing the view and one of them probably is the landscape painter Esaias van der Velde (Matsier, Goudswaard & Bakker, 2015, Op ’t Duin, Thoth).

80

J. A. M. VAN DEN ANCKER & P. D. JUNGERIUS

Fig. 7. Sand road in the dunes, c. 1650– 55, as depicted in an oil painting by Jacob van Ruisdael (Rijksmuseum, Amsterdam). In Haarlem alone hundreds of paintings of dune landscapes were produced in the seventeenth century, revealing the growing interest in the ‘wild’. The dunes in those days were considered a wild landscape. Jacob Ruisdael’s paintings show the most geomorphological detail, in this case a blowout with water erosion, road erosion and a valley with water. Water erosion as a dominant erosional process in the vegetated dunes only became apparent to scientists in the 1980s.

Geotourism – implicit in the sixteenth century and more explicit in the seventeenth century Geotourism was first defined by Hose (1995) as a new form of Earth science –based tourism that was recognized in the closing decades of the twentieth century, when Earth scientists started to explain the science of geology to a broader public. Apart from interesting geosites and geomorphosites, geotourism by definition requires the provision of geo-interpretation and the implementation of geoconservation measures (Hose 1995, 2012). Geotourists can be divided into ‘casual’ and ‘dedicated’ geotourists (Hose 2008, 2012); the casual geotourists appreciate the sites and facilities as something that comes across their path, the dedicated geotourists actively seek specific sites and geological information. Hose (2008) describes how modern geotourism provision has its antecedents in

late eighteenth-century and nineteenth-century field excursions, tourist guides and populist geology books in England. He further mentions how the development of these was influenced by earlier aesthetic movements from the seventeenth century onwards. In this it might be argued (T. A. Hose pers. comm., 2015) that many artists, in choosing to visit and select particular landscapes and to accurately depict their geomorphological features, were themselves perhaps dedicated geotourists. The examples discussed herein of sixteenth- and seventeenth-century Dutch tourism on the beach and in the dunes are illustrations of landscapebased tourism, in which the depicted geomorphology is perceived as part of the total system. In going to the beach, people in those centuries, like most modern visitors, were like casual geotourists, implicitly appreciating the geological forces of sea and wind on the beach. The people visiting the dunes in the seventeenth century more explicitly

FIRST EVIDENCE OF GEOTOURISM ON THE DUTCH COAST

81

Fig. 8. Dune landscape with oaks and resting figure, c. 1655, etching by Claes van Beresteyn (Rijksmuseum, Amsterdam). It is the first known etching in which the human figure is reduced to next to nothing.

considered aspects of the geosciences when appreciating and thinking about the importance of the dunes as part of God’s creation. For example they considered how the dunes seemed to be a wasteland but were in fact important for human society, as they produced clean water; how the dune morphology created attractive viewpoints; and finally, how the poor soils of the dunes hosted plants not found elsewhere that had important medicinal properties and uses. They were, then, more than just casual geotourists.

Conclusion The selection of Dutch etchings and paintings of the sixteenth and seventeenth centuries specifically referred to in this paper illustrate several types

of tourists who went to the beach and walked amongst the dunes. The earliest drawings and prints suggest that people, apart from going to the beach to buy fish or to watch special events, just walked there for recreational purposes, appreciating the landscape and active processes. As we have demonstrated, the valuation of the dune landscape started in the early seventeenth century with people actively seeking, as more than just casual geotourists, inspiration in the dunes. It gave rise to the famous school of landscape painters around Haarlem. Their drawings, prints and paintings are an important source of information about Dutch coastal landscapes and their visitors of this period. In making these visual representations it is probable that the Dutch landscape painters, in choosing to visit and select particular landscapes and to accurately depict their geomorphological features, were

82

J. A. M. VAN DEN ANCKER & P. D. JUNGERIUS

themselves dedicated geotourists. The buyers of their works belong to this same category. As the prints and paintings both provide a record of early Dutch coastal tourism and present scientific information of value to modern geomorphologists: these paintings then should be rightfully regarded as part of the geoheritage of the Netherlands. Dr. John E. Gordon persuaded us to participate in the original HOGG meeting at the Geological Society. Dr. Thomas A. Hose encouraged us to publish a paper based on our poster contribution and provided additional information, especially on the development of geotourism and the significance of Dutch artists in the development of landscape painting in England. Without the assistance of Mr. Huigen Leeflang, who facilitated our access to the different collections of etchings, prints and paintings of the Rijksmuseum in Amsterdam and also generously shared with us his knowledge of seventeenth-century landscape painting, together with his studies of the landscape painters around Haarlem, this paper could not have been written. Finally, we would like to acknowledge the advice of the two anonymous reviewers, which was invaluable in helping us to hone this paper.

References Hose, T. A. 1995. Selling the story of Britain’s stone. Environmental Interpretation, 10, 16–17. Hose, T. A. 2008. Towards a history of geotourism: definitions, antecedents and the future. In: Burek, C. V. & Prosser, C. D. (eds) The History of Geoconservation. Geological Society, London, Special Publications, 300, 37–60. Hose, T. A. 2010. The significance of aesthetic landscape appreciation to modern geotourism provision. In: Newsome, D. & Dowling, K. (eds) Geotourism, the Tourism of Geology and Landscape. Goodfellow, Oxford, 13–26. Hose, T. A. 2012. 3 G’s for Modern Geotourism. Geoheritage, 4, 7 –24. Jelgersma, S., De Jong, J., Zagwijn, W. H. & Van Regteren Altena, J. F. 1970. The coastal dunes

of the western Netherlands, geology, vegetational history and archaeology. Mededelingen Rijks Geologische Dienst, Nieuwe Series, 21, 93– 167. Jungerius, P. D. 1997. De zeereep rond 1600; potvisstrandingen aan de kust. Duin 20, 1, 14–15. Jungerius, P. D. & Dekker, L. W. 1990. Water erosion in the dunes. In: Bakker, TH. W. M., Jungerius, P. D. & Klijn, J. A. (eds) Dunes of the European Coast, Geomorphology, Hydrology, Soils. Catena Supplement, Catena Verlag Gmbh, Reiskirchen, Germany, 18, 185– 193. Jungerius, P. D., Verheggen, A. J. T. & Wiggers, A. J. 1981. The development of blowouts in ‘De Blink’, a coastal dune area near Noordwijkerhout, The Netherlands. Earth Surface Processes and Landforms, 6, 375–396. Leeflang, H. 1998. Dutch landscape: the urban view. Haarlem and its environs in literature and art, 15th– 17th century. In: Falkenburg, R. et al. (eds) Natuur en landschap in de Nederlandse kunst 1500–1850, Nederlands Kunsthistorisch Jaarboek 48, Zwolle (Waanders), 52– 115. Leeflang, H. 2002. Ruisdaels natuur. In: Sitt, M. & Biesboer, P. (eds) Jacob van Ruisdael, De Revolutie van het Hollandse Landschap. Waanders, Zwolle/ Frans Hals Museum, Haarlem. McInnes, R. & Stubbings, H. 2010. Marine Estate Research Report: Art as a Tool in Support of the Understanding of Coastal Change in East Anglia. The Crown Estate, London. Ossing, F. & Brauer, A. 2006. Contrived Reality: Weather and Geology in Jacob van Ruisdael’s Painting ‘View of Ootmarsum’. GFZ Helmholz Centre, German Research Centre for Geosciences, Potsdam, Germany. Rutin, J. 1983. Erosional processes on a coastal dune, De Blink, Noordwijkerhout. PhD thesis, University of Amsterdam, The Netherlands. Scott, E. 1907. The Travels of the King, Charles II in Germany and Flanders 1654– 1660. Constable & Co., London. Sitt, M. & Biesboer, P. (eds) 2002. Jacob van Ruisdael, De Revolutie van het Hollandse Landschap. Waanders, Zwolle/Frans Hals Museum, Haarlem.

Visitors to ‘the northern playgrounds’: tourists and exploratory science in north Norway W. BRIAN WHALLEY1* & ANNE F. PARKINSON2 1

Department of Geography, University of Sheffield, Sheffield S10 2TN, UK 2

Ilsington, Devon, UK

*Corresponding author (e-mail: [email protected]) Abstract: This paper outlines some significant visits made to north Norway by geologists and mountaineers from Britain and Ireland from the early to late nineteenth century. These visitors wrote up their travels and climbing experiences in a region in north Norway that was difficult to get to other than by sea: Øksfjordjøkelen and Lyngen. Early travellers revealed the sights of the fjord areas and thereby promoted the region for subsequent travellers. Leopold von Buch’s Travels though Norway and Lapland during the Years 1806, 1807, and 1808 probably prompted J. D. Forbes to visit and produce Norway and Its Glaciers and Archibald Geikie’s Geological Sketches at Home and Abroad as part of the contemporary discussions about the ‘glacial theory’. In the latter years of the nineteenth century the British climbers William Cecil Slingsby and George Hastings, with local climber Josef Caspari, explored the Lyngen Peninsula. Elizabeth Main (Mrs Aubrey Le Blond) also climbed in Lyngen. As well as providing written summaries of their exploits, the early explorers included photographs in their books. Some of these images are helpful in the reconstruction of the glacierized landscapes at the end of the Little Ice Age. It is suggested that present-day travellers might leave their observations available, in digital media, for future investigators.

Unlike many other mountainous parts of Europe, the mountains of north Norway are still relatively remote. Although never more than 2000 m in height, but with glaciers descending from plateau tops to valleys, they provide a serious challenge for climbers even today. Perhaps for these reasons they have attracted both climbing and scientific investigations, especially from UK universities, from the 1950s onwards. Two areas in particular have been visited, the Lyngen Alps and the plateau region of Øksfjordjøkelen on the Bergsfjord Peninsula (Fig. 1). As part of university expeditions to these two areas, the authors carried out bibliographic research on the legacies of early tourists and climbers to north Norway, especially to locate evidence of glacier positions recorded in images from the last 150 years. At the time, the significance of what the early visitors drew or photographed was not really known. In this paper we highlight the importance of some of these travellers and climbers and the images they made with particular reference to investigations of upland glaciation in north Norway. Present-day tourists might also be able to assist future research, in topics yet to be disclosed, by making images and recordings available for researchers.

The end of the ‘Little Ice Age’ in Norway At a time in the twenty-first century when glacier mass loss is a known consequence of ‘climate

warming’, it is important to determine some baseline data as a starting point for modern-day observations and measurements. Grove’s Little Ice Ages (Grove 2004) provides a compendium of investigations, especially of early reports and observations of former glacier extents and effects such as rock avalanches and debris flows. Proxy data had been used by the economic historian Ladurie (1972), who compiled vintage records in Haute Savoy to link glacier recession and climate change in the Alps. By the end of the nineteenth century, travellers, cartographers and mountaineers had been to most of the mountain areas of Europe; their maps, sketches and photographs have been used to trace volumetric and length changes of glaciers through to the present day (Grove 2004). However, the Little Ice Age, although roughly synchronous in the mountains of both hemispheres (approximately 1600–1900 CE), is characterized by variable glacier responses at its end in many locations (Grove 2004; Matthews & Briffa 2005). The proxy climatic signal marked by glacier retreat from a maximum extent may be different according to geographic location, latitude, altitude and continentality (essentially, distance from the sea) as well as local aspect. In southern Norway there are good datasets of glacier extents and volumes following the work of Matthews and co-workers (e.g. Griffey & Matthews 1978; Matthews & Briffa 2005). Present-day massbalance measurements in southern Scandinavia

From: Hose, T. A. (ed.) 2016. Appreciating Physical Landscapes: Three Hundred Years of Geotourism. Geological Society, London, Special Publications, 417, 83– 93. First published online September 17, 2015, http://dx.doi.org/10.1144/SP417.12 # 2016 The Author(s). Published by The Geological Society of London. All rights reserved. For permissions: http://www.geolsoc.org.uk/permissions. Publishing disclaimer: www.geolsoc.org.uk/pub_ethics

84

W. B. WHALLEY & A. F. PARKINSON

Fig. 1. Location of the main glacier areas mentioned in the text: Øksfjordjøkelen on the Bergsfjord Peninsula in the north and the Lyngen Peninsula farther south. Seilandsjøkulen has had visits in the twentieth century, but no historical records exist, as far as is known.

show a distinct west–east continentality (Chorlton & Lister 1971; Whalley 2004). In the far north of Scandinavia there are far fewer records. Some exist from interior valley glaciers such as Storglacia¨ren in Sweden (e.g. Holmlund et al. 1996). However, little is known about glacier positions, and hence glacier volumes, at the end of the nineteenth century. In maritime north Norway regular mass-balance measurements have only been made on Langfjordjøkelen since 1989 from the work of NVE, the Norwegian Water Resources and Energy Directorate (Kjøllmoen et al. 2000), and some more general observations in north Norway (Hoel & Werenskiold 1962; Andreassen 2000).

Thus, there is a general lack of glacier extent information in north Norway from the end of the Little Ice Age to the end of the twentieth century. The following sections present a summary of observations by some early travellers in Norway, especially from Britain and Ireland. A brief review of their writings and images shows how they have helped to amplify later glaciological work. In particular, investigations have shown how these early observations have facilitated reconstructions of late-nineteenth-century glacier limits. This is especially important, as the plateau origin of many of the glaciers in this region reveals a distinctive manner of response to altitude and area of accumulation as well as more general climatic controls

TOURISTS AND EXPLORATORY SCIENCE IN NORTH NORWAY

(Gellatly et al. 1986). Furthermore, such early observations help to elucidate the timing of events at the end of the Little Ice Age in north Norway (Whalley et al. 1989). Finally, these can be seen as past observations by visitors who would today be considered ‘dedicated geotourists’ (Hose 2008), whose records are of considerable importance to modern science; their records also provide background material for both present-day geotourists and any future geotourism provision.

Glaciology and glacial geology in the mid-nineteenth century The Irish physicist John Tyndall had visited the Alps in 1856 to carry out scientific work, and he continued to climb there in subsequent years. He started to speculate on the flow of glaciers in his Glaciers of the Alps (Tyndall 1860) but did not travel farther north into Scandinavia to extend his investigations. Tyndall was involved in the considerable debate about glaciers; their origin, flow and formation of features such as crevasses. For a fuller view of the sometimes acrimonious discussion, see Rendu & Wills (1874). These debates were largely predicated upon Louis Agassiz’s great work Etudes sur les glaciers of 1840 (Agassiz & Carozzi 1967) at a time when the ‘glacial hypothesis’ was not everywhere accepted in Britain and Ireland. Agassiz’s work is almost entirely restricted to the Swiss Alps. However, he does introduce a brief mention of work elsewhere by Leopold von Buch, noting that ‘The data published by Leopold von Buch on the snowline in northern Europe leave no doubts on the identity of the mode of formation of glaciers in polar regions and those of Switzerland’ (Agassiz & Carozzi 1967, p. 82). Christian Leopold von Buch enters our story by way of his Reise durch Norwegen und Lappland, or Travels through Lapland and Norway during the Years 1806, 1807, and 1808 (von Buch 1810), an early geotourist publication for the area. As well as being the first foreign member of the Geological Society of London, introducing the term ‘caldera’ into the geological literature, and recognizing the volcanic nature of basalt, he visited Øksfjordjøkelen during his travels. Von Buch’s report appears to have prompted other scientists to venture north.

1

85

Early observations of glaciers on the northern and southern sides of Øksfjordjøkelen On the cliffs below the southern side of the Øksfjordjøkelen plateau glacier or icefield (c. 40 km2) a ‘regenerated’ glacier cone (fall-jøkull) is fed by avalanches from the plateau above and descends to the sea at the head of the fjord (Fig. 2). Von Buch visited this by sea and included a description of the glacier in his book. His account influenced others to head into this confined fjord head, including Everest (Everest 1829) and von Bayer (von Bayer 1889).1,2 The latter included a sketch of its size and the descent to the fjord (Gellatly et al. 1989). Another visitor to the Bergsfjord Peninsula influenced by von Buch was Sir Archibald Geikie. He reported in Geological Sketches at Home and Abroad (Geikie 1866), an early text for well-travelled geotourists, that he visited the fall-jøkull, which he called a ‘glacier remanie´’. Adolf Hoel, in his report on the glaciers of this area, records Geikie’s visit and includes his sketch (Fig. 3) (Hoel & Werenskiold 1962). The plateau ice both was clearly thicker in Geikie’s day and deposited far more ice on the falljøkull than at present. Comparing the sizes of the fall-jøkull in Figures 2 and 3 shows this difference. Trawlers used this remnant glacier as a convenient source of ice blocks for freezing their catches for export abroad. Hoel & Werenskiold (1962, pp. 116–117) include a brief history of this ‘ice trade’ and export of ice blocks (but see also Smith 1989). Further observations at the end of the nineteenth century were made by the mountaineers George Hastings and Elias Hogrenning (see ‘Climbers in the Lyngen Peninsula’ below), who visited the fjord in 1898 (Hastings 1899). Their photograph (Gellatly et al. 1989) shows, by comparison with Figure 3, a reduction in the altitude of the skyline of about 20 m. James D. Forbes, as well as contributing to the early ‘glacial’ controversies mentioned earlier, was also involved in journeys to north Norway, presumably to see evidence of the ‘glacial theory’ for himself. This he did by steamer, although his account (Forbes 1853) indicates he was not always successful in making visits to points of interest suggested to him and was prevented from accessing the southern fjord. However, he did land on the north side of the Øksfjordjøkelen ice cap in

The usual Norwegian word for glacier is ‘bre’ or ‘breen’ with the definite article. Sometimes, traditionally, the Icelandic equivalent jøkull (or jo¨kull) is used. Some places, and particularly mountains and passes, are known by Sami (Lapp) names and the spellings vary across the years. For example, Mrs Main uses Balkisvarre and Joekkevvare, whereas the current Norwegian maps, to which we have adhered, use Balgesvarri and Jiekk’varri respectively. 2 Th. von Bayer was the pseudonym of Therese Charlotte Marianne Auguste von Bayern, a noted scientist and travel writer.

86

W. B. WHALLEY & A. F. PARKINSON

Fig. 2. The lower, triangular, ice mass is a ‘regenerated glacier’ (fall-jøkull) derived from the accumulation of ice and snow avalanching from the glacier above. The latter is an outlet glacier of Øksfjordjøkelen, an icefield or plateau glacier that stretches back several kilometres from the skyline. (Photo: Whalley taken in 1986)

Nuvsfjord in 1851 (Fig. 4). He recorded that ‘it will be seen that these glaciers have the true alpine character, being outlets of extensive snow-fields whose general level exceeds 2000 feet – the glaciers descending the natural ravines in the midst of slopes of the brightest verdure’ (Forbes 1853, p. 81). From later observations, it has been deduced that the glaciers were at, or very nearly at, their Little Ice Age maxima positions by association with the terminal

and lateral moraines (Gellatly et al. 1989; Whalley et al. 1989; Evans et al. 2002). The same outlet glaciers at the present-day can be seen in Figure 5. Whereas the glacier snouts in the valleys shown by Forbes have retreated many hundreds of metres from the Little Ice Age termini, there is very little recession apparent at the plateau rock edges. The figure in the photograph is on a low moraine, barely visible on aerial photographs,

Fig. 3. Sketch of the southern outlet of the Øksfjordjøkelen plateau glacier and the fall-jøkull or glacier remanie´ from Archibald Geikie’s visit of 1862. Compare with Figure 2. From Geikie (1866).

TOURISTS AND EXPLORATORY SCIENCE IN NORTH NORWAY

87

Fig. 4. Glaciers descending from the north side of the Øksfjordjøkelen plateau ice cap (highest point about 1140 m asl) as depicted from more than one location by James D. Forbes in 1851. From Forbes (1853).

at the edge of the plateau, with the glacier edge some 50 m behind the camera position. The moraine probably corresponds to the extent of the glacier on the plateau as shown by Forbes (see Fig. 4) at or near the end of the Little Ice Age in the area. It is estimated (Gellatly et al. 1989) that between 1851 and 1986 the ice cap down-wasted (vertically) some 20 –30 m, with the outlet glaciers in the northern valleys retreating some 0.7 km from the Little Ice Age moraines. We surmise that glacier recession is the result of climatic warming that affected the glacier termini rather than any substantial reduction in precipitation that affected the plateau accumulation area. In the twentieth century, it was possible

to map glacier extents and moraines of the Øksfjordjøkelen area (Evans et al. 2002). However, glacier recession, and particularly rates of recession since the end of the Little Ice Age, would not be as clear without the good sketches provided by nineteenthcentury traveller –scientists, or as they would be termed today, ‘dedicated geotourists’ (Hose 2008), such as Geikie and Forbes.

Climbers in the Lyngen Peninsula Geotourists to north Norway in the nineteenth century primarily visited the region by sea, because the roads were poorly engineered and only linked

Fig. 5. The northern edge of the Øksfjordjøkelen plateau (1991) with glacier outlets in a very similar state to those shown by Forbes (Fig. 4). (Photo: Whalley)

88

W. B. WHALLEY & A. F. PARKINSON

Fig. 6. Strupvatnet in 1975, looking east towards Strupskar, showing Strupvatnet in its fully drained state after a drainage of c. 3 m earlier in the summer. The drop in lake level can be identified by the pale grey band above the lake in the foreground and the corresponding gap below the snowbanks on the far side of the lake. Although there is water in the lake, it does not drain below this level. Field observations have shown that it drains over a bedrock sill when no longer dammed by the glacier. (Photo: T. P. L. King)

the coastal fringes. As well as the just-outlined Øksfjordjøkelen, they also passed by glaciers in the mountains of the Lyngen Peninsula (Lyngshalvøyen) (Fig. 1). The glacial tongue of Strupbreen, like that of the fall-jøkull of southern Øksfjordjøkelen, was also used to provide ice for trawlers (to preserve their catches) in the days before mechanical refrigeration. A photograph of this regenerated glacier, taken by the local photographer Egeret Wilse in 1900, can be found in Johnsen & Skjerven (1984). Unlike the earlier ‘coastal’ scientific explorers, it was necessary for the climbers to get up high on these relatively remote peaks. In doing so, they

provided some observations and photographs of glaciers at the turn of the nineteenth century. One of these mountaineers was the Norwegian Josef Caspari, who climbed with the noted English climber William Cecil Slingsby in Lyngen. In 1898 they climbed with George Hastings and Elias Hogrenning (Johnsen & Skjerven 1984), and they explored the deep valley of the Strupskar (Hastings 1899). They discovered (and probably not even the Sami, the local inhabitants, had seen it before) Strupvatnet, a lake impounded between the valley walls and Strupbreen at its eastern extremity (Fig. 6). Returning a few days later, they found that the level of the lake had suddenly dropped. Hence, this was a

Fig. 7. Part of a very similar view photographed by the Slingsby– Hastings–Caspari party in 1898. The rock island in the centre is the same outcrop seen in the centre of Figure 6. The vertical space between each of the three pairs of black bars identifies the snowbank–lake gap (cf. Fig. 6). (Photo: Alpine Club Archives)

TOURISTS AND EXPLORATORY SCIENCE IN NORTH NORWAY

true ice-dammed lake that was suddenly drained, as was confirmed by Aitkenhead (1960) on a Durham University expedition and again in 1969 by a Leicester University expedition (Whalley 1971). A subsequent paper (Whalley 1973) was able to use the descriptions provided by Hastings, Slingsby and Caspari to piece together the drainage events of Strupvatnet as well as the size of Strupbreen that was effectively damming it; a photograph in the Alpine Club archives (Fig. 7) showing the level of

89

the lake in 1898 was later discovered and confirmed this interpretation. At the same time as Hastings, Slingsby, Caspari and Hogrenning were in north Lyngen, Elizabeth Main ne´e Hawkins-Whitsed was exploring the southern part. Mrs Main (Fig. 8) had climbed in the European Alps and was the first president of the Ladies’ Alpine Club (Nugent 2013). Although here referred to as Elizabeth Main, her Mountaineering in the Land of the Midnight Sun (Le Blond

Fig. 8. Mrs Main (Mrs Aubrey Le Blond) as depicted in her Mountaineering in the Land of the Midnight Sun (Le Blond 1908).

90

W. B. WHALLEY & A. F. PARKINSON

Fig. 9. A photograph taken in 1974 showing the very thin remnant of ice cover on the plateau of Bredalsfjellet (1300 m). Compare this with Figure 10. The latest Google Earth imagery (9 July 2011; 698 30′ 22′′ N, 198 58′ 42′′ E) shows there is still limited blue ice exposed on the plateau, even at the end of the ablation season. (Photo: Whalley).

1908) was published under her subsequent married name, Mrs Aubrey Le Blond. Not only was Elizabeth Main an intrepid climber, she and her Swiss guides, Joseph and Emil Imboden, were carrying the bulky photographic equipment of the day.

Several of her photographs illustrate glacier change in Lyngen (Gellatly et al. 1986), but two will have to suffice here. Figure 9 shows a view from the edge of the Jiekk’varri plateau taken in 1974. Mrs Main had taken the photograph shown

Fig. 10. Bredalsfjellet (distant), showing thicker ice cover than at present but with no ice flowing from the plateau. The edge of the Jiekk’varri plateau glacier is on the right. This was taken by Mrs Main, probably in 1898, image facing p. 187 of Le Blond (1908).

TOURISTS AND EXPLORATORY SCIENCE IN NORTH NORWAY

91

Fig. 11. The terminus of Fugledalsbreen (Fauldalen) taken by Mrs Main in 1898. This is a glacier remanie´ in the same manner as the fall-jøkull at Øksfjordjøkelen but here looks like a valley glacier, even though all of the ice comes from the plateau glacier of Jiekk’varri some 1000 m above. (Le Blond 1908).

in Figure 10 probably in 1898. Although from farther away, when cropped to give as near a match as possible, it can be compared to the far plateau of Bredalsfjellet (c. 1300 m) of Figure 9 (Gellatly et al. 1986). Visiting the site from which Figure 9 was taken in 1974 prompted investigations of the plateau glacier extent, which led to the records in the literature of early climbers in the area such as Mrs Main. The photographs from Mrs Main’s book (Le Blond 1908) allow us to compare glacier behaviour over a period spanning about 100 years. Indeed, we see the expected recession of the glacier termini; in Figure 11 (taken in 1898) the glacier terminated in the lake, but had moved some 200 m back from its edge by 1985 (Gellatly et al. 1986). Inspection of Google Earth images (9 July 2011; centered at 698 29′ 12′′ N, 198 50′ 05′′ E) suggests the recession is now some 500 m from the edge of the lake, obscured by the ice tongue shown in Figure 11. This glacier snout recession is due to some climatic warming and consequent pronounced melting of the snout area. However, in the case of large plateaus with ice on top (such as Øksfjordjøkelen), middle-sized plateaus (such as Jiekk’varri), or even less extensive areas (such as Bredalsfjellet), a slightly reduced amount of precipitation/accumulation has apparently meant a reduction in ice supply from the plateau ice down the cliffs. If the plateau is ‘high enough’ (i.e. is in the glacier’s accumulation area) and the plateau area ‘sufficiently’ large, then

a glacier can continue to exist (such as Bredalsfjellet, Fig. 9) even if in a somewhat diminished volume. This has allowed us to make a model of glaciers on plateaus (Whalley et al. 1995) developed from the idea of Manley (1955) to inform their reconstruction in areas where glaciers no longer exist, as in the British Isles (McDougall 2001).

Conclusions Through the observations of nineteenth-century travellers, some of whom were very much geotourists in the modern sense, and ascents of the first mountaineers in what Slingsby (1904) called, ‘the northern playground’, we can use their images to fill in for the ready lack of (especially 1:100 000 topographic) maps until the late- nineteenth century of these remote glaciers. More recent glaciological surveys and meteorological measurements have enabled comparisons with the early observations. Together, these have assisted with the modelling of glacier response and behaviour to climate change since the Little Ice Age in northern Scandinavia. This would not have been possible had not these mountaineers and scientists made their images available in books and journals, a valuable resource potentially also underpinning modern geotourism provision, so enabling ‘baseline’ data to be constructed with relevant geographical information and dates.

92

W. B. WHALLEY & A. F. PARKINSON

Although this paper has concentrated on the recording of mountain glacier features, it is worth noting that the books left by travellers such as Geikie, Forbes, von Buch and Main were not intended to be ‘scientific’ works but were intended to convey something of the overall character of the regions in which they travelled. So, as well as geological and glaciological observations, there are also fragments of historical and ethnographic information that might be of use to researchers in other disciplines. The travel and climbing books of the nineteenth century were, as today, written by a few travellers for the benefit of many. However, with modern adventurer–travellers, relatively easy worldwide travel and lightweight (especially digital) cameras, just about anyone can now explore and record the world’s wilderness areas. Many modern digital cameras and smartphones can now affix notes, ratings, image location and time data (called ‘metadata’). The written record can be in the form of website notes or web logs and even tweets with photographs uploaded to one of several repositories on the Internet. Those present-day travellers interested in recording their exploits should perhaps also make an effort to provide and archive their images with such image metadata made ‘discoverable’. Future researchers will no doubt still use traditional archives and libraries to retrieve catalogued image data, but searching of digital archives will almost certainly be done online. We do not know how our digital photographic images might be used by later generations of researchers, but we can make a contribution to research, even as casual geotourists (Hose 2008), ecotourists or just travellers, by digitally recording our images and field notebooks and making them accessible for use in the future by researchers in several disciplines. We thank colleagues and members of the various expeditions to north Norway: University of Leicester to Lyngen (1969), University of Sheffield to Øksfjordjøkelen (1986), British Schools Exploring Society to Lyngen (1979, 1984) and Øksfjordjøkelen (1989) and Earthwatch to Øksfjordjøkelen (1990). We also thank the librarians at the Royal Geographical Society and the Alpine Club for their assistance with searches, T. A. Hose for his editorial support, and the two anonymous referees who made useful contributions to the paper.

References Agassiz, L. & Carozzi, A. V. 1967. Studies on Glaciers. Hafner, New York. Aitkenhead, N. 1960. Observations of the drainage of a glacier dammed lake in Norway. Journal of Glaciology, 3, 607–609. Andreassen, L. M. 2000. Regional change of glaciers in northern Norway. Norwegian Water Resources and Energy Directorate (NVE), Oslo, 1/2000.

Chorlton, J. C. & Lister, H. 1971. Geographical controls of glacier budget gradients in Norway. Norsk Geografisk Tidsskrift, 25, 159– 164. Evans, D. J. A., Rea, B. R., Hansom, J. D. & Whalley, W. B. 2002. Geomorphology and style of plateau icefield deglaciation in fjord terrains: the example of Troms-Finnmark, north Norway. Journal of Quaternary Science, 17, 221 –239. Everest, R. 1829. A Journey through Norway, Lapland, and Part of Sweden with Remarks on the Geology of the Country. Underwood, London. Forbes, J. D. 1853. Norway and Its Glaciers. Black, Edinburgh. Geikie, A. 1866. Geological Sketches at Home and Abroad. Macmillan, London. Gellatly, A. F., Whalley, W. B. & Gordon, J. E. 1986. Topographic control over recent glacier changes in southern Lyngen Peninsula, north Norway. Norsk Geografisk Tidsskrift, 41, 211–218. Gellatly, A. F., Whalley, W. B., Gordon, J. E., Hansom, J. D. & Twigg, D. S. 1989. Recent glacial history and climatic change, Bergsfjord, TromsFinnmark, Norway. Norsk Geografisk Tidsskrift, 43, 21–34. Griffey, N. & Matthews, J. A. 1978. Major Neoglacial glacier expansion episodes in southern Norway: evidences from moraine ridge stratigraphy with 14C dates on buried palaeosols and moss layers. Geografiska Annaler, 60A, 73– 90. Grove, J. M. 2004. Little Ice Ages: Ancient and Modern. 2nd edn. Routledge, London. Hastings, G. 1899. The Lyngen district. Alpine Journal, 19, 611– 615. Hoel, A. & Werenskiold, W. 1962. Glaciers and snowfields in Norway. Norsk Polarinstitutt Skrifter, 114, 1– 291. Holmlund, P., Karle´n, W. & Grudd, H. 1996. Fifty years of mass balance and glacier front observations at the Tarfala Research Station. Geografiska Annaler, 78A, 105– 114. Hose, T. A. 2008. Towards a history of geotourism: definitions, antecedents and the future. In: Burek, C. V. & Prosser, C. D. (eds) The History of Geoconservation. Geological Society, London, Special Publications, 300, 37– 60, http://doi.org/10.1144/SP300.5 Johnsen, B. & Skjerven, O. 1984. Lyngsalpene. Universitetsforlaget, Tromsø. Kjøllmoen, B., Olsen, H. C. & Svaerd, R. 2000. Langfjordjøkelen i Vest-Finnmark: norges vassdrags- og energidirektorat (NVE), Oslo. Glasiohydrologiske undersøkelser 2000/3. Ladurie, E., Le R. 1972. Times of Feast, Times of Famine: A History of Climate Since the Year 1000. Allen and Unwin, London. Le Blond, A. 1908. Mountaineering in the Land of the Midnight Sun. Fisher Unwin, London. Manley, G. 1955. On the occurrence of ice domes and permanently snow-covered summits. Journal of Glaciology, 2, 453– 456. Matthews, J. A. & Briffa, K. R. 2005. The ‘Little Ice Age’: re-evaluation of an evolving concept. Geografiska Annaler, 87A, 17– 36. McDougall, D. A. 2001. The geomorphological impact of Loch Lomond (Younger Dryas) Stadial plateau

TOURISTS AND EXPLORATORY SCIENCE IN NORTH NORWAY icefields in the central Lake District, northwest England. Journal of Quaternary Science, 16, 531– 543. Nugent, F. 2013. In Search of Peaks, Passes & Glaciers: Irish Alpine Pioneers. Collins Press, Cork. Rendu, M. L. C. & Wills, A. 1874. Theory of the Glaciers of Savoy. Macmillan, London. Slingsby, W. C. 1904. Norway: The Northern Playground. David Douglas, Edinburgh. Smith, R. A. 1989. A historical perspective of the nineteenth century ice trade. In: Oerlemans, J. (ed.) Glacier Fluctuations and Climate Change. Kluwer, Dordrecht, 173–182. Tyndall, J. 1860. The Glaciers of the Alps: Being a Narrative of Excursions and Ascents, an Account of the Origin and Phenomena of Glaciers and an Exposition of the Physical Principles to which They are Related. Murray, Edinburgh. von Bayer, Th. 1889. Ueber den Polarkreis. Brockhaus, Leipzig. von Buch, L. 1810. Reise durch Norwegen und Lappland, Bd 2. Nauck, Vienna; 1813. Travels through Norway and Lapland during the Years 1806, 1807, and 1808. J. Black (trans.). Henry Colburn, London.

93

Whalley, W. B. 1971. Observations of the drainage of an ice-dammed lake, Troms, North Norway. Norsk Geografisk Tidsskrift, 25, 165– 174. Whalley, W. B. 1973. A note on the fluctuations of the level and size of Strupvatnet, Lyngen, Troms, and the interpretation of ice loss from Strupbreen. Norsk Geografisk Tidsskrift, 27, 39–45. Whalley, W. B. 2004. Glacier research in mainland Scandinavia. In: Cecil, L. D., Green, J. R. & Thompson, L. G. (eds) Earth Paleoenvironments: Records Preserved in Mid- and Low-Latitude Glaciers. Kluwer, Dordrecht, 121– 143. Whalley, W. B., Gordon, J. E. & Gellatly, A. F. 1989. Effects of topographic and climatic controls on 19th and 20th century glacier changes in the Lyngen and Bergsfjord areas, north Norway, In: Oerlemans, J. (ed.) Glacier Fluctuations and Climate Change. Kluwer, Dordrecht, 153–172. Whalley, W. B., Gordon, J. E., Gellatly, A. F. & Hansom, J. G. 1995. Plateau and valley glaciers in north Norway: responses to climate over the last 100 years. Zeitschrift fu¨r Gletscherkunde und Glazialgeologie, 31, 115–124.

The role of local societies in early modern geotourism: a case study of the Chester Society of Natural Science and the Woolhope Naturalists’ Field Club CYNTHIA V. BUREK1* & THOMAS A. HOSE2 1

Centre for Science Communication, Department of Biological Sciences, University of Chester, Parkgate Road, Chester, CH1 4BJ 2

School of Earth Sciences, University of Bristol, Wills Memorial Building, Queens Road, Bristol, BS8 1RJ, UK *Corresponding author (e-mail: [email protected]) Abstract: Local voluntary natural science societies played an important role in the development of early modern geotourism. This chapter explores the development of field, especially geological, excursions and their popularity in two local natural science societies – The Chester Society of Natural Science and the Woolhope Naturalists’ Field Club – from the 1850s to the 1950s. Both societies were established in the borderlands between England and Wales and had a strong emphasis on local and regional scientific studies. They exemplify broader trends in public engagement in the natural sciences and associated fieldwork consequent upon the British socio-political environment. Further, they draw out comparisons between the attitudes of society to excursions and scientific fieldwork, as well as involvement by social status and gender.

The early context for local natural history societies and fieldwork This chapter examines the development and influence on landscape (geo)tourism of two local natural history societies between the 1850s and 1950s. Inevitably, with voluminous material available for study and comment, it is necessary to confine analysis to the detail that best underpins a particular matter or viewpoint. Historically this timeframe post-dates the influence of the Napoleonic period and is completed by the post-war austerity; the events associated with these helped mould the British tourism experience. The beginning of the timeframe also coincides with the spread of the railways, which opened up travel opportunities for sections of society previously denied this (such as women), and closes with the eve of their decline in the 1960s. It coincides with the emerging availability of medium-scale walking maps, tourist guidebooks and geology field guides (Hose 2007), the prices of which were rapidly becoming affordable to the populace due to improvements in printing (especially for illustrations) and binding technology. Most people, except perhaps the gentry, some professionals (such as clerics and doctors) and the military, stayed close to home until after World War I. They related mainly to their local area because distant travel was difficult and expensive and many lacked much free time. This fostered an interest in local and regional landscape studies;

field excursions became a major feature of local society programs. However, it must be remembered that there have always been travellers from other sections of society: pilgrims, migrant workers and entertainers. Only the aristocracy (and later, the offspring of wealthy merchants) originally had the time, leisure and money to travel far afield and abroad; for the latter, the Grand Tour was a feature of early adulthood from the seventeenth century. Landscape tourism dating from the Enlightenment and the Grand Tour, explored the spiritual elements of landscape. However, Britain’s industrial heritage, partly owing its origins to geology as well as engineering, attracted early geotourists who were curious and compared these anthropogenic sites with wild and natural landscapes; thus Romantic landscape tourism is the antecedent of modern geotourism (Hose 2008). The closure of Europe to British tourists during the Napoleonic era led to the re-evaluation and exploration of their own Romantic landscapes. Later, the railway companies encouraged domestic tourism and promoted these home-grown landscapes. These were so established by the close of the nineteenth century that the twentieth century motor tourist followed in their tracks and continues to do so today. The social upheavals of the Great War led to increasing political pressure, with the emergence of hiking as a major leisure activity, for access to the countryside, especially that in the ‘wild’ places which was jealously guarded by the

From: Hose, T. A. (ed.) 2016. Appreciating Physical Landscapes: Three Hundred Years of Geotourism. Geological Society, London, Special Publications, 417, 95– 116. First published online February 6, 2015, http://dx.doi.org/10.1144/SP417.6 # The Geological Society of London 2016. Publishing disclaimer: www.geolsoc.org.uk/pub_ethics

96

C. V. BUREK & T. A. HOSE

major landowners. The populace was then interested in the countryside for aesthetic recreational rather than scientific enquiry purposes. In the period immediately after World War II, the establishment of National Parks and National Nature Reserves probably seemed a natural culmination of the activities initiated by a small and mainly male social elite in late Georgian Britain. The role of local philosophical societies, naturalists’ societies and field clubs was critical in the rise of early modern geotourism in Britain. They began to appear towards the end of the Age of Enlightenment in the late eighteenth and early nineteenth centuries, although a few were founded as late as the last quarter of the nineteenth century. The establishment of The Woodstock Field Naturalists’ Club in 1934 was unusually late. The first philosophical society was established in Derby in 1783. It served as a regional forum for male professionals (especially the gentry, clerics, doctors, industrialists and manufacturers) and formed a large scientific library. In this respect it was the model that many kindred bodies later emulated. Likewise, the society was founded by a leading figure of the day, Erasmus Darwin. This prominent founder approach was followed by just about every other similar society. Post-1830, these societies were generally entitled naturalists’ societies or field clubs rather than philosophical societies. The Darlington Society for Promoting the Study of General and Natural History and Antiquities, founded at the latest in 1793, was then unusual in its choice of name, but it was a short-lived venture; having seemingly long ceased to exist, its role was taken on by the Darlington Naturalists’ Society formed in 1860. In turn the Society’s assets were transferred in 1891 to the newly formed Darlington Naturalists’ Field Club which in 1896 renamed itself the Darlington and Teesdale Naturalists’ Field Club.1,2 Contemporaneously, the Belfast Field Naturalists’ Club (now the Belfast Naturalists’ Field Club) was formed as the first of a series of such clubs in response to the increasing interest in natural sciences in Victorian Ireland. Its first field trip on 6 April 1863 saw 88 members travel to Islandmagee to collect fossils; 1

this was a popular activity at the time, as geology was at the leading edge of scientific interest in the popular imagination. The first local geological society, the Royal Geological Society of Cornwall, was established in 1814. The second was a local society in Newcastle-upon-Tyne in 1829 (although it followed on from an earlier philosophical society established in 1793) and the third was the Edinburgh Geological Society in 1834 (Burek 2008). The Dudley and Midland Geological Society was formed in 1842, but disappears from the record after 1843; however, in 1862 its place was taken by the Dudley & Midland Geological & Scientific Society & Field Club which just survived into the twentieth century. The Yorkshire Geological Society was established in 1837 and the Liverpool Geological Society in 1859. All these bodies still run lecture and field trip programmes. The Woolhope Naturalists’ Field Club (WNFC) was founded in 1851 and The Chester Society for Natural Science (CSNS) in 1871; the case study organizations therefore straddle the period of major growth in interest in natural history and geology and the rise of early modern geotourism. By the 1850s many natural history societies had begun to found museums to house their burgeoning collections (Knell 2000) and libraries and in which to deliver their lectures. By the 1920s, with a considerable demise in interest in the natural sciences but an increasing interest in archaeology, the majority of these had passed into the care of local authorities. However, this was not always to the benefit of their natural science collections or libraries.

Geotourism considered Geotourism was first defined in the mid-1990s by Hose (1994, 1995) and recently redefined as: ‘The provision of interpretative and service facilities for geosites and geomorphosites and their encompassing topography, together with their associated in-situ and ex-situ artefacts, to constituency-build for their conservation by generating appreciation, learning and research by and for current and future

A 33-page manuscript book, with entries dated 1793 and 1794, in the possession of the Darlington and Teesdale Naturalists’ Field Club, concerns a Society for Promoting the Study of General and Natural History and Antiquities. Two signatures in the book are by its Secretary (William Hutchinson of Barnard Castle) and Treasurer (Edward Robson of Darlington). Hutchinson is the author of the ‘History and Antiquities of the County Palatine of Durham’ within which Edward Robson’s local flora, presented to the Society in 1794, is noted. 2 The present society was founded on 29 April 1891 in the Club Room of the Mechanics’ Institute. Twenty-two people joined as founder members, and the annual subscription was fixed at five shillings (50p). The Mechanics’ Institute was asked to provide a suitable room for the use of the club. They did so but at an annual rent of £5 which is why, to get their money’s worth, the society started holding weekly meetings – the practice of which continues today. The Club’s objects were the study of natural history of the district, compiling a catalogue of the recent and fossil fauna of the same, and the procuring and arrangement of specimens of local natural history.

THE ROLE OF LOCAL SOCIETIES IN EARLY MODERN GEOTOURISM

generations’ (Hose 2011). It therefore comprises (Hose 2012) three elements: geoconservation, defined by Burek & Prosser (2008) as unfulfilled in this chapter’s timeframe; geohistory, which is intrinsic; and geointerpretation, which began in the chapter’s early timeframe (Table 1). Hose (2000, p. 136) recognized the different motivations and expectations of two major geotourist groups: Dedicated Geotourists, individuals who purposefully select to visit geosites and exhibits for the purpose of personal educational or intellectual improvement and enjoyment; and the Casual Geotourists, individuals who visit geosites and exhibits primarily for the purpose of pleasure and some limited intellectual stimulation. For this chapter, the differences between fieldwork, field trips, excursions and hence casual/dedicated geotourism lie within the proportion of the event devoted to scientific endeavour and social interaction for participants as mentioned in Burek (2012). While true modern geotourism provision can be recognized as beginning around the 1970s (Hose 2011, p. 346), early modern geotourism provision can be recognized as beginning in some earnest around the 1840s with the emergence of the numerous naturalists’ societies and field clubs together with the publication, most notably those of De la Beche and Gideon Mantell, of the first geology field manuals and field guides (Hose 2007, 2008). Fieldwork was initially undertaken in the naturalists’ societies and field clubs in a scientifically rigorous fashion, recording observations for publication or, often with some assessment, for educational purposes – essentially, dedicated geotourists. Enjoyment was present for some but it was not the main aim of the outing; this is particularly so for the early WNFC field excursions. With casual geotourism however, as in the majority of the CSNS excursions, social interaction and pleasure were often the primary reasons for venturing out; while education was often part of the experience, it was not the main aim in participants’ minds. This pleasure/leisure and scientific data collection distinction underpins this chapter. The drivers of travel and early modern geotourism

were curiosity, influence and power, with the collection of scientific data. With poor regional access to the few and mainly metropolitan national scientific and conservation bodies, most people belonged to their local natural science society and discovered the world close to home, examined in the incorporated WNFC and CSNS case studies.

The Woolhope Naturalists’ Field Club The precursor of the WNFC was the Herefordshire Natural History, Literary, Philosophical and Antiquarian Society established in Hereford in 1836. Following a lecture to that society by the Reverend W. S. Symonds, in which he mentioned the work of the Cotteswold, Tyne-side and Berwick Field Clubs in developing the fauna and flora of their districts, it was intimated that something similar was needed for Herefordshire. Shortly afterwards, W. H. Purchas delivered a lecture ‘On the Ferns of Herefordshire’ and suggested that ‘. . . the progress of natural science might be aided by the labours of local observers’ (Edmunds 1852) and M. J. Scobie set about organizing a natural history society. When established in the winter of 1851, it was subsequently named (Edmunds 1852) ‘The Woolhope Naturalists’ Field Club, its first object being the investigation of the locality which is geologically known as “The Woolhope Valley of Elevation,” a district which represents, as it were, an epitome of the history of immense tracts of the earth’s surface. To the Geologist, its having been the birthplace, so to speak, of the Silurian System of Murchison, must make it ever classic ground. The operations of the Club, however, will not be confined to the Woolhope District; the vicinity of Whitchurch, and the limestone beds of Aymestry, have been selected for future excursions.’ Its earliest meeting Minutes are dated 13 April 1852 when R. M. Lingwood was in the chair. However, since 30 members were elected at this meeting, together with seven distinguished honorary botanical and geological members, it can be assumed that its genesis was somewhat earlier. Lingwood was elected its first

Table 1. Comparison of methods of geointerpretation communication in late Victorian/Edwardian Times and the present day Middle eighteenth/early twentieth centuries Literary Oral Human communication Mobile

97

Twenty-first century Graphical Visual Non-human communication; sometimes electronic/digital Static

98

C. V. BUREK & T. A. HOSE

President and Scobie, for his exertions in establishing it, was elected Honorary Secretary. The Club’s initial Rules included three of some importance: (1) ‘For the practical study, in all its branches, of the Natural History of Herefordshire and the districts immediately adjacent’; (2) ‘That the Club consists of 40 Members, with such Honorary Members as may be admitted from time to time’; and (3) ‘That the Annual Subscription be Ten Shillings’. The Club’s initial Rules covering fieldwork noted: ‘That the Members of the Club shall hold three Field Meetings during the year in the most interesting localities for investigation of the Natural History of the district’; ‘That those Members to whom it may be convenient shall breakfast together at the nearest country Inn at 9 o’clock, after which the researches of the day shall commence’; and ‘That 4 be the hour appointed for Dinner, after which any papers shall be read by the respective authors. Each Member may introduce a friend on such occasions who must pay his own expenses.’ It is clear from these rules that from the outset the Club had a men-only and socially elite membership dedicated to serious scientific research in Herefordshire and its adjacent counties. Over time, and reflecting changing local and national interests, the WNFC amended its rules; for example, in 1903 its First Rule then included archaeology. The Club held its first 1852 field meeting, to the Woolhope Valley and Stoke Edith, on Tuesday 18 May. Attendance, although ‘goodly’, was affected by it being the wettest day for several weeks. Following a 9 am breakfast at Tarrington, the Reverend W. S. Symonds was voted into the chair because of the President’s absence due to illness. The Honorary Secretary then read several letters from members unable to attend together with a note from Lady Emily Foley welcoming the Club. She is particularly significant in early modern geotourism because she promoted the town’s hydrotherapy developments and, indirectly, geoconservation.3 Her horticulturist met the members and acted as their guide through the estate’s gardens and conservatories. Heavy rain, however, prevented any field investigations until midday. Meanwhile, some time was spent examining a collection of Pleistocene mammals discovered during the Hereford and Gloucester Canal’s construction. The second field meeting of 1852 was to Whitchurch on 20 July. The third and final field meeting of 1852, on Tuesday 21 September, was to Amestry to explore 3

the geology of the Silurian Amestry Limestone and botany. The party met at the Mortimer’s Cross Inn for a 9 am breakfast; those from Hereford having, as unrecognized geotourists, ‘greatly enjoyed the ride of fifteen miles through the charming scenery of that district.’ While the weather was initially wet and stormy it gradually cleared to a fine day for the season; seemingly ‘the high equinoctial S.W. winds having carried off the masses of rain-cloud which gathered so ominously in the morning.’ The chair was filled by the Club’s president and, significantly, also present were the Reverend T. T. Lewis, Mr Scobie, F.G.S. and W. H. Purchas. The first had greatly assisted, although was unacknowledged by, Murchison in his geological exploration of the Welsh Borderland; with Dr J. Lloyd he had discovered the Ludlow Bone Bed. At its first and third meeting in its inaugural year, the WNFC botanical and geological interests, so commonly intertwined in Victorian natural history, were therefore explored. In many instances the first populist accounts of an area’s geology were often to be found within newly published floras.4 In the following years the mid-week dates for field meetings were chosen on the basis of convenience for an elite social group; attendees were listed within the WNFC’s Proceedings. In its early years the WNFC’s activities were reported by Flavell Edmunds (a local amateur botanist) in the Hereford Times, of which he was Editor. Its activities from 1852 to 1865 were summarized in the President’s annual address as required by its initial rules. Six pamphlets were issued during the same period; the Club’s first Transactions volume for 1866 incorporated them. In 1907 it sent out 300 copies of its Transactions to its 260 ordinary members, honorary members and corresponding societies; the costs of any included lithographic or other illustrations were covered by the authors, probably explaining their paucity.

The Chester Society of Natural Science There were two attempts to set up local scientific societies in Chester before the CSNS. The first, The Chester Literary and Philosophical Society (1812– 1813), lasted barely a year and had no prominent figurehead. The second, The Chester Natural History Society (1858– 1859), had a local scientific backing and 80 members (Siddall 1911). In both

Lady Emily Foley had a well-house built at St Ann’s Well. She influenced the route and design of the railway across her lands and ensured, because she did not like tunnels, that the bulk of the line was in cuttings to mask its route in the landscape. She was also a key sponsor of Great Malvern Station (at which she had her own private waiting room, now ‘Lady Foley’s Tea Room’). 4 A Flora of Herefordshire (W.H. Purchase & A. Ley) was published by the WNFC in 1889.

THE ROLE OF LOCAL SOCIETIES IN EARLY MODERN GEOTOURISM

cases, the fees for entry were high and membership seems to have been limited to men. The CSNS was therefore the third attempt to set up a society in Chester. It was initiated by Canon Charles Kingsley (1815–1875) after pressure from local people while he was resident in Chester Cathedral during 1870–1873. The Society started life on 26 May 1871, with ‘All persons eligible to become members’ (Siddall 1911). It is interesting to note here that at the Third Conversazione on 18 July 1873, Walker states that ‘He had seen one Natural Science Society started in Chester [1858] and make a feeble effort to live, and die in one year. This was the only one that had flourished, and this was entirely due to Canon Kingsley, for whom he called for three hearty cheers’. This early history is discussed in detail by Siddall (1911), Robinson (1971) and Burek (2008). Its initial aim was ‘the promotion of the study of natural science by lectures, field meetings, the reading and discussion of papers and other suitable means’ (Siddall 1911). Its prime purpose was to instruct and enable the ‘mutual communication of knowledge’; three sections were consequently formed: ‘Geological, Botanical and Zoological’ (Robinson 1971). Excursions were a key element in achieving this aim. As the CSNS grew, so did the interests of its members and additional sections were added such as Photography, Literature, Astronomy and Microscopy; each had its own chair and secretary and held lectures and sometimes field excursions. By 1908 there were eight sections, and this increasing diversity of interest is reflected in its change of name (Table 2). By 1880 there were 561 members of the Chester Society of Natural Science and 954 members by 1901 (Stolterfoth 1902). On its inauguration the Society had a strong geological presence, but over time this became more diffuse. The initial emphasis on geology in the field and lecture theatre partly accounts for the Society’s initial growth (Burek 2008). Several famous geologists were associated with the Society, all promoting field excursions, such as Prof. McKenny Hughes, Dr Ethel Skeat, Sir Philip de Malpas GreyEgerton and Prof. Boswell. Ethel Skeat and her work at the Queen’s school teaching geography and Table 2. Chester Natural Science Society name changes Year

Name

1871 1888

Chester Society of Natural Science Chester Society of Natural Science and Literature Chester Society of Natural Science, Literature and Art

1898

99

science and as a researcher in Northeast Wales are noteworthy (Burek & Malpas 2007; Burek 2014). Prof. McKenny Hughes, from the geology department of Cambridge, was the Society’s second President after Kingsley died in January 1875 (Fig. 1); he accepted the Presidency because his father was Bishop of St Asaph and his family connections in the area. He held it for 16 years, when geology and geomorphology blossomed in the CSNS, and he frequently led members’ excursions; he believed they were an effective method of learning about the natural history of an area. The role of geology within the Society, and the consequent geotourism represented by the excursions, displays a trend evident in similar societies at this time and one which was representative countrywide. At its inception in 1871, Dr Henry Stolterforth, a keen geologist and microscopist himself, became the Scientific Secretary, preparing the annual reports for the next 36 years until his death in October 1907; this might account for their very full geological records. In 1871, Charles Kingsley had encouraged membership of the CSNS through his series of geology lectures in the King’s School, Chester. Published in 1873 as ‘Town Geology’ and dedicated to his class members (Kingsley 1873), its subjects were made relevant to a wide audience (as can be seen from the titles) and lent themselves to trips of personal observation: † † † † † †

The soil of the field; The pebbles in the street; The stones in the wall; The coal in the fire; The lime in the mortar; and The slates on the roof.

In the book’s Preface he explained the necessity of teaching general natural history and specifically geology: ‘I know few studies to compare with Natural History; with the search for the most beautiful and curious productions of Nature amid her loveliest scenery. I have known . . . working men who in the midst of smoky cities have kept their bodies, their minds and their hearts healthy and pure by going out into the country at odd hours, and making collections of fossils, plants, insects, birds or some other objects of natural history’ (Kingsley 1873). In his first lecture he starts by explaining geology’s importance: ‘The most important facts of geology do not require, to discover them, any knowledge of mathematics or of chemical analysis: they may be studied in every bank, every grot, every quarry, every railway-cutting, by anyone who has eyes and common sense . . . geology is (or ought to be) in popular parlance, the people’s science.’ This surely is early geotourism and forms the background and context with which the CSNS’s first President sought its establishment.

100

C. V. BUREK & T. A. HOSE

Fig. 1. Professor McKenny Hughes was the second President, after Charles Kingsley, of the Chester Natural Science Society. Picture from the archive of the Geologists’ Association.

THE ROLE OF LOCAL SOCIETIES IN EARLY MODERN GEOTOURISM

The first excursions were often led by Kinglsey and, after his death, by the second President Prof. McKenny Hughes; they had a strong geological/ botanical bias. The first trip went to Helsby Crag to look at the Triassic sandstone; the second went to Nannerach to look at the Carboniferous Limestone and its associated flora. Both were extremely well attended (Siddall 1911). During the summer of 1872, local excursions went to Llangollen, Delamere and further afield to Dolgellau. Five hundred people attended the latter, where Kingsley gave a lecture on the area’s geology and topography. It is debateable how many people heard him, especially as the band of the 14th regiment was playing dance music at the nearby hotel (Siddall 1911). However, this is true geotourism as people enjoyed the whole experience. By 1873, train excursions were very popular. Over 200 members and friends attended the Church Stretton outing; the geologists took charge and led the excursion up the Lightspout Valley and Longmynd (Fig. 2). There were at least two geology excursions a year organized by the Society when the presidents were interested in geology; these included applied geology (e.g. slate mines, coal mines, salt mines,

101

castles, big houses and gardens). Most excursions were fairly local at this time because of the transport available at a reasonable cost in Cheshire, Shropshire, Liverpool and North Wales. There were at least another six trips a year of either a day or half-day. From 1875, evening rambles aimed at younger members were held. The first long excursion was in 1898 (10 –13 June) to Bull Bay, Anglesey; the record photograph (Fig. 3) shows 36 people including two children and several unmarried women, but only the men are named. It is as if the women do not exist, a common plight of many early women geologists (Higgs et al. 2005; Burek & Higgs 2007; Higgs & Wyse-Jackson 2007).

Geotourism in the Woolhope Club and the Chester Society The WNFC’s initial restricted membership of 40 was eventually relaxed and had reached 260 ordinary members by 1907. Its excursions were initially focused on geology, but very quickly had a more countryside and historic focus with especial attention paid to local horticultural matters. Its

Fig. 2. Field trip to Lightspout Valley on 15 July 1873. This drawing of the party scrambling up Lightspout Valley is by Mr Alfred Sumner and shows the esker of glacial sand, which choked the Church Stretton Valley. Courtesy of the Grosvenor Museum, Chester.

102

C. V. BUREK & T. A. HOSE

Fig. 3. Chester Natural Science Society First long excursion,10– 13 June 1898, Bull Bay, Anglesey (Siddall 1911).

Transactions include various geological accounts, and those on botany commonly included geological summaries; the earliest of these were published, to a wider audience than the Transactions, in the Hereford Times. Particularly popular classic geology localities for its excursions included the local Woolhope inlier and farther afield: the Ludlow area of Shropshire; the Malvern Hills of Worcestershire; the Forest of Dean in Gloucestershire; and Abergavenny in Monmouthshire. Mapping the WNFC’s excursions for its first 30 years (Fig. 4) shows its wide-ranging regional coverage; at least one and often two of the four or five field meetings a year had a geological interest during that time. Often a full account of the geology was provided in the Transactions. The Club’s formal breakfast and luncheon arrangements encouraged, as reported in its Transactions, much social and scientific discussion. The excursions also included an appropriate lecture, often covering topography, towards the end. Many excursion accounts reflect a scientific and aesthetic interest in, and acquaintance with, the landscapes in which they took place; for

example, the report on the 20 August 1868 excursion to Woolhope noted that its leader: ‘. . . delivered his address . . . with the aid of diagrams which he supplied to the members, supplemented by his clear explanations, rendered the subject so interesting to his hearers as to teach them that geology consisted in something more than collecting fossils to put upon their shelves, for he led them to observe the wonderful processes of Nature in forming the hills and valleys and soils upon which we had been treading, and of which we now saw such a charming landscape’ (Anon. 1886, p. 55). Such addresses were often delivered to some 30–40 ordinary members, all of whom were listed in the individual excursion reports; of the 29 excursionists on the 1886 Woolhope Friday outing, one was an aristocrat with his two young sons, three were clergy with one guest, three were women and the remainder (all male) were classed as visitors. This healthy number of excursionists and ordinary members continued through the nineteenth century, but had significantly declined by the end of the Great War such that women membership was partly proposed

THE ROLE OF LOCAL SOCIETIES IN EARLY MODERN GEOTOURISM

103

Fig. 4. Woolhope Naturalists’ Field Club field trip locations 1852– 1882.

as a means to address the problem. The WNFC membership lists within its Transactions are inconsistently listed, but a general pattern of growth in the nineteenth century can be noted (Table 3). Table 3. Woolhope Naturalists’ Field Club membership in the nineteenth century Year

Number of Ordinary Members

1852 1866 1870 1881 1889

40 (max) 85 149 177 152

The initial 40-strong membership of the CSNS grew to 454, mostly interested amateurs, at the end of 1872 (Fig. 5); Kingsley stated that only ‘4– 5% of members have scientific instincts’ (Siddall 1911). The excursions were not all about geology but were often linked with other sciences, providing a holistic approach, and there was a social aspect to them. Some localities were very popular with Storeton Woods and Quarries, the site of an ancestral dinosaur footprint find in 1838, being visited in August 1876, July 1888, July 1894 and June 1907. The slate mines at Blaenau Ffestiniog were another favourite site and the first trip in August 1882 was very well attended. Holt and Farndon villages, on either side of the River Dee and navigable by rowing boat from Chester, were favourite places to

104

C. V. BUREK & T. A. HOSE

1400

Chester Natural Science Society Membership levels 1871-1931

1200

1000

Total Membership

800

Ord Hon Corresponding 600

Junior (from1907) Life (from 1914)

400

200

0 1

3

5

7

9 11 13 15 17 19 21 23 25 27 29 31 33 35 37 39 41 43 45 47 49 51 53 55 57 59 Year from start of Society

Fig. 5. Chester Natural Science Society Membership figures 1873–1930.

visit and picnic and saw excursions in May 1877, July 1885, July 1888 and 1901. In 1902, excursions were run to: Beeston Castle in May; Rhydymwyn and the Leet in June; West Kirby and Hilbre Island in July; and Delamere Forest and the Meres in July. Table 4 lists the excursion locations during the first year and 20 years later (Fig. 6). The CSNS promoted enjoyable education for themselves and others at all levels. Its field outings were not limited by locality, geology type, distance from Chester or purpose as its annual reports make clear. For example: ‘It is no small advantage that we always take with us a pleasure section . . .’ and ‘We have again to thank those noblemen and gentlemen who kindly open their grounds to us, and who Table 4. Chester Society of Natural Science excursion locations of 1st and 20th years Summer 1872

Summer 1891

Llangollen Delamere Eaton Gardens Cefn-y-Bedd Dolgellau

Vale of Gresford Vaynol Park, Gwynedd Manchester Ship Canal Rhydymwyn and Leet Nine evening rambles

offer facilities for obtaining rooms where we can take our refreshments’. Further, it developed local evening excursions and reported, for example, that: ‘Our Evening Rambles were again continued in 1884 . . . The committee hope to continue the Evening Rambles this year’ (Stolterfoth 1885). It is interesting to note the importance of food on its excursions, although a much less formal affair than the WNFC: ‘Tea arrangements on all these occasions have been under the management of Mr and Mrs. Blake, and it has been found that by taking out caterers with us we have fared better than when trusting to those at a distance often in remote localities’ (Robinson 1971). In the early years attendance for whole- and half-day trips was ‘seldom less than hundred and the weather with few exceptions has been favourable’ (Robinson 1971). During the 1870s and 1880s there were many weather problems with rain often forcing cancellations, often commented on in the annual reports, for example the ‘1878 excursion cancelled due to flooding and railway bridge collapse’ (Stolterfoth 1879). In 1896 several evening rambles were cancelled due to bad weather. There was a gradual name change for these evening outings; rambles were perhaps not as taxing as field trips and excursions. When the Society’s Museum

THE ROLE OF LOCAL SOCIETIES IN EARLY MODERN GEOTOURISM

105

Fig. 6. Map of Chester Natural Science Society excursion 1872– 1892.

was under construction (1886–1887) attendance at field meetings reduced slightly but increased afterwards: ‘On the whole the excursions were not well attended; on two or three of the days the weather was deplorably wet, and the fact that many of our leading members could not give the time and attention to the excursions which they had done in previous years, was another drawback. This was owing to the extra work entailed in the moving and arranging of our possessions in the new . . . Museum’ (Stolterfoth 1887). At the start of the twentieth century it was noted that: ‘The increased attendance at the Excursions and Field meetings were even more marked and at these many members displayed a keen interest in Botany, Geology and Entomology’ (Miln & Shepheard 1901), but this did not last. Figure 6 depicts the locations of these early field excursions. The declining popularity of CSNS excursions can be attributed to several reasons. They were no longer a novelty, by 1897 the Society was 25 years old and the obvious locations had already been visited several times. In the early years locations were rarely visited more than once. New locations were needed and five new destinations

had been added in 1885. By 1908 the Honorary Secretary had cause to write: ‘Members do not sufficiently realise that these excursions though not perhaps of much use for scientific observations or for collecting purposed, do yet give the best possible opportunities for bringing together those who are interested in the same branches of Natural History, for the comparison of observations, and for the interchange of ideas’ (Fish 1908). After the end of the Great War in 1919, there was an appeal for help in leading rambles because many of the original leaders had died. There were also other entertainments or distractions in Chester: the cinema, radio in the 1920s and eventually early television. By the 1920s, archaeology was also a field of interest attracting much attention by those interested in the countryside. By 1922 there was again popular support for excursions and the summer sessions were well attended, perhaps due to the improved railway facilities. By 1929 there was a decrease in active workers in the Society and most members only attended the winter lectures; there were a few summer excursions, but the popularity of evening walks was generally maintained (Fig. 7). However, purely geological excursions finished in 1939

106

C. V. BUREK & T. A. HOSE

20

Number of excursions

18

16

14

12

Excursions

10

Evening total

8

1915-1916 WWI 6

4

Noinformaon BA meeng 1897 Liverpool

2

0 1

3

5

7

9 11 13 15 17 19 21 23 25 27 29 31 33 35 37 39 41 43 45 47 49 51 53 55 57 59 Years from 1874 (3)-1930 (59)

Fig. 7. Chester Natural Science Society number of Excursions per annum (1874–1930).

as the CSNS became a popular lecture body only, ending nearly 70 years of excursions, field trips and evening walks for non-experts.

Women in the Woolhope Club and the Chester Society The role and involvement of women in the two organizations was markedly different. Women often felt intimidated on excursions (Burek & Ko¨lblEbert 2007a) and some societies discouraged their membership. For a century the WNFC did not permit women members. The CSNS embraced women within their membership and activities from its inception. At the outset the WNFC was for men (of rank) only and was reluctant to embrace women members, despite many of their evening lectures being delivered by women with considerable expertise in their subject. Instead, women’s engagement with the Club was restricted to the one ‘Ladies Day’ field excursion a year, for which admission was by ticket only, and was commonly conducted as a picnic meeting. The first reference to Ladies Days was during the field meeting of 23 August 1853, when it was proposed that ‘ladies should be

admitted to one of the yearly field meetings of the Club’. This was adjourned to the following Annual Meeting, held on 24 January 1854, at which the proposal was ‘deferred for further consideration’. This deferment obviously took some considerable time because women are only recorded as actually attending a field meeting in 1865 on 28 August when the Club visited Usk. The Ladies Days were generally organized for late July or early August, probably in the hope that the expected good weather would not cause them too much discomfort! Unsurprisingly however, such good weather was not always forthcoming; the women (perhaps to the surprise of some of the men?) seemingly coped with the inevitable underfoot wet and mud and were sometimes admirably commended by their hosts. The report of the 18 July 1867 Craig-Y-Pwll-Ddu field meeting included a comment from the manager of a nearby field club: ‘You will never be able to go on with “Ladies’ Days”. We had them at one time but the weather was invariably wet, and we were obliged to give them up.’ The 11 July 1870 Ladies’ Day was reported in the Transactions in an illuminating way ‘. . . ladies arrayed in a bewitching varieties of . . . costume, bespoke at once the meeting of the year when absorbed naturalists are

THE ROLE OF LOCAL SOCIETIES IN EARLY MODERN GEOTOURISM

107

Table 5. Chester Natural Science Society women’s membership Year

Total membership

Women (%)

Number of women

Honorary members

1871 1872 1873 1884 1890

356 454 502 616 661

41.8 41.6 39.4 28.3 33.7

142 189 198 174 223

0 0 0 0 0

conspicuous by their absence’. It seems that choice of venue for the Ladies Days took some account of the consideration the terrain, especially the feasibility of clambering around in cumbersome long skirts, and walking distance.

When the matter of women’s WNFC membership was again discussed in 1918 it was rejected because the ‘indiscriminate admission of ladies would seriously interfere with the scientific objects for which the club had been formed’. Additionally,

Table 6. Woolhope Field Naturalists’ Club joint excursions in its first 30 years Year

Date

Location

1853 1854 1855

7 June 13 June 12 June

Eastnor Forest of Dean Malvern, Worcestershire Beacon Tarrington Ledbury Eastnor, Dog Hill, Ledbury Ludlow, Mocktree, Leintwardine

1857 1858 1860

2 June 3 June 15 May

1861

23 – 24 May

1862 1862 1863

22 May 18 September 7 July

Ledbury Church Stretton Craven Arms, Ludlow

1863 1864 1865

7 September 21 July 18 – 19 July

1866

12 September

Malvern Link Ross Ludlow and Craven Arms Malvern

1867

6 August

Craven Arms for Clun

1868

19 June

1868

28 July

1870 1871 1872 1875 1876 1876

19 August 25 May 17 May 23 May 17 July 17 July

1877

19 June

1880 1881

20 May 19 May

Crumlin Bridge and Pontypool Ludlow for Titterstone Clee Hill and Oakley Park Longmynd Hay on Wye Great Malvern Symond’s Yat Stoke Edith Old Radnor for Stannor Rocks Hollybush Pass, Midsummer Hill Herefordshire Beacons Tewkesbury Abbey and battlefield

Notes Malvern Field Club and Cotteswold Field Club Malvern Field Club and Cotteswold Field Club Cotteswold, Malvern, Warwickshire, Worcester Malvern Field Club Malvern Field Club and Worcestershire Naturalists’ Club With Malvern Field Club and Worcestershire Naturalists’ Club With Malvern Field Club. Poorly attended by WFNC due to poor train service as the summer timetable was not running. Malvern Field Club Dudley Field Club and the Oswestry Field Club Bridgenorth Club, Caradoc Field Club, Dudley Field Club and Oswestry Club Meeting of United Field Clubs Malvern Field Club and Cotteswold Field Club Caradoc Field Club, and Dudley & Midland Geological Society Malvern Field Club and Cotteswold Field Club, Worcestershire Naturalists’ Club Caradoc Field Club invitation to meet with the British Archaeological Association Cardiff Naturalist’s Society Caradoc Field Club Caradoc Field Club Malvern Field Club Malvern Field Club Cotteswold Naturalists’ Field Club Dudley and Midland Geological and Scientific Society Caradoc Field Club Caradoc Field Club and Cotteswold Field Club Malvern Field Club Malvern Field Club

108

C. V. BUREK & T. A. HOSE

admitting ladies would likely increase the membership, thus making field days unwieldy and reduce the existing ladies days to mere picnics. A Mr Watkins, in supporting the admission of women, drew attention to the decreasing number of members who were competent microscopists and photographers; indeed, he considered the craft of the latter to be dead and likewise the study of botany and mycology. He noted that women were then taking an interest in the very things that men were dropping and also taking a greater interest in crafts, techniques and science than had been the case in 1918. A Special General meeting was held on 6 May 1931 to consider the admission of women to the WNFC. At this meeting the Honorary Secretary pointed out that the Club’s then existing rules barred ladies from membership, although there was no mention of ladies or gentlemen in them ‘it was certainly the intention of the founders of the Club that . . . ladies should not become members’ but that in the minutes of 1866 ‘a resolution had been unanimously passed that ladies be invited to attend the First Meeting to be held at Ross’. He

further conceded that ladies’ names had several times been submitted but the Central Committee would not sanction their election and stridently argued that ‘if you are going to drive a coach and horses through the rules . . . You would be able to admit ladies, gorillas, and babies in arms!’ Women continued to attend just one annual field meeting until eventually admitted to membership at a special meeting on 14 January 1954 when the proposal was carried by 43 votes to 22; it was endorsed at a General Meeting on 21 January 1954. After 100 years, the Club ceased its open discrimination. The CSNS was much more enlightened in its approach to women’s membership. An analysis of its membership shows that, between 1873 and 1911 for example, the female membership remained at roughly 25–30% (Table 5). This allowed women to participate in geotourism through an acceptable social medium in Victorian and Edwardian England where protocol was paramount. The inclusion of women from the very beginning of the Society allowed them to participate in all activities

Fig. 8. Chester Natural Science Society Joint trip with Liverpool Geological Society (1927). Courtesy of the Grosvenor Museum, Chester.

THE ROLE OF LOCAL SOCIETIES IN EARLY MODERN GEOTOURISM

including lectures (presenting as well as attending), management (the first member of the Management Committee serving from 1907– 1925 was Beatrice Clay, headmistress of the Queen’s School, Chester, who was made an honorary member when she moved away from Chester) and excursions (Burek 2008). The importance of locality and the difficulty of travel for women during the nineteenth century may account for an imposed local interest (Burek & Ko¨lbl-Ebert 2007a, b). However, women also influenced geotourism by involving their children and indeed the whole family in the outings as an enjoyable way of spending time together, as happened in the CSNS; there are also examples of spinster sisters enjoying the freedom of outings through its excursions. Again, the CSNS’s open acceptance of women fostered enhanced travel opportunities. It is not freedom of travel which restricts geotourism activities today for women, but perhaps time and family commitments as in the past.

Significant similarities of the Woolhope Club and the Chester Society While focusing on their specific counties, both the Club and Society also adopted a regional role in their scientific endeavours and trips, occasionally enjoining with kindred local bodies for fieldwork. For example, the WNFC jointly met with the

109

Caradoc and Cotteswold Clubs in 1877 and the Malvern Field Club in 1881 (Table 6). The first joint annual trip for the CSNS was with the Chester and North Wales Archaeological and Historic Society (CNAHS) in 1894 to Bryntisilio, and then again with them in 1896 to the Bethesda Slate quarries. The Naturalists’ Field Club of Liverpool was very supportive. The Society also entertained other local groups and the Liverpool Geological Association visited its Grosvenor Museum in 1886 (Tresise 2013), acting as the focus for other geotourism events (Fig. 8). A close association was established not only with the Liverpool Societies but also the Wrexham Society which became affiliated in 1872. This allowed free access to each other’s lectures and meetings. The links between the latter two societies went even further: the Annual reports were printed alongside each other but as separate sections during 1874 –1879. Like the WNFC they also entertained the Dudley & Midland Geological Society in September 1878 (Table 6). The interaction between these local societies allowed a greater understanding of the context of the local area. Both organizations employed a variety of transport methods including private horse carriage (especially for the WNFC) and trains (especially for the CSNS). The WFNC employed a private coach for its third field geology-focused field meeting in 1853 (to Kington, Presteign and Stannor Rocks) when ‘. . . the exceeding beauty not less than

Fig. 9. Dee river steamer (Welsh government educational sources).

110 C. V. BUREK & T. A. HOSE Fig. 10. Left: Honorary Members (1866) of the Woolhope Field Naturalists’ Club, as published in the Club’s first full published set of Transactions in 1866. Notable on the list are the names of key early geologists such as Murchison, Sedgwick, Lyell, Lindley, and Phillips. Also of interest, in fostering joint work, are the inclusion of other field club officials. Right: Honorary Members (1873) of the Chester Society of Natural Science.

THE ROLE OF LOCAL SOCIETIES IN EARLY MODERN GEOTOURISM

111

Fig. 11. Hereford Free Public Library and Museum (1874 WFNC Transactions frontispiece). Originally published in the Gardeners’ Chronicle and presented as the frontispiece of the 1874 Transactions with the note that it showed some Woolhopians, presumably the figures on the pavement.

the scientific interest of the country, made the meeting altogether a most delightful one. The Hereford party of members assembled at 6 a.m. at the Green Dragon Hotel, from which they speedily departed on a small coach appropriated to them . . .’ (Edmunds 1853). The railways fully reached Hereford in 1853 and Chester in 1847. This opened up good communications from Hereford to Ludlow, Malvern, Shrewsbury, Ross, Worcester, Newport and Abergavenny and similarly from Chester to the whole of the North Wales coast and to the Wirral, Liverpool, Crewe and Birmingham (Figs 4 & 6). For the CSNS: ‘In some instances, short journeys by train took us to lanes and fields and ponds in free air, and less frequented than the immediate neighbourhood of the City. These short trips were well appreciated, attracting even larger numbers

than our more distant and elaborately planned excursions’ (Stolterfoth 1885). Bicycles were used occasionally, for example for the CSNS excursions to the Wirral in 1900 and Tarvin in 1903. It was never very popular however as the Geologists’ Association also contemporaneously discovered between 1898 and 1903; the bike excursions ceased in 1907 when ‘No special cycling excursions have been arranged, they not having been well attended in former seasons’ (Sweeting 1958). This was exacerbated with the advent of the motor car and charabanc. CSNS participants also used the River Dee steamer (Fig. 9) for the 1878 trip to Hilbre island in the Dee estuary (only accessible on foot at low tide) and in 1880 for the trip to Eaton. Both organizations elected distinguished honorary members to establish their social and academic

112

C. V. BUREK & T. A. HOSE

Fig. 12. The Grosvenor Museum, Chester. ‘The work of the Contractor, Mr Gabbutt, has been carried out under the superintendence of our able Architect, Mr Lockwood, who has spared no pains to meet the varied requirements for which the whole building is designed. The united Committees of the various Societies have met frequently, and a special Sub-Committee has met every week, to see that the work was properly carried out’ (Stolterfoth 1887, pp. 6 –7). Courtesy of the Grosvenor Museum, Chester.

status (Fig. 10) and published accounts of their activities. The WNFC more or less continuously published its Transactions from 1866 (although these incorporated material from 1852 onwards); publication continues today. The CSNS published Annual Reports from its inception until 1931 (59th annual report) and scientific proceedings intermittently during 1874–1907 (Robinson 1971). The CSNS briefly restarted publication of these in 1947 with a memorial volume (Report and Proceedings, December 1947) dedicated to the late Robert Newstead, its first museum curator. However, the venture was short-lived with the last volume (subtitled Cheshire, North and Mid-Wales Natural History, Vol. 5) published in 1954, covering the period 1951– 1953. The volumes were especially focused on Lepidoptera and the only geological

paper was on the non-marine mollusca of the Caerwys Holocene tufa deposit. Following suggestions at several 1870 field meetings, the Reverend R. H. Williams worked up a proposal that the WNFC should offer annual prizes for specimen collection of dried plants, insects, etc., made in the course of each year in any parish connected with the Club and submitted by the 15 December. The proposal was discussed and adopted at its 1873 Ladies day meeting at Haye Park: ‘. . . but as the time between the August meeting and the end of the year would be somewhat short to begin the prize-system with next year, and not the present, each prize to be not less than one guinea’ (Salwey 1873, p. 86). Interestingly, this was also the meeting at which T.J. Salway delivered his lecture ‘The Geology, History, and

THE ROLE OF LOCAL SOCIETIES IN EARLY MODERN GEOTOURISM

Natural Features of the Neighbourhood’ on the Ludlow area. By 1875 the prize had been amended, on admirable conservation grounds, to exclude birds’ eggs. The CSNS instituted its Kingsley Memorial Prizes in 1877. One, up to the value of £10 was for a collection of objects in a single branch of natural history from within the Society’s area. To further fieldwork, another prize of up to £10 was offered for collections, essays or notes of observations illustrative of the natural history of the Society’s area (Burek 2008). A Medal was also awarded for outstanding contributions, with Charles Kingsley being its first and post-humous recipient in 1877. Both organizations therefore promoted fieldwork and scientific endeavour through monetary rewards. Sir James Rankin, a wealthy member and past President of the WNFC, offered to pay for a public library and museum in 1871. This building, as the Hereford Free Public Library and Museum

113

with a purpose-built meeting room for the Club, was opened in 1873 (Fig. 11). Prior to this, The Hereford Free Library and Museum Committee had organized a temporary Free Library and Reading-room in King Street. In 1871 the WNFC appointed a committee to act with the Town Council with a view to establishing a scheme for a Free Library Museum; under the powers of the Public Libraries Act 1850 a site was purchased following approval at a public meeting in July 1871. It became apparent however that while the Town Council would pay in full for a library, the museum (allowed under the Museums Act 1850) would only be possible as an addition through the endeavours of WNFC. Somewhat later, the 1st Duke of Westminster (who owned much of central Chester) donated a plot of land inside the city walls along with £4000 towards the establishment of a new museum building (the total cost being £11 000). Opening in 1886 (Figs 12 & 13), this

Fig. 13. Grosvenor Museum Building Committee (1886) (Siddall 1911). This photo contains many of the leading members of the CSNH of the time.

114

C. V. BUREK & T. A. HOSE

museum served both the CSNS and the CNAHS. Within a year of its opening Dr Howson, one of the CSNS’s first Vice-Presidents who had been associated with the Society and its Founder from its earliest days, died. He had been Secretary to the Chester Museum of Natural History and had a major role in raising funds towards its completion. However, it was reported that: ‘The year 1886 will mark a crisis in the existence of the Society. Since the first formation of the Society we have had no permanent home in which to place our

collections. . . . There is still, we regret to say, more than £1500 to be collected before the structure is paid for . . .’ (Stolterfoth 1887, p. 6). Both organizations established museums and associated libraries, despite the challenging and ongoing financial situations involved in the last quarter of the nineteenth century that survive to the present day. The CSNS’s Curator could report by 1894 that the museum housed 11 000 specimens and a new wing was under construction. Due to funding issues, both museums were managed by their respective

Table 7. The Woolhope Club and Chester Society compared Society When formed Purpose

Woolhope Naturalists’

1st President

May 1851 ‘. . . for the practical study, in all its branches, of the Natural History of Herefordshire and the districts immediately adjacent.’ Mr R.M. Lingwood (1852)

2nd President

Reverend T.T. Lewis (1853), a geologist

3rd President

Reverend William Samuel Symond Williams (1854), founder member and geologist These early presidents set the Club’s tone and its field excursions, but over time . . .

Comments

Chester Society 26 May 1871 ‘To promote the study of natural science by lectures, field meetings, the reading and discussion of papers and other suitable means.’ Canon (Sir) Charles Kingsley (1871– 1875), a botanist and geologist Prof. Thomas McKenny Hughes, a geologist (1875– 1891) Alfred Oston Walker J.P. F.L.S. (1891 – 1892), meteorologist and founder member These early presidents set the Society’s tone and its field excursions: ‘educate the people’

Logo

The logo has a geology hammer and collecting bag, and a cross-section of the Woolhope inlier

Museum buildings

Publications

1871: Sir James Rankin offered to pay for a ‘Public Library and Museum in connection with the Woolhope Naturalists’ Field Club’ 1873: Opened with a purpose-built meeting room for the Club Transactions, more or less continuous throughout its existence

The logo incorporates the initials of Charles Kingsley. The research work of founder member and the 6th president, microscopist Dr Henry Stolterfoth, led to the diatom Eucampia zodiacus being used to encircle these initials in the badge. He was a geological enthusiast. 1885: 1st Duke of Westminster (4th President) gave the land and £4000 of £11 000 cost. 1886: Opened by Professor Robert Newstead as 1st Curator Annual Reports, continuous until end of nineteenth century then a few in late 1940s. Scientific Proceedings from 1874 and five further volumes until 1907

THE ROLE OF LOCAL SOCIETIES IN EARLY MODERN GEOTOURISM

town councils by the 1930s. Table 7 draws together a comparison between the two societies, illustrating the wider applicability of local natural history societies to early geotourism.

Concluding comments The CSNS offered countryside evening and weekend excursions with a holistic approach to nature to a wide range and number of people, encouraging landscape tourism. The majority of the society members were not experts but casual excursionists. The WNFC offered weekday excursions, with a holistic approach to landscapes, to a small number of professional and elite casual excursionists. For both, the numbers on excursions grew during the nineteenth century and significantly declined after the Great War. For both, the focus of excursions turned away from geology and especially towards archaeology after the Great War. The excursions fulfilled the original founders’ and local organizations’ aims to promote a greater knowledge and understanding of the countryside and its conservation. The Great War also coincided with the arrival of other forms of entertainment, especially cinema and then, in the 1920s, the radio. Local societies had by then fulfilled the role of predecessors to modern organized geotourism; their legacy of publications and sites is still around today. They helped too when the environmental movement emerged in the post-1950s to underpin modern geotourism, which was finally recognized in the 1990s. The authors freely acknowledge the support and assistance of colleagues from the two organizations. CB would especially like to thank Dr K. Riddington of the Grosvenor Museum for her unfailing support. TAH would especially like to thank J. Jonson for information on key aspects of the WNFC’s Ladies’ Days and women’s membership. Both authors are grateful for assistance from various libraries and their staff; for TAH that of the Geological Society is noteworthy.

References Anon. 1886. Third Field Meeting, August 20th – Geology, Perton Lane to St. Ethelbert’s Camp (with plan of Camp, facing page 60). Transactions of the Woolhope Naturalists’ Field Club, 1886, 52– 56. Burek, C. V. 2008. The role of the voluntary sector in the evolving geoconservation movement. In: Burek, C. V. & Prosser, C. (eds) The History of Geoconservation. Geological Society, London, Special Publications, 300, 61–89. Burek, C. V. 2012. The role of LGAPs (Local Geodiversity Action Plans) and Welsh RIGS as local drivers for geoconservation within geotourism in Wales. Geoheritage, 4, 45– 63.

115

Burek, C. V. 2014. The contribution of women to Welsh geological research during the later nineteenth and early twentieth century. Proceedings of the Geologists’ Association, 125, 480 –492. Burek, C. V. & Higgs, B. 2007. The role of women in the history and development of geology: an introduction. In: Burek, C. V. & Higgs, B. (eds) The Role of Women in the History of Geology. Geological Society, London, Special Publications, 281, 1 –8. Burek, C. V. & Ko¨lbl-Ebert, M. 2007a. The historical problems of travel for women geologists, Women in the History of Geology V. Geology Today, 23, 30–32. Burek, C. V. & Ko¨lbl-Ebert, M. 2007b. The historical problems of travel for women undertaking geological fieldwork. In: Burek, C. V. & Higgs, B. (eds) The Role of Women in the History of Geology. Geological Society, London, Special Publications, 281, 115–122. Burek, C. V. & Malpas, J. A. 2007. Rediscovering and conserving the Lower Palaeozoic ‘treasures’ of Ethel Woods (nee´) Skeat and Margaret Crosfield in northeast Wales. In: Burek, C. V. & Higgs, B. (eds) The Role of Women in the History of Geology. Geological Society, London, Special Publications, 281, 203– 226. Burek, C. V. & Prosser, C. D. 2008. The history of geoconservation: an introduction. In: Burek, C. V. & Higgs, B. (eds) The History of Geoconservation. Geological Society, London, Special Publications, 300, 1– 6. Edmunds, F. 1852. Woolhope Field Naturalists’ Field Club from the ‘Hereford Times,’ 22nd May, 1852. Transactions of the Woolhope Naturalists’ Field Club, 1852, 1– 2. Edmunds, F. 1853. Third Field Meeting of the Woolhope Naturalists’ Field Club. Hereford Times, Hereford (23 August). Fish, A. H. 1908. Thirty seventh Annual report and Proceedings for the year 1907– 1908. Chester Society of Natural Science, Literature and Art, Chester. Higgs, B. & Wyse-Jackson, P. 2007. The role of women in the history of geological studies in Ireland. In: Burek, C. V. & Higgs, B. (eds) The Role of Women in the History of Geology. Geological Society, London, Special Publications, 281, 137– 153. Higgs, B., Burek, C. V. & Wyse Jackson, P. N. 2005. Is there gender bias in the geological sciences? Irish Journal of Earth Sciences, 23, 132– 133. Hose, T. A. 1994. Telling the story of stone – assessing the client base. In: O’Hallaran, D., Green, C., Harley, M., Stanley, M. & Knill, S. (eds) Geological and Landscape Conservation. Geological Society, London, 451–457. Hose, T. A. 1995. Selling the story of Britain’s stone. Environmental Interpretation, 10, 16–17. Hose, T. A. 2000. European geotourism – Geological interpretation and Geoconservation promotion for geotourists. In: Barretino, D., Wimbledon, W. P. & Gallego, E. (eds) Geological Heritage: Its Conservation and Management. Instituto Technologico Geominero de Espana, Madrid, 127– 146. Hose, T. A. 2007. Leading the field: a contextual analysis of the field-excursion and the field-guide in England. In: Wickens, E. W., Hose, T. A. & Humberstone, B. (eds) Proceedings of the Critical Issues in Leisure

116

C. V. BUREK & T. A. HOSE

and Tourism Education Conference. Buckinghamshire Chilterns University College, High Wycombe, 115– 132. Hose, T. A. 2008. Towards a history of geotourism: definitions, antecedents and the future. In: Burek, C. V. & Prosser, C. D. (eds) A History of Geoconservation. The Geological Society, London, Special Publications, 300, 37–60. Hose, T. A. 2011. The English origins of geotourism (as a vehicle for geoconservation) and their relevance to current studies. Acta Geographica Slovenica, 51, 343– 359. Hose, T. A. 2012. 3G’s for modern geotourism. Geoheritage, 4, 7 –24. Kingsley, C. 1873. Town Geology. Appleton & Co., New York. Knell, S. J. 2000. The Culture of English Geology 1815– 1851: A Science Revealed Through Its Collecting. Ashgate, Aldershot. Miln, G. P. & Shepheard, W. F. J. 1901. Hon. Secretaries report. 30th Annual Report and Proceedings for the year 1900– 1901. Chester Society of Natural Science, Literature and Art, Chester, 5– 7. Purchas, W. H. & Ley, A. 1889. A Flora of Herefordshire. Woolhope Naturalists’ Field Club, Hereford. Robinson, H. 1971. Chester Society of Natural Science, Literature and Art – The First Hundred Years 1871– 1971. Chester Society of Natural Science, Literature and Art, Chester.

Salwey, T. J. 1873. The meeting at Ludlow, Richard’s Castle, and Haye Park: the geology, history, and natural features of the neighbourhood. Transactions of the Woolhope Naturalists’ Field Club, 1873, 83–88. Siddall, J. D. 1911. The Formation of the Chester Society of Natural Science, Literature and Art and an Epitome of Its Subsequent History. Chester Society of Natural Science, Literature and Art, Chester. Stolterfoth, H. 1879. Hon. Scientific Secretary’s Report. Eighth Annual Report and Statement of Accounts for the year 1884– 5. Chester Society of Natural Science, Chester. Stolterfoth, H. 1885. Hon. Scientific Secretary’s Report. Fourteenth Annual Report and Statement of Accounts for the year 1884– 5. Chester Society of Natural Science, Chester. Stolterfoth, H. 1887. Hon. Scientific Secretary’s Report. Sixteenth Annual Report of the Chester Society of Natural Science and Statement of Accounts for the year 1886– 7. Chester Society of Natural Science, Chester. Stolterfoth, H. 1902. Hon. Scientific Secretary’s Report. Thirty first Annual Report and proceedings for the 1901–1902. Chester Society of Natural Science, Chester. Sweeting, G. S. 1958. The Geologists’ Association 1858– 1958. Geologists’ Association, London. Tresise, G. 2013. The Liverpool Geological Association 1880– 1910. North West Geologist, 26– 34.

Geotourism: an early photographic insight through the lens of the Geologists’ Association JONATHAN G. LARWOOD Natural England, Unex House, Bourges Boulevard, Peterborough PE1 1NG, UK (e-mail: [email protected]) Abstract: From its earliest days in London in 1858, the Geologists’ Association (GA) brought together people from all backgrounds – amateur and professional geologists, men, women and children – to share their enthusiasm for geology and their desire to seek out and explore the geological world around them. The travels of the Geologists’ Association, in search of geological enlightenment, are documented in the Proceedings of the Geologists’ Association. These include accounts of organized excursions, detailing the geology seen and describing the discussion had, the refreshments taken and the transport used. Bringing these accounts to life is the Geologists’ Association’s Carreck Archive, which provides a rare insight into the world of the early geotourist documenting both familiar and lost places. Much is owed to the skill of the photographers such as T. W. Reader, whose albums document the field meetings between 1907 and 1919, while the spirit of the early travelling Geologists’ Association is captured in the albums of Miss M. S. Johnston. This paper explores the early travels of the Geologists’ Association through the literal views of the Carreck Archive and accounts in its literature, and the establishment of the GA as an inadvertent geotourism agent.

The Geologists’ Association (often referred to as the GA) was founded in 1858 with the aim ‘. . . to enable the practical student in geology to find a congenial place where doubt may be stated and experience exchanged’. These are the words of the Provisional Committee of the GA in their prospectus issued to attract its first members. The intention was, in particular, to provide ‘a means of intercommunication among those who, while not devoting their lives to the pursuit [of geology], yet take an active interest in its facts and teachings’, making a distinction from the more advanced and strict ‘scientific method and treatment’ of the Geological Society. From the onset, the Association embraced members from town and country and ladies were eligible for election. It is under this ethos that the seed of the GA as a promoter and enabler of geotourism was sown. Uniquely, the GA has brought together the professional geologist, often leading field visits, and the amateur geologist from a range of backgrounds. It has been this blend of the ‘dedicated’ (the professional) and ‘casual’ (the amateur) geotourist (as defined by Hose 2000, 2008) that has enriched the GA field visits with many of the casual becoming dedicated, often leading excursions and becoming significant contributors to geology. Detailed accounts of the history of the Association can be found in Sweeting’s (1958) centenary account and The Wyley History (Leake et al. 2013) of the following 50 years. Here the focus is the role of the GA in the early development of geotourism, in particular, the visual record of field activities documented through the Geologists’ Association Carreck Archive.

Field meetings: the establishment of GA geotourism Field meetings have always been central to the GA. Professor W. W. Watts (1909), then GA President, wrote in celebration of 50 years of the Association ‘If the papers read represent the mind of the Association the field work is its soul’. Green (1989a) provides a detailed account of the development of excursions and a brief summary follows here. The Reverend Thomas Wiltshire (1859), then GA President, wrote in response to a request that the GA ‘institute field-meetings’. Wiltshire acknowledges that ‘. . . a lecture given at a natural section, and on the fossils in situ, is far more valuable and instructive than one illustrated by the most expensive diagrams’. He then proceeds to invite members to a future ‘. . . geological ramble, and to spend a summer’s-day both pleasantly and profitably’, and so is established the ‘soul’ of the Geologists’ Association. It is also noted in the GA Annual Report of 1866 (Jones 1880) that ‘. . . good Excursions require 1. Good planning; 2. Good leaders; and 3. Good reporting; and that it is a pity to lose the results of the Excursion for want of the last’. It is the value placed on reporting that has secured the GA record of field excursions (and geotourism) that exists today. The first reported field meeting is to Folkestone on 9 April 1860 to examine ‘the chalk, greensand and gault, as exhibited in Eastwear Bay, the Warren and Copt Point’ (Wiltshire 1860). Field meetings become a regular part of GA activity from then on and by 1871 a pattern developed of Easter,

From: Hose, T. A. (ed.) 2016. Appreciating Physical Landscapes: Three Hundred Years of Geotourism. Geological Society, London, Special Publications, 417, 117–129. First published online December 18, 2014, http://dx.doi.org/10.1144/SP417.7 # The Geological Society of London 2016. Publishing disclaimer: www.geolsoc.org.uk/pub_ethics

118

J. G. LARWOOD

Whitsun and ‘Long’ or summer field meetings (Green 1989a). In 1886 the formal role of the ‘Field Meetings Secretary’ was established with the appointment of J. Foulerton. Initially concentrated in the south, field meetings soon extended northwards; the first two-day meeting (27–28 May 1863) to the West Midlands was undertaken when a party attended the Anniversary Meeting of the Dudley Geological Society (Sweeting 1958). In 1867 some GA Members (by invitation) took part in a meeting of geologists in Paris, field excursions and a visit to the International Paris Exhibition (Jones 1880). This seeded the early idea of the benefit of foreign field meetings and an important French connection. In 1878 the first foreign organized field meeting to the Boulonnais, northern France, was announced and reported (Barrois 1879). By 1910 over 1000 field meetings are recorded in the Proceedings and Sweeting (1958) provides a full list fitting the ‘tripartite’ pattern up until 1957.

Field meetings: ‘good reporting’ and the development of the ‘guide’ Good reporting was established as critical and there are three important and, mutually supportive, aspects to recording GA field meetings. First of these aspects is the announcements of planned field meetings. Initially (between 1858 and 1870), this was through notices delivered by post to members. With the need to supply more information about forthcoming meetings and excursions the Geologists’ Association Circular was established in 1871 and issued on a monthly basis (Robinson 1989b). It provided detailed announcements of field meetings covering itinerary, travel and accommodation arrangements, costs, where necessary an indication of recommended field clothing and the contact details of the field meeting organizer (the Field Meeting Secretary). The second important aspect is the reports of field meetings in the Proceedings of the Geologists’ Association, Watt’s ‘mind’ of the GA (Robinson 1989a). For most field meetings a full report was written up in the Proceedings by the leader of the visit. Typically, they would outline what was visited and when, who the field party met and were hosted by, an account of the geology, what was found and discussed (often naming who contributed to the debate) and anecdotal information such as the mode of transport, weather conditions and the refreshments consumed. The final aspect of the records is the development of geological guides. In 1891 the GA aggregated in one volume a Record of the field excursions made between 1860 and 1891 (Holmes & Sherborn 1891). This was more than a simple compilation.

Excursions were ordered by geographical region, additional illustrations added to the original accounts and the volume indexed by location. Holmes notes that though he did not consider this to be an adequate geological guide, it did fill a gap. Where other guides provided general descriptions and comparisons of county geology (Ramsay’s Geology of Great Britain, 4th edition 1874, Woodward’s Geology of England and Wales 1887 and Harrison’s Geology of the Counties of England and Wales 1882 are cited), the GA’s Record provided detailed descriptions of individual localities and details of how to spend your visit most profitably with the inclusion of itineraries. The value of the Record in the field was also recognized with deliberate page breaks included so that the volume could be split into three and rebound at a size easier to fit into the pocket. It was also suggested that the black-and-white map and section illustrations could be made more distinctive by the reader with the addition of colour. In 1910 the Geologists’ Association published ‘The Jubilee Volume’: Geology in the Field (Monckton & Herries 1910). Covering England and Wales the publication is again arranged geographically. It uses previous excursions as a basis for new descriptions compiled by geologists, with the relevant specialist knowledge of each area covered (29 contributors in total; Sweeting 1958). Occasional pamphlets were also published to accompany the long field meetings. One of the earliest was issued in 1885 for the field meeting to Belgium and the French Ardennes (Greensmith 2008). In 1934, pamphlets were published for the Easter visit to the Gloucester District and the summer field meetings to Norway and NE Yorkshire (Sweeting 1958). The Norway pamphlet was 71 pages long with 23 plates and text figures and included a pocket for a duplicate set of plates and tables for use in the field. Marking the centenary of the Association in 1958, the formal publication of a set of excursion guides was embarked upon: The Centennial Guides of the GA (Greensmith 2008; Hose 2008; Leake et al. 2013). The first guide published was to the Birmingham District (Hardie et al. 1958) and the most recent, the 71st, to the The Coast of the Bristol Region: Quaternary Geology and Geomorphology (Case 2013).

The Carreck Archive: seeing the GA as geotourist The GA Circular provides the announcement of field meetings and the Proceedings has traditionally reported what was seen and discussed. Uniquely, however, the Geologists’ Association has a photographic archive of its field activities: the Carreck

THROUGH THE LENS OF THE GEOLOGISTS’ ASSOCIATION

119

Fig. 1. Pages from Miss M. S. Johnston’s two albums depicting the 1925 GA visit to Shropshire (Watts 1925): (a) the group with numbered reference (M. S. Johnston is third from left on the back row); and (b) key to (a), geologists inspecting a section and the group having lunch on the Longmynd.

120

J. G. LARWOOD

Archive. Dating back to the 1860s the Carreck Archive contains photographs (many compiled in albums), letters, postcards and a range of ephemera that record the field meetings and key events of the Geologists’ Association. Photography has always been an important part of the record of the GA (Green 1989b). The earliest photograph to appear in the Proceedings is a stereoscopic pair of the Iguanodon Quarry in Maidstone (Bensted 1860). This is also believed to represent the first use of a photograph to illustrate a geological journal in Britain (Green 1989b). Expense prevented the further inclusion of photographs in the Proceedings until 1885 when the simplification of photography and consequent popularization, alongside adaptation of photographs for printing, established photography as a key illustrative tool. From then on the Archive grew significantly with the development of ‘film’, replacing the need to carry boxes of glass plates and chemicals in the field.

With advances such as the introduction of the Kodak Brownie in 1901, the camera became more portable and widely available. In 1892 the Geologists’ Association’s Photography Sub-Committee was established with the aim of making available to members photographs taken during field meetings (Green 1989b). This was the beginning of the Association’s photographic archive.

Key people in the development of the Carreck Archive Willliam Pinckard Delane Stebbing (1873– 1961) joined the GA in 1893; he was a polymath and, among his many interests, was a keen photographer. In 1902 he placed a request for photographs in the GA Circular. His intention was to provide a record of changing field sections and establish a collection to inform members and, eventually, be used

Fig. 2. Typical page from a T. W. Reader Album illustrating the excursion to Hertford and Stevenage (2 May 1914) (Hill 1914) and the journey on the railway line (under construction) from Hertford to Knebworth.

THROUGH THE LENS OF THE GEOLOGISTS’ ASSOCIATION

by members at lectures. In 1903 Stebbing became Curator of the Album and can be considered the first archivist of the Geologists’ Association. Miss Mary Sophia Johnston (1875– 1955) joined the Association in 1898 and was secretary to the Illustrations and Photographic Committee from 1910 to 1925 (Sweeting 1958). From Stebbing she took responsibility for maintaining the Association’s albums and exhibited them at the Association’s Annual Conversazione (subsequently the GA Reunion and now the Festival of Geology). Her own personal albums of photographs, drawings, autographs, letters from friends and other personalia reflected her involvement with the Geologists’ Association and the British Association for the Advancement of Science. These two albums were bequeathed to the Association and provide a record of field meetings she attended between 1890 and 1937 (Leach 1956; Carreck 2007). M. S. Johnston can be considered the second archivist of the

121

GA. Uniquely, M. S. Johnston’s albums provide a detailed record of field meeting participants with careful labelling of the many groups and individuals (Fig. 1a, b). Thomas William Reader (1860– 1923) joined the Association in 1903. He attended every GA field meeting between 1907 and 1919, meticulously photographing and recording the localities visited which he compiled in a series of 12 albums that were subsequently donated to the Association (Figs 2 & 3). Reflecting the skill, quality and contribution of T. W. Reader’s photography, more than 100 of his photographs were published in the Proceedings. In 1920 he became the first recipient of the Geologists’ Association’s Foulerton Award (Leach 1934) for work of merit connected to the Association. Following the death of M. S. Johnston, Marjorie Carreck took responsibility for the care of the GA albums including those of M. S. Johnston. Marjorie

Fig. 3. T. W. Reader Album, excursion to the Royal Albert Docks, 21 March 1914, showing excavations and ‘. . . Snags or Yew trunks brought down by the old river’ being left to dry.

122

J. G. LARWOOD

can be considered the third archivist of the GA and fulfilled this role until 2010. Marjorie continued to grow the archive putting together two new albums, one of which is devoted to field meetings from the 1920s to the 1980s (Fig. 4). Over this period many members donated photographs to the Archive encouraged by Marjorie, who exhibited the collection at every GA Reunion. On her retirement from the role the archive was renamed the Geologists’ Association Carreck Archive, recognizing Marjorie’s contribution over 55 years. From 2010 the author became the fourth archivist.

Carreck Archive: examples of GA geotourism The GA’s Carreck Archive brings together over 150 years of contribution and includes albums, photographs, postcards, letters, menus and a range of associated ephemera. It provides a view of the Geologists’ Association as geotourist and, in particular, the growth of the organized field meeting.

The activities of the GA are comprehensively documented through the Proceedings, Circular and minutes and reports from meetings of the Association’s Council. It is however the Carreck Archive that allows us to see these records through the ‘lens of the Geologists’ Association’. What distinguishes the Carreck Archive as a record of geotourism is the written record that accompanies the Archive. Successive archivists and contributors have ensured the photographs are accurately labelled, identifying geology, places and dates and, in many cases, the people seen in the photographs. Adding further depth there are many letters and postcards written to the archivists, particularly Miss M. S. Johnston, as well as contemporary newspaper cuttings providing local accounts of GA visits.

Modes of transport Documented within the GA Circular and Proceedings are many modes of transport. The growing

Fig. 4. Detail from Marjorie Carreck’s album of field meetings. Caption. ‘Digging out an elephant at Aylesford, Kent, September 1955. From left, standing Mrs. J. N. Carreck [Marjorie], Dr. M. P. Kerney, Mr. T. Harris, Mr. J. N. Carreck (crouching [Marjorie’s husband]), Mr. Goodfellow standing second from right, extreme right Mr. J. F. Wyley (H. N. Wright photo).’

THROUGH THE LENS OF THE GEOLOGISTS’ ASSOCIATION

123

Fig. 5. ‘Our rendezvous’, the meeting place for the excursion to Upware, T. W. Reader Album, 16 May 1914.

rail network provided trains and, more particularly, excursion trains run by many rail companies from the mid-nineteenth century onwards. These were widely used and on many occasions afforded direct observation of newly cut railway sections. Green (1989a) records references to carriages, brakes, trams, light engines and a ‘special electric car’ (used in an excursion to Essex, Youens & Priest 1908). At the turn of the nineteenth century there was a brief experiment with bicycles (Monckton 1899, 1903), although this did not prove as popular as hoped. The announcement in the GA Circular of the 1914 field meeting to Upware to the north of Cambridge is fairly typical: Excursion to Upware, Saturday May 16th, 1914 Director: T. McKenny Hughes F.R.S Excursion secretary: Miss E. Pearse, 3, Bessborough Mansions, Westminster, S.W. Leave Liverpool Street Station 11.05 a.m. Arrive Cambridge 12.21 a.m. Meet Miss Pearse in 1

main line booking office not later than 10.50 a.m., to obtain special tickets, price 6/2. At Cambridge take motor omnibus (2 miles) to the river near Victoria Bridge, where a steamer will be waiting to take the party to Upware. Omnibus fare 2d. each way. Steamer fare 2/return. Members should bring lunch. 1 There then follows a brief description of what will be seen on the voyage: the fens, archaeological sites linked to springs issuing from the chalk rock and river terraces. Then: The party will land at Upware opposite the public house known by the sign “Five miles from anywhere and no hurry” [Fig. 5] which will be the rendezvous for the day and where tea will be ready at 4 o’clock. (Plain tea 6d., with eggs 9d.)

The first record of a ‘motor’ vehicle is an excursion to Farnham Gravel Pits (in Surrey) in 1904 (Monckton 1904) when a motor was among carriages and cycles used to travel from Farnham Station to the extensive gravel pits (described as ‘a little to

Pre-decimalization (15 February 1971) British currency was divided into pounds (£), shillings (/-) and pennies (d.). One pound was worth 20 shillings and one shilling worth 12 pennies.

124

J. G. LARWOOD

the south east of Farnham’). Motor vehicles were increasingly used by individuals to rendezvous and as the main mode of transport for a wider itinerary, the earliest reported being to Surrey in 1914 (Young & Leighton 1915). This excursion was planned to afford members the opportunity to traverse a large tract of country (previously visited on different occasions) in one day (Fig. 6). A ‘char-a-banc’ was hired departing from Trafalgar Square in central London at 10 am and a detailed description of the route, variation in geology and topography and the stops taken is provided in the Proceedings, which concludes that ‘Notwithstanding that mist and rain at times spoilt the distant views, the whole excursion was pronounced a successful experiment’.

H. Dixon Hewitt: an individual geotourist Described in his obituary (J. E. W. R. 1967) as from southeast London and ‘towards the end of his life an

unmistakeable cockney’, Henry Dixon Hewitt (seen in Fig. 1a – number 19 in key) joined the GA in 1920. Among his interests was photography and he recorded a number of the Long Excursions of the Association. H. Dixon Hewitt was noted as one of the last old-fashioned field naturalists considered as a keen observer in many branches of science. He had a particularly keen eye ‘. . . missing nothing on a walk whether it was a chalk or greensand fossil, a fragment of Roman pottery, a new locality for a rare plant or a new species of snail in his garden’. He used ‘. . . an old-fashioned quarter-plate box camera with an excellent lens and as he always processed his own plates, his photographs were always of first quality’, carefully labelled in copper plate Indian ink and compiled in albums reflecting a personal view of his travels. He exhibited photographs at the annual meetings and donated many to the Association’s album. His albums of field meetings to Scotland, Ulster and the Boulonnais (Pruvost &

Fig. 6. ‘Motor Excursion to Surrey’ at Newlands Corner. T. W. Reader Album, 6 June 1914.

THROUGH THE LENS OF THE GEOLOGISTS’ ASSOCIATION

Pringle 1924; Fig. 7a, b), an album of photographs spanning 1920–1938 and the original glass plates were donated to the GA by Chichester Museum in 1982.

125

A geological ABC of North Cornwall The first recorded visit to Cornwall by the GA was during 8 –13 August 1887. Led by F. W. Rudler

Fig. 7. (a) Title and facing page of H. Dixon Hewitt’s Album from his trip to Boulonnais in 1923 (Pruvost & Pringle 1924). The party for this summer or Long Excursion (25 August–1 September) were based in the Grand Hotel de la Plage in Wimereux. H. Dixon Hewitt (and others) stayed on and the two photographs are dated post the organized excursion. (b) H. Dixon Hewitt album page of the Napoleon Carboniferous limestone quarries. Detail of workings at the He`naux Quarry with an additional photograph (below) by Mr W. L. Turner.

126

J. G. LARWOOD

(President of the GA), William Thomas (Secretary to the Mining Association and Institute of Cornwall) and A. K. Barnett of Penzance, the group visited the districts of Truro, Penzance and the Lizard (Rudler 1887). At the end of the visit it was concluded that ‘Ultimately they bent their way homewards, not simply laden with heavy loads of minerals and rocks, but carrying away the most pleasant recollections of their visit to the West country, while they left behind them a legacy of sincere gratitude for the cordiality with which they had been everywhere received by their Cornish friends’. The GA returned to the Lizard in Easter 1913 (Flett & Hill 1913) travelling from Paddington railway station on the Cornish Riviera Express. The party numbered 82 on several occasions, added to by local visitors joining throughout the tour. The Association then returned to North Cornwall in 1914 for their Easter field meeting (9–18 April). The visit is reported in the Proceedings (Dewey

1914; Hall 1914) but is brought to life by the compilation of a special album A “geological” A.B.C. of North Cornwall by W. A. McIntyre F.G.S and T. W. Reader F.G.S., Easter 1914 (Fig. 8a –c) which recounts their visit in rhyming couplets: A is St Austell where the China Clay is found it’s a much altered granite and lies white all around B is Miss Bauer whom all of us bless for her untiring efforts to make things a success C is the Carraway seed that we find in the silky phyllite of Woolgarden kind D is for Dewey describing a Spilite talking and teaching until it is twilight E is the Egg that was laid by a Pig it’s an orthoclase crystal sometimes very big F is for Frasnian the goniatite beds with layers of lavas in black, brown and reds

Fig. 8. (a) Title page of A “Geological” A.B.C. of North Cornwall by W. A. McIntyre F.G.S and T. W. Reader F.G.S., Easter 1914. (b) The Lantern China Clay works near St Austell with caption ‘A is St Austell where the China Clay is found, it’s a much altered granite and lies white all around’. (c) Elvan Quarry, Camelford with caption ‘Z is the Zeal the Geologists show, and that is what makes our excursions go.

THROUGH THE LENS OF THE GEOLOGISTS’ ASSOCIATION

G is George Young our President gay who holds us all under his most gracious sway H is for Hall who gases on gases and thinks all opponents a lot of he asses I is the Industry many display as up and down cliffs we hammer away J is for Jewell who lent us his quarry when he sees the result he will feel rather sorry K is the Kaolin pit beloved of Hall down goes the Director, the members and all L is for Lowe whose hammer’s the best when he’s once smote the rock there’s none left for the rest M is for Minverite of a rather dark tone it’s only a rechristened good old Greenstone N are the Nuts who stayed out till night leaping hedges and ditches after Luxulyanite O is the Order which all of us keep so the President’s whistle has fallen asleep

127

P stands for Padstow a sleepy old town where we waited an hour for a railway break down Q is the Quest for serrato striata which none of us found but were all of us after R stands for Reader the man with the camera indispensable quite but not as a hammerer S are the Students who joined us with Hall they ride motor bikes while the rest of us crawl T is the Time we got up in the morning it’s sometimes made earlier without any warning U is the ‘Ump which none of us get for why? Because the weather’s not wet V is Verneuili from Delabole high a Spirifer known as a black butterfly W is the Weather a dream of delight it’s only rained once and then in the night X is the Xercise which all of us take riding about in a motor brake

Fig. 9. (a) Cleaning and repair of album, reattaching loose leaf to be reinforced with Japanese hinging tape and (b) construction of new custom-made album portfolios where repair is not feasible. Individual pages are inserted into loose-leaf sleeves and bound together in a ring-bound portfolio. Albums (and all other material within the Archive) have been rehoused in bespoke grey drop spine boxes. (Photographs: Richard Weedon).

128

J. G. LARWOOD

Y is for Young A.C. I mean when he sits down beside you, you find he’s not lean Z is the Zeal the Geologists show and that is what makes our excursions go A note on the people mentioned illuminates their role on the excursion. Miss Grace M. Bauer was a member of the Association’s Council and assisted in the excursions. Henry Dewey was one of the two Directors of the excursion. George W. Young was President of the Association during 1914– 1915. T. C. F. Hall (Thomas Clifford Fitzwilliam Hall) was the second Director of the excursion. Mr and Mrs Jewell of Tregoodwell owned quarries at Tyland and Grey Lake. Lowe is unknown, though is possibly E. Lowe who directed excursions to the Leicester district in 1909 and 1912. T. W. Reader (Thomas William Reader) is photographer and one of the authors of the ABC. A. C. Young (Alfred Collett Young) was Field Meeting Secretary during 1905–1912 and 1918–1921. The humour of the ABC demonstrates the enjoyment of the field trips which have always been as much a social outing as a visit to see geology. There is reference to hammers and hammering, enthusiastic collecting, the weather, transport (and delays) and insights into the characters of the party. This reflects the inclusive, widely drawn and crosscutting GA membership (setting it apart from similar ‘scientific’ societies). In fact, in the 1920s this led the GA into a brief dispute with the Charity Commission over the object of field meetings, then known as ‘Field Excursions’. The Charity Commission argued that the field meetings were recreational, with ‘Excursions’ having an association with day trips and short holidays often with specially organized ‘excursion trains’. The resolution of this dispute is reflected in the fact that from 1930 onwards, what were previously referred to as Field Excursions became Field Meetings. Visits to Cornwall have continued (there are at least six recorded in the Proceedings since 1914) and there are now two GA Guides, one to the North Cornwall Coast (Dearman et al. 1970) and one to West Cornwall (Hall 2005).

The Carreck Archive: today and in the future Between 2008 and 2010, coinciding with the 150 years celebration of the GA, the Carreck Archive underwent a complete conservation assessment (Weedon 2008) and restoration funded by the GA’s Curry Fund. Photographs were cleaned and damaged albums repaired (Fig. 9a). Loose photographs were repackaged in transparent conservation envelopes and all the material (including albums) re-housed

in conservation-grade boxes, improving the overall order and storage of the collection (Fig. 9b). The T. W. Reader albums and others, which over the years had started to fall apart, were rebound in new portfolios (Fig. 9b). Having co-ordinated this restoration project, in 2010 the author became the fourth archivist and, with the handover of responsibility from Marjorie Carreck, recognizing her contribution, the archive was renamed the Geologists’ Association Carreck Archive. The Carreck Archive is now stored at the British Geological Survey (BGS, at Keyworth, Nottingham) which, through agreement with the GA, have taken responsibility for its long-term care. The Carreck Archive joins two other national geological photographic archives: the British Geological Survey’s own field photographs and the geological photographs of the British Association. Currently, the Carreck Archive is undergoing digitization with the plan to make more widely available much of the Archive through the BGS’s Geoscenic and GA websites. The record of field meetings within the GA has evolved and they are no longer reported in the Proceedings. The role of reporting has, in part, been taken on by the membership GA Magazine where, in particular, accounts of foreign trips appear. The Archive will continue to be exhibited at what is now the GA Festival of Geology. As the Carreck Archive is digitized and becomes more widely available, it is also hoped that interest will be reinvigorated and new contributions of photographs and now digital images encouraged, so that it will continue to document the activities of the GA and the future GA geotourist. The author thanks the GA’s Curry Fund grant support for the conservation of the Carreck Archive, securing its longterm future. Involved in the conservation and care of the archive are: R. Weedon (paper Conservator), who undertook the conservation work and continues to advise on the care of the Archive; J. Pendlebury (book binder) who re-bound albums, most notably the T. W. Reader volumes; and the British Geological Survey, who have entered into an agreement with the GA over the future care of the archive and are currently digitizing the Archive. M. Carreck should always be recognized (and the previous archivists) as it is through her 55 years of unstinting care that the archive exists today. Lastly, the time taken by two reviewers has been invaluable in honing this paper.

References Barrois, C. 1879. Excursion to Boulonnais. Monday, August 5th, 1878, and five following days. Proceedings of the Geologists’ Association, 16, 101– 132. Bensted, W. H. 1860. On the Kentish Ragstone as exhibited in the Iguanodon Quarry at Maidstone. Proceedings of the Geologists’ Association, 1, 57–60.

THROUGH THE LENS OF THE GEOLOGISTS’ ASSOCIATION Carreck, M. 2007. The Geologists’ Association’s photographic Archives. GA Magazine, 6, 11–12. Case, D. J. 2013. The Coast of the Bristol Region: Quaternary Geology and Geomorphology. GA Guide No 71, Geologists’ Association, London. Dearman, W. R., Freshney, E. C., King, A. F., Williams, M. & Mckeown, M. C. 1970. The North Coast of Cornwall from Bude to Tintagel. GA Guide No 10, Benham & Company Ltd, Colchester. Dewey, H. 1914. Report of an excursion to North Cornwall, April 9th to 18th (Easter), 1914, part 1. Proceedings of the Geologists’ Association, 25, 154–179. Flett, J. S. & Hill, J. B. 1913. Report of an excursion to the Lizard, Cornwall: April 16th to 18th, Easter, 1914. Proceedings of the Geologists’ Association, 24, 313–327. Green, C. P. 1989a. Excursions in the past: a review of the Field Meeting Reports in the first one hundred volumes of the Proceedings. Proceedings of the Geologists’ Association, 100, 17–29. Green, C. P. 1989b. The illustration of the Proceedings: the first one hundred volumes. Proceedings of the Geologists’ Association, 100, 31–54. Greensmith, T. 2008. History of the Geologists’ Association Field Guides (1958– 2008). GA Magazine, 7, 6 –9. Hall, T. C. F. 1914. Report of an excursion to the St. Austell District, Cornwall, April 16th to 18th, Easter, 1914. Proceedings of the Geologists’ Association, 26, 34– 46. Hall, A. 2005. West Cornwall, Geologists’ Association Guide No 31, City Print Ltd, Milton Keynes. Hardie, W. G., Bennison, G. M., Garrett, P. A., Lawson, J. D. & Shotton, F. W. 1958. The Area Around Birmingham. GA Guide No 1, Benham & Company Ltd, Colchester. Harrison, W. J. 1882. Geology of the Counties of England and Wales. Kelly and Co., London. Hill, W. 1914. Report on an excursion to Knebworth and Hertford: May 2nd, 1914. Proceedings of the Geologists’ Association, 25, 288– 291. Holmes, T. V. & Sherborn, S. D. (eds) 1891. A Record of Excursions made between 1860 and 1890. Geologists’ Association, Edward Stanford, London. Hose, T. 2000. European Geotourism – Geological interpretation and geoconservation promotion for tourists. In: Barretino, D., Wimbledon, W. P. & Gallego, E. (eds) Geological Heritage: Its Conservation and Management. Instituto Tecnologico Geominero de Espana, Madrid, 127–146. Hose, T. A. 2008. Towards a history of geotourism: definitions, antecedents and the future. In: Bureck, C. W. & Prosser, C. D. (eds) The History of Geoconservation. Geological Society, London, Special Publications, 300, 37–60. J. E. W. R. 1967. Obituary notice [Henry Dixon Hewitt]. Proceedings of the Geologists’ Association, 78, 371–372. Jones, T. R. 1880. The Geologists’ Association; its origin and progress. Proceedings of the Geologists’ Association, 7, 1 –57. Leach, A. L. 1934. The Foulerton Award: recipients 1920–1932. Proceedings of the Geologists’ Association, 45, 72– 77.

129

Leach, A. L. 1956. Mary Sophia Johnston (Obituary). Proceedings of the Geologists’ Association, 67, 197– 199. Leake, B. E., Bishop, A. C. & Howarth, R. J. 2013. The Wyley History of the Geologists’ Association in the 50 years 1958– 2008. Geologists’ Association, City Print Ltd, Milton Keynes. Monckton, H. W. 1899. Cycling excursion from Winchfield to Wokingham. Proceedings of the Geologists’ Association, 16, 153– 155. Monckton, H. W. 1903. Cycling excursion to the Aldershot District. Proceedings of the Geologists’ Association, 18, 184– 188. Monckton, H. 1904. Excursions to the Farnham Gravel Pits on April 23rd, and to the brickfields and gravel pits at Dawley, between Hayes and West Drayton on April 30th, 1904. Proceedings of the Geologists’ Association, 18, 409– 414. Monckton, H. W. & Herries, R. S. (eds) 1910. Geology in the Field – the Jubilee Volume of the Geologists’ Association (1858–1908), 2 parts. Stanford, London. Pruvost, P. & Pringle, J. 1924. Excursion to the Boulonnais: August 25th to September 1st, 1924. Proceedings of the Geologists’ Association, 35, 56– 67. Ramsay, A. C. 1874. The Physical Geology and Geography of Great Britain a Manual of British Geology. 4th edn, Stanford, London. Robinson, E. 1989a. The origin and early years of the Proceedings: context and content. Proceedings of the Geologists’ Association, 100, 5– 15. Robinson, E. 1989b. ‘Clarion o’er the dreaming earth’: a personal view of the GA Circular since 1885. Proceedings of the Geologists’ Association, 101, 101– 117. Rudler, F. W. 1887. Excursion to Cornwall: Monday, August 8th to Saturday, August 13th 1887. Proceedings of the Geologists’ Association, 10, 196– 216. Sweeting, G. S. 1958. The Geologists’ Association, 1858– 1958. Benham & Company Ltd, Colchester. Watts, W. W. 1909. The Jubilee of the Geologists’ Association. Proceedings of the Geologists’ Association, 21, 119 –49. Watts, W. W. 1925. Excursion to South Shropshire: July 23rd to 30th, 1925. Proceedings of the Geologists’ Association, 36, 394– 405. Weedon, R. 2008. A conservation Survey and Preservation Plan for the Geologists’ Association Archive. Unpublished report available from the Geologists’ Association. Wiltshire, T. 1859. Field meetings of the Geologists’ Association. The Geologist, 2, 369. Wiltshire, T. 1860. Ordinary meeting, 5th March 1860. Proceedings of the Geologists’ Association, 1, 38–48. Woodward, H. B. 1887. The Geology of England and Wales. George Phillip and Son, London. Youens, E. C. & Priest, S. 1908. Excursion to Dartford and Stone. Proceedings of the Geologists’ Association, 20, 127 –128. Young, G. W. & Leighton, D. 1915. Report of a motor excursion in Surrey. Proceedings of the Geologists’ Association, 26, 118– 20.

The contribution of maps to appreciating physical landscape: examples from Derbyshire’s Peak District C. JOHN HENRY1* & THOMAS A. HOSE2 1

Nineteenth Century Geological Maps, 71a Oxford Gardens, London W10 5UJ, UK 2

School of Earth Sciences, University of Bristol, Wills Building, Queens Road, Bristol BS8 1RJ, UK *Corresponding author (e-mail: [email protected])

Abstract: It is only in the last 100 years or so that most of Britain has been covered by accurate, published, topographic and geological maps. Although travellers’ guides were available from the late seventeenth century, they lacked adequate maps. Whilst fairly accurate maps of the major roads were published in the early seventeenth century as strip maps, topographic maps were not generally available until the nineteenth century. Cartographers, usually when preparing county maps, struggled with the representation of Britain’s varied topography. In the nineteenth century, medium-scale (1-inch-to-the-mile) topographic maps initially developed through the agency of the prizes offered by the Royal Society of Arts but primarily due to the Ordnance Survey. Geological maps benefitted from improved base maps – those of John Cary and the Ordnance Survey. This paper especially explores and illustrates the development of maps and the role they played in the depiction and understanding of landscape and promotion of the major early geotourism region of the Peak District from 1780 to 1930.

Given the wealth of material available for study, any necessarily condensed account of the mapping, depiction and description of Derbyshire’s Peak District landscape for travellers and geologists (or geotourists) will inevitably suffer from the vagaries of informed selection. As one of the earliest modern guidebooks for the region noted ‘Very remarkable is the county of Derby for its great variety of surface . . . we first behold a champaign (sic) country, then gentle eminences which, by gradual transition are succeeded by hills that increase, as we advance, in height and extent and terminate at length in that mountainous tract called the Low Peak, or Wapontake, and the High Peak’ (Ward 1827, p. 5). The region early attracted the attention of travellers, artists and geotourists (Hose 2008, pp. 43–45). The publications and maps developed since the mid-nineteenth century for their guidance usefully reflect contemporary national developments; the region’s varied landscape proved a cartographer’s challenge from the outset. During the period of this account, 1780–1930, it was widely considered, as a Castleton guidebook noted, that ‘Mineralogy forms the study equally of the Gentleman and the Artist’ (Hedinger 1839, preface). Mapping at the same time was going through major changes that merged the artistic, aesthetic and technological approaches to topographic depictions and descriptions, as seen in contemporary maps and guidebooks. We so much accept today’s ready availability, comparative cheapness and accuracy of topographic

maps (with coordinate grid systems and generally fixed place-name spellings) that it is hard to conceive of a time when no such tool for landscape appreciation was available. In the 1800s it was difficult for the traveller to fully appreciate the nature of the physical landscape. By the 1900s the aesthetic appreciation of landscape was much easier. Initially the realm of a small leisured class, it then extended to a much wider circle. Over the course of the nineteenth century, the advent of railways, the shorter working week, annual and bank holidays, rising incomes and increasing literacy levels led to increased personal mobility and interest in places beyond the immediate locale. This generated demand for more useful maps (encouraging design improvements and greater availability) and stimulated guidebook developments. For travellers the most important aspect of landscape depiction was navigation. Uppermost in the eighteenth-century appreciation of landscape by established landowners and industrialists were the possibilities for agricultural improvement and mineral extraction. The available maps were inadequate, because the systematic appraisal of terrain for such purposes was in its infancy. By 1800 the coverage of England by 1-inch-to-the mile (1:63 360) county maps was fairly comprehensive; by the 1820s half-inch (1:126 720) reductions of these were also readily available. Cary’s New Map of England and Wales with Part of Scotland of 1794 had continuous mapping, at a scale of 5 miles to the inch (1:316 800); this was the first

From: Hose, T. A. (ed.) 2016. Appreciating Physical Landscapes: Three Hundred Years of Geotourism. Geological Society, London, Special Publications, 417, 131–155. First published online September 7, 2015, http://dx.doi.org/10.1144/SP417.13 # 2016 The Author(s). Published by The Geological Society of London. All rights reserved. For permissions: http://www.geolsoc.org.uk/permissions. Publishing disclaimer: www.geolsoc.org.uk/pub_ethics

132

C. J. HENRY & T. A. HOSE

departure from the county format that, modified, became the basis for William Smith’s iconic 1815 geological map, A Delineation of the Strata of England and Wales with Part of Scotland. Cary returned to the county format at a larger scale in 1809 with his New County Atlas; this had county maps at various scales, typically 3 miles to the inch (1:190 080). He adapted the 1818 second edition of this atlas for William Smith’s County Geological Atlas, which was issued in parts between 1819 and 1824 but not completed. By the close of the nineteenth century, Ordnance Survey (OS) maps on a scale of 1 inch to 1 mile (1:63 360), colloquially ‘one-inch’ maps, were the most accurate available. They provided the base maps for the Geological Survey’s maps. By the early nineteenth century fairly accurate small-scale road maps were included within many tourist guidebooks. Whilst this account is focussed on nineteenthcentury maps for the Peak District, a broader temporal and geographical framework is provided to contextualize their development. The Peak District was the first region in England promoted to travellers for its natural wonders (Hose 2008, p. 43), which are influenced by the underlying geology. The natural wonders were significantly recognized by the ‘picturesque’ and later ‘romantic’ movements. Naturally, visits to these landscapes needed, apart from transport and accommodation, the services of either competent guides or the combination of reliable maps and guidebooks. The earliest regional maps for England and Wales, dating from the late sixteenth century, invariably covered a county and were essentially decorative (see Table 1). Topographically imprecise and generally without roads, they showed places and features considered important at the time of their printing; they emphasized cultural and historical prejudices and likely traveller destinations. The first bound volume of English county maps was published in 1579 by Christopher Saxton (c. 1542– c. 1610), the ‘father of English cartography’; his Atlas of the Counties of England and Wales was the first such regional coverage of any European country. His map of Derbyshire is noteworthy for depicting its major rivers, woodlands, estates, moorlands and dales with different symbols, but no roads – even between the major settlements. Later maps by John Speed (1552– 1629) improved on topography slightly but changed little else (see Fig. 1).

Mapping prior to the nineteenth century: the county map Up to the late eighteenth century, only strip (see Table 2) and county maps were available. County maps up to those of Emanuel Bowen

(1694?–1767) were unhelpful for travellers’ navigation, because roads were generally omitted; these maps depicted, at relatively small scales, drainage and the boundaries of the administrative areas known as ‘hundreds’ or ‘wapontakes’ into which English counties were subdivided. Topographic relief was depicted by stylized ‘plum pudding’ hills that, by their relative size and frequency or absence, gave a crude impression of the terrain. The maps often focussed on the coats of arms of the great county families and their estates, which were shown as simple rounded enclosures. Towns and villages were approximately located dots, and roads were generally omitted. These early county maps, which were often bound into atlases, provided limited substance for landscape appreciation and, away from rivers, were of no navigational use. This last significant failing was addressed by route or strip maps pioneered by John Ogilby (1600–1676) for the post-roads of England and Wales. His maps, accurate and detailed, were depicted as an unwinding ribbon on which the stagecoach route corridor was centred; stately homes visible from the road and towns, villages, and staging posts were depicted pictorially. Map orientation was indicated by a compass rose and varied along each strip in order to fit the maximum distance on the sheet; measured distances were given. His maps were published in Brittania in 1675. This less than portable atlas of 300 pages, 13.5 in × 18 in (34 cm × 46 cm), at a scale of 1 inch to the mile (1:63 360) was the first specifically to portray roads from London to and across the counties of England and Wales. Figure 2 illustrates the route from Manchester to Derby through the Peak District. Ogilby, and later Bowen, greatly assisted travellers’ navigation on their selected linear routes, showing the main sights nearby – rather like the maps in the present-day multi-volume series on Britain’s waterways published by Collins/Nicholson since the 1970s. Hills with a steep gradient were depicted across the strip, but otherwise there was negligible terrain guidance. County and strip maps were commercially published by surveyors and cartographers. They improved only gradually with much, often unchecked, plagiarism and a relative absence of rigour until John Cary. From the late seventeenth century, topographic maps were small scale, typically between 5 (1:316 800) and 12 (1:760 320) miles to the inch. Surveyors employed relatively crude compass bearings and distance estimation to triangulate from church towers and hill-tops. The maps produced sufficed as sketch maps within guidebooks but were inadequate for mediumand large-scale mapping. John Cary’s first atlas, of 1787 was produced with distances measured by a wayfarer’s wheel and with frequent compass bearings. His subsequent atlases of 1794 and 1809

ROLE OF MAPS IN PEAK DISTRICT GEOTOURISM

133

Table 1. Prominent county map-makers up to the nineteenth century Years

Cartographer/publisher

1579

Christopher Saxton (c. 1542– c. 1610)

Atlas of the Counties of England and Wales Decorative: no roads, not triangulated, generic profile hills; emphasis on drainage, boundaries and large estates

1610

John Speed (1552 – 1629)

The Theatre of the Empire of Great Britaine, published in 1610 – 11 Decorative: no roads, not triangulated, generic profile hills increasing in size to the northwest and massed more realistically than Saxton; emphasis on drainage, boundaries and large estates (see Fig. 1)

1695

Robert Morden (c. 1650– 1703)

Maps in Camden’s Britannia in 1695 and later editions of 1722, 1753 and 1772 Decorative: no roads, not triangulated, generic profile hills increasing in size and steepness to the northwest and massed more realistically than Speed; emphasis on drainage, boundaries and large estates

1720

Emanuel Bowen (c. 1694– 1767) and John Owen (1720 – ?)

Britannia Depicta or Ogilby Impov’d; Being a Correct Coppy of Mr. Ogilby’s Actual Survey of all y Direct & Principal Cross Roads in England and Wales: Wherein are exactly Delineated & Engraven, All y cities, Towns, Villages, Churches, Seats & scituate on or near the Roads with their respective Distances in Measured and Computed Miles. And to render this work universally usefull & agreeable, [beyond any of its kind] are added in a clear & most compendious method. A full & particular description & account of all the cities, borough-towns, towns-corporate & their arms, antiquity, charters, privileges, trade, rarities, & with suitable remarks on all places of note drawn from the best historians and antiquaries County maps in miniature with more than 200 strip maps printed both sides for compactness Decorative map: straight main roads, generic profiled hills of uniform size in northwest, but otherwise omitted; distances between main towns in table above map; road detail on strip maps; Derby to Buxton in strips on reverse side of map

1752

Thomas Kitchen (1718 – 1784)

Henry Boswell’s Antiquities of England & Wales Decorative: main roads, generic profiled hills of uniform size evenly distributed; compiled, not surveyed

1767 1767

Comments

In John Bowles’ Post-Chaise Companion through England and Wales Not surveyed; main roads straight and approximate; hill icons Peter Perez Burdett (1734 – 1793)

First 1-inch-to-the-mile map of Derbyshire and second successful claimant of the 1759 Royal Society of Art prize

Table 1 is based on Handford (1971).

show evidence of triangulation between major centres (Wigley 2015, www.strata-smith.com/ ?page_id=433) by the relative accuracy of their positioning, although coastal and remote positions are not accurate. His maps accurately depicted a well-developed road network with hachures fairly convincingly showing topographic relief before rigorous national mapping at a larger scale was available.

Topographic depiction and surveying accuracy The illustration of relief was restricted to icons of hills, as ‘plum puddings’ (see Fig. 1) from Saxton

until Cary and, in Derbyshire, P. P. Burdett (1734– 1793). Variations in hill size and topographic ruggedness were indicated by the size, shape and spacing of the ‘plum puddings’. Hachures were a considerable improvement in the depiction of hilly topography. Hachures (see Fig. 3) are short strokes of the pen or the engraver’s stylus in the direction of slope with the slightly wider head of the stroke at the top of slope tailing off towards the base of the slope; they were a considerable improvement in the depiction of hilly topography. Drawn crudely or used symbolically, hachured ridges resembled, as they were known colloquially, hairy caterpillars. However, in skilled hands, the spacing and splaying of the hachure strokes sensitively illustrated

134

C. J. HENRY & T. A. HOSE

Fig. 1. Darbieshire described, John Speed, 1610. Crests of aristocratic families are featured; their parks are shown as circles enclosing trees. Buxton town plan and St. Anne’s Well – already tourist attractions – are in the lower corners. The ‘plum pudding’ representations of hills increase in size to the north and west to simulate the actual topographic variation within the county. Roads are absent. Maps were published as monochrome engravings and then taken to a colourist if desired.

the actual hill form and gave a very good visual impression of the topography (Andrews 2009, pp. 238– 239); the engravers relied on surveyors’ field notes and sketches. Often the fair drawn copy of the map sent to the engraver was a water-colour rendering of the hill form that used the direction of brush stroke and density of colour to indicate topography. Hachures were very effective in depicting terrain, although they did not show heights, but in hilly and mountainous areas they darkened the map such that other information – place names, symbols, routes and drainage – could be obscured or did not stand out clearly. When geologists applied water-colour to hachured base maps, there were often further problems of legibility; dense hachuring could so darken geological colouring as to be misleading in hilly and mountainous areas. However, the very success of their impressions of hill form delayed the introduction of more quantitative techniques such as contours and altitudinal

colour shading. Contours are lines of equal elevation along which every point is at the same height above a datum, usually sea-level. With a suitable interval, contours can give an impression of topography; however, the impression is not so immediately visual as good hachuring and takes more ‘reading’ of the map. Contours enabled accurate profiles to be drawn in any direction to give heights and slope angles, which were of practical use to engineers, geologists and travellers venturing onto the hills. The addition of colour layering to contoured maps, as seen on Bartholomew’s half-inch-to-themile(1:126 720)maps,colloquially‘half-inch’maps, introduced another type of topographic impressionism and one that is readily ‘readable’. Typically, shades of green depicted low-lying areas and browns represented higher ground. Mauves and purple or white were often used for high peaks and ridges. The colours approximated those seen in the Alpine world at such altitudes from a distance.

ROLE OF MAPS IN PEAK DISTRICT GEOTOURISM

135

Table 2. Early strip maps Years

Cartographer/publisher

Comments

1533 – 1543

John Leland (1503– 1552)*

Private notes on travels as Thomas Cromwell’s agent prior to the Dissolution of the Monasteries and subsequently informed Camden’s Britannia – see Robert Morden in Table 1. It was presented to Henry VIII* but not published until the eighteenth century (Toulmin Smith 1910).

1675

John Ogilby (1600– 1675)

Britannia, Volume the First: or An Illustration of the Kingdom of England and the Dominion of Wales: by a Geographical and Historical Description of the Principal Roads thereof. Actually admeasured and Delineated in a Century of Whole-Sheet Copper-Sculps. Accomodated with the Ichonography of the Several Cities and Capital Towns; and Completed by an Accurate Account of the More Remarkable of Passages of Antiquity, Together with a Novel Discourse of the Present State. Ribbon-style strip maps in Britannia are at a scale of 1 inch to the mile (1:63 360). It includes Derby to Manchester via Buxton.

1720

Emanuel Bowen (c. 1694– 1767) and John Owen (1720– ?)

See Table 1.

*‘I have so travelid yn yowr dominions booth by the se costes and the midle partes, sparing nother labor nor costes, by the space of these vi. yeres paste, that there is almoste nother cape, nor bay, haven, creke or peere, river or confluence of rivers, breches, waschis, lakes, meres, fenny waters, montaynes, valleis, mores, hethes, forestes, wooddes, cities, burges, castelles, principale manor placis, monasteries, and colleges, but I have seene them; and notid yn so doing a hole worlde of thinges very memorable’ (Toulmin Smith 1910, xli).

Advanced cartography, as in Switzerland, applied hill shading, a more delicate shadow effect than hachures, to colour layering and contours to produce maps of great beauty and utility. This degree of sophisticated cartography was not really demanded by Britain’s non-Alpine terrain and did not appear until the mid-twentieth century on selected OS tourist maps, such as for that for the Lake District, which was first published in 1948 and in various versions up to 2001.

Late eighteenth-century mapping developments In the mid-eighteenth century, detailed and accurate maps were needed for then-developing administrative and legal purposes; the planning and construction of turnpike roads and then canals, estate management and the enclosure of formerly common land. Measured baselines with precise bearings to establish absolute positions were not widely employed until the late eighteenth century. Improvements in equipment and more accurate angle measurements with theodolites developed in the late eighteenth century to meet surveyors’ needs. Their work began in the 1720s and increased dramatically from the 1770s (Hobley 1964, p. 18); more than 5 million acres (2 023 428 hectares) were enclosed with new hedges and service

tracks that had to be accurately set out in accordance with the Parliamentary bills for which maps had to be presented. Large-scale enclosure and tithe maps were prepared from around 1750 to 1850. The improved technology also aided levelling, that is, the measurement of relative heights. Initially, isolated hills were measured trigonometrically relative to their surroundings; later, barometric altimeters relying on differences in air pressure to give relative heights above sea-level were employed. The benefit of the detailed mapping of Scotland’s terrain to the English army in its operations in the 1740s was promoted by Major-General William Roy (1726– 1790). His maps were fundamentally sketch maps lacking trigonometric rigour but realistically depicting the terrain. Roy’s work and advocacy led eventually to the establishment of the OS in 1791. Travellers were a comparatively small market for such detailed maps prior to 1791, because relatively few people journeyed much beyond the immediate neighbourhood of their birth or work. The first depiction on a map of a Derbyshire geotourist venue is that of Buxton’s St. Anne’s Well on Speed’s 1610 Derbyshire map (see Fig. 1). Simpson’s 1746 Derbyshire map is especially noteworthy for its depiction of Peak Cavern and Poole’s Hole. Moule’s 1837 Derbyshire map also included an image of Peak Cavern (see Fig. 3).

136

C. J. HENRY & T. A. HOSE

ROLE OF MAPS IN PEAK DISTRICT GEOTOURISM

137

Fig. 3. Derbyshire’s geotourist sites as depicted on county maps and guidebooks; there has always been some traveller and tourist interest in the county’s spas, caves and mines expressed in many maps and guidebooks: (a) Simpson’s 1746 county map: probably the earliest published representations of Poole’s Cavern, Buxton and Peak Cavern (or the Devil’s Arse), Castleton; (b) Moule’s 1836 county map: Peak Cavern, Castleton (original is coloured); (c) Hedinger’s 1839 A Short Description of Castleton, in Derbyshire: Peak Cavern, Castleton; frontispiece to the book, which also includes an illustration of Peak Cavern identical to (b) above; (d) Knight’s 1871 Journey-Book of England: Derbyshire: Odin’s Mine, Castleton – the oldest documented lead mine in Derbyshire. Fig. 2. John Ogilby 1675. The Continuation of the Road from York to West Chester, p. 90 of Britannia. Two unconnected routes are on this map. From left to right, the top of each strip connects with the bottom of the next. The miles count from bottom left to top right. The first route begins at Warrington in the lower left corner and ends in West Chester on the River Dee midway up the second strip at a line across the strip. The second route begins at Manchester. At Stockport, near the bottom of the third strip, Manchester Hill is like a butterfly across the route, with the ascent from base to peak and the descent on the ‘upside-down’ hill from peak to base. Other hills are drawn in the direction of gradient. Buxton is at the bottom of the fifth strip. The route ends at Derby, mile 55, in the top right corner. On each compass a crown points north and a cross points east.

138

C. J. HENRY & T. A. HOSE

In 1759 the Royal Society of Arts1 (RSA) announced its intention to award a £100 prize to surveyors for ‘an accurate Survey of any County upon the Scale of one Inch to one Mile’ (Harbeson et al. 1963, p. 45), envisioning that the resulting maps might fit together to produce a national map. Such mapping was beyond the resources of most surveyors given the 1-year time limit, and the prize was far from adequate to cover the necessary expenses. The competition was unsuccessful in its larger intention, awarding only £460 plus medals for 13 county maps by 1802. However, it did stimulate surveyors like Burdett in Derbyshire (see Fig. 3). The absence of a national triangulation network based on rigorous surveying ultimately defeated the optimistic vision of the RSA to compile a national map at a scale of 1 inch to the mile (1:63 360). To produce such a network was beyond the financial means of even the largest commercial surveyors and government resources were required.2 The Trigonometrical Survey of the Board of Ordnance was established in 1791; the threat of Napoleonic invasion was partly the catalyst (Hewitt 2010, p. 112) for creating a national survey framework for which Roy, the RSA and others had been pressing. It was soon renamed the Ordnance Survey. Plagiarism was a major problem for anyone desiring to innovate and improve maps, such as those based on accurate triangulation, because it reduced potential profits from the sales of newly surveyed maps. In 1766, the Engravers’ Copyright Act made more effectual the 1735 Engravers’ Act in extending and reinforcing copyright protection to cartographers and surveyors.3 Possibly not coincidentally, the last of the great English county mapmakers, John Cary, published his first engraved map in 1779, and his first county atlas, Cary’s New and Correct English Atlas in 1787 (Fordham [1925] 1976, p. 23). It had typical small-scale county maps (approximately 11 miles to the inch) but was noteworthy for their clean uncluttered appearance. In it, he advertised as ‘now engraving’ his New Map of England and Wales and part of Scotland, first published in 1794 at 5 miles to the inch (1:316 800). This departed significantly from available atlases, because it provided continuous mapping across county boundaries at a constant 1

scale (Fordham [1925] 1976, p. 45). It was not laid out county by county with varying scales to accommodate each to a page; Cary resolved the edge-matching defects along county boundaries and provided a more national perspective. Cary completed the Cary’s New and Correct English Atlas in 1809. In this large atlas – page size 21.5– 22.4 in (54–56 cm) high by 35.2 – 37.6 in (88 –94 cm) wide – he returned to the popular county atlas format, but the map scales were large, varying from 1 to 4 miles to the inch (Fordham 1925, p. 82). Although the maps were presented as counties, he numbered the turnpikes at county borders so that they linked to the neighbouring county sheets. It went through several editions issued as revised whole volumes and as individual sheets. John Cary’s sons, John and George, later enlarged their father’s 1794 national atlas in Cary’s Improved Map of England and Wales with a Considerable Part of Scotland Planned on a Scale of Two Statute Miles to One Inch in 1832 (Fordham 1925, p. 120). It combined the counties as continuous mapping on 65 plates each 22.5 × 27.5 in (58 × 70 cm) and was later published as a boxed set of folding maps. In southern England, this late Cary atlas benefitted from the early triangulation work done by the OS. In the north, the Carys had to rely on their own and others’ piecemeal triangulation. Cary’s sons sold the map plates and copyright to G. F. Crutchley in 1844; he continued to publish the maps under his own imprimatur. In turn, he sold them in 1877 to Messrs Gall & Inglis, who published them under their own names into the 1890s. Perhaps the continuation of the Carys’ maps under other guises was the sincerest form of flattery, unacknowledged but evidently profitable. The RSA’s approach fostered the continuing use of the county as the base mapping unit for regional topographic mapping. It also led to the accurate re-surveying of numerous counties. Between 1759 and 1808 a total of 13 maps were awarded the prize, the first in 1765 being for Devon. Burdett’s 1767 Survey of Derbyshire (see Fig. 4) is typical of the accuracy of mapping then achievable and of the limitations of contemporary printing technology; roads, in particular, were accurately portrayed, together with the new industrial features appearing

The Society for the Encouragement of Arts, Manufactures and Commerce, popularly referred to at the Society of Arts, received its royal charter in 1847 to become the Royal Society of Arts. 2 By comparison in France the national map, produced by four generations of the Cassini family between 1756 and 1815, at 1:86 400 scale in 182 sheets was a formidable achievement. Based on a rigorous national geodetic triangulation, the accuracy of its depiction of France’s road network is such that it registers well with current satellite imagery. See David Rumsey Historical Map Collection: |whttp://rumsey.geogarage.com/maps/cassinige.html. 3 An act to amend and render more effectual an act made in the eighth year of the reign of King George the Second, for encouragement of the arts of designing, engraving, and etching; historical and other prints; and for vesting in, and securing to, Jane Hogarth widow, the property in certain prints, 1766, 7 Geo. III, c.38.

ROLE OF MAPS IN PEAK DISTRICT GEOTOURISM

in the countryside such as water-mills and factories. The map covered three full and three half-sheets and was produced from original surveys. It was the most accurate of any county map produced up to that time (Harley 1975, preface). Burdett’s plates passed through several hands after his death, undergoing partial revisions; his map was re-issued in a 1791 edition published by Snowden and sold by John Cary. The first published 1-inch OS map in 1801 was of Kent in four sheets with adjacent counties left blank; subsequently the OS dropped the county as the basic unit of regional mapping. The initial layout and numbering reflected the ad hoc military context: sheet number 1 was London, with the numbering initially following the south coast; the map sheets were large, generally covering 35 × 23 miles (56 × 37 km). After the Napoleonic wars, the unwieldy large folio sheets already completed were retained, but subsequent maps were issued as quarter-sheets covering approximately 18 × 12 miles (29 × 19.3 km). The ungainly layout of variable sheet sizes and bizarre numbering continued until the New Series was adopted in 1885 by the OS and 1900 by the Geological Survey; the latter used OS copper plates on which it then engraved geological information; the printed maps were hand coloured to show the geological formations. The OS 1-inch maps were especially good at showing roads, canals and railways. They provided a broad accurate overview of topography and settlements. Their move from hachures to contours in the 1860s and 1870s was significant in promoting accurate land form studies.

Nineteenth-century geological mapping Accurate baselines for national survey were established by the OS in the early nineteenth century; they were used for triangulation to locate major features accurately. By the 1860s, OS heights were measured for contouring and plotted relative to a national datum. In 1885 a standardized sheet format and renumbered layout were introduced as the New Series. Towards the end of the nineteenth century improvements in printing technology enabled improved and revised maps to be reproduced cheaply and in quantities to meet the travelling public’s demands. 4

139

One of the earliest instruction manuals for field geology made the point that it is of the very first importance that the geologist should, before he proceeds to the examination of a country, be provided with the best physical map that can be procured, so that his observations may be recorded on that which will not deceive him. It must be admitted that such maps are sufficiently rare . . . too much praise cannot be given to the late sheets of those published by the Ordnance, remarkable not only for their general fidelity, but also for the shading of the hills. . . . With these maps in his hands the geologist feels that his time is not thrown away, and by noticing various minute circumstances upon them, he is subsequently enabled to soar, as it were, above the country he has examined; and by combining his various observations, he may arrive at general conclusions, with which he might not otherwise feel satisfied, and to which he might never have been led without an exact document of this nature. (De la Beche 1832, pp. 544 –5)

The OS did not begin its work in Derbyshire until the 1850s. Before this time maps used by travellers and tourists in Derbyshire, whether in separate sheets, atlases, or incorporated within guidebooks, were commercial offerings. It is therefore unsurprising that William Smith (1769– 1839) employed Cary’s 1796 ‘one-sheet England’, compiled from the assembled pages of his 1794 atlas, a New Map of England and Wales and Part of Scotland, as a base map for his A Delineation of the Strata of England and Wales and Part of Scotland. Cary and Smith clearly conferred to produce a most effective base map for showing geology; Cary re-engraved his 1796 version, de-cluttering it by removing administrative boundaries and many place names. At Smith’s request he developed more detail in the drainage system (see Fig. 5). After Smith had published his (financially unsuccessful) geological map in 1815, he began work on two other related projects. These were geological cross-sections and atlases of county geological maps. Both were expensive major undertakings that stretched his resources and were never completed. His county maps were based upon those of Cary’s New and Correct English Atlas 1809. These maps with Smith’s geological information, were published in six ‘parts’ of four maps each, 21 counties with Yorkshire requiring four sheets, between 1819 and 1824.4

Part I (January 1819) – Norfolk, Kent, Wiltshire, and Sussex; Part II (September 1819) – Gloucestershire, Berkshire, Surrey and Suffolk; Part III (February 1820) – Oxfordshire, Buckinghamshire, Bedfordshire and Essex; Part IV (1821) – Yorkshire (in four sheets); Part V (1822) – Nottinghamshire, Leicestershire, Huntingdonshire and Rutland; Part VI (1824) – Cumberland, Durham, Northumberland and Westmoreland (Davies 1952, p. 388). As there was insufficient room on the county atlas pages for a conventional geological legend, Smith arranged the colours in ‘tablets’, together with the formation descriptions, in the blank spaces around the mapped area and opposite the geological formations to which they referred.

140

C. J. HENRY & T. A. HOSE

ROLE OF MAPS IN PEAK DISTRICT GEOTOURISM

141

Fig. 5. (a) Map extract of the Peak District from Cary’s New Atlas of England and Wales and part of Scotland, 1794, p. 41, which marked a change from traditional county atlases in terms of accuracy and clarity. It used type size and font to signify the size and importance of towns and showed minor as well as main roads. It was the first modern atlas. (b) An extract from William Smith’s A Delineation of the Strata of England and Wales with part of Scotland, 1815. To prepare a less-cluttered base map to this first geological map of a nation, Cary deleted boundaries, minor roads and many place names and reduced hachuring. He introduced topographic names and increased drainage detail to Smith’s specification. (Courtesy of Daniel Crouch; http://www.crouchrarebooks.com)

Cary partly completed three additional county geological maps before the county geological atlas aborted. These maps appear in some later editions of Cary’s conventional atlas, coloured to show administrative divisions but with geological boundaries and/or legend tablets visible as line work and

lettering but effectively ignored.5 Derbyshire was unfortunately not amongst the published or the ‘recycled’ geological sheets. In addition to Smith’s national map and prior to the availability of Geological Survey maps in the mid-1850s, an increasing number of national

Fig. 4. Extract of P. P. Burdett’s The Survey of Derbyshire, second edition, 1791. (a) First published in 1767, it won the prize offered by the RSA to encourage accurate topographic mapping at 1 inch to the mile. The RSA is acknowledged in the title block (b). The map extract is reproduced at approximately the original scale. Hachuring gives a reasonable impression of the topography but in places conflicts with the lettering. Details of roads, individual buildings and minor drainage could be represented at this scale.

5

William Smith was working on number of county maps that survived ‘in a state of forwardness’ in his papers in various stages of completion (Phillips 1844, p. 149). Manuscript copies of Northamptonshire, Staffordshire, Monmouthshire, Lancashire and Cambridgeshire are preserved in the Smith Collection of the Oxford University Museum of Natural History. In Cary’s New and Correct English Atlas editions of 1824, 1828 and 1834, the maps for Somerset, Lincolnshire and Northamptonshire appear with some of Smith’s work engraved, but they are coloured as administrative units (Davies 1952, p. 390).

142

C. J. HENRY & T. A. HOSE

maps became available in textbooks and as large wall maps. Noteworthy amongst the latter was the Geological Map of the British Isles and Part of France, Showing Also the Inland Navigation by Means of Rivers and Canals, the Railways and Principal Roads (Knipe 1843), at a scale of 1 inch to 12 miles (1:760 320) by James Alexander Knipe (1803?–1882). Its second edition of 1859 was engraved and hand coloured on four sheets each measuring 31.75 × 27 in (80.65 × 68.58 cm); an inset map of the Orkney and Shetland Isles and nine geological sections are in the area occupied by the map’s German Ocean (North Sea), and its coverage of France extends southwards to include Paris and westwards to include the Channel Islands and part of Brittany. Knipe published a series of maps between 1837 and 1881 that included sheets of England and Wales, the British Isles and Scotland; it appeared in at least 50 different issues over 44 years (Toland et al. 2013, appendix). He was probably the most prolific non-OS publisher of British geological maps in the mid-nineteenth century. Table 3 lists those national maps available to the Derbyshire-bound geotourist that provided a geological context with some limited detail due to limitations of their scale.

Geological texts available to nineteenth-century geotourists The earliest ‘textbooks’ on English geology were by William Phillips (1773–1828), A Selection of Facts from the Best Authorities, Arranged so as to Form an Outline of the Geology of England and Wales (1818), William Phillips and William Conybeare (1787–1857), Outlines of the Geology of England and Wales (1822), and Robert Bakewell (1767– 1843), Introduction to Geology (five English editions, 1813–1838). They all incorporated reduced variations of William Smith’s 1815 map as a small foldout.6 Phillips was a founder member of the Geological Society and reputed by Henry B. Woodward, in his History of the Geological Society (1907), to be the best mineralogist among the founders. Woodward rated Bakewell’s text as the best of the early textbooks; it was the third edition that inspired Charles Lyell (1797–1875) to take up geology. After Lyell’s three-volume Principles of

Geology: Being an Attempt to Explain the Former Changes of the Earth’s Surface, by Reference to Causes Now in Operation (1830– 33), the publishing floodgates opened for geological texts to satisfy the growing public enthusiasm for geology and the controversies it generated (see Table 4). Texts well known to students of geology included: Horace B. Woodward’s The Geology of England and Wales: A Concise Account of the Lithographical Characters, Leading Fossils, and Economical Products of the Rocks; with Notes on the Physical Features of the Country (1876; and a second edition in 1887) had a large, 25 in by 21 in (62.5 × 52.5 cm) colour-printed geological map of England and Wales (at a scale of 1 inch to 69 miles, 1:4371 840). Physical geography at a national scale was not neglected with the publication of Sir Andrew Crombie Ramsay’s Physical Geology and Geography of Great Britain, which included a fine colour-printed geology map of Britain (1862; and to a sixth edition revised by Woodward in 1894), and D. Mackintosh’s Scenery of England and Wales (1869). The culminating text was Lord Avebury’s The Scenery of England and the Causes to which it is due (1902), a well-illustrated (with photos, diagrams and maps) volume of a style similar to modern textbooks. In the late nineteenth century, county geology accounts appeared in trade directories; one set of these accounts was the Geology of the Counties of England and of North and South Wales, 1882 by William J. Harrison (1845–1909); it did not include maps but referred to the relevant Geological Survey sheets at the head of each county’s chapter. By then, the Geological Atlas of Great Britain had been published by James Reynolds (1817– 1876) in 1860. It was a commercial response to the public’s interest in geology and ability to travel widely via the expanding railway network; it was published in a greatly expanded edition with revised maps in 1889. In Reynolds’s editions the double-page maps were the focus, while the text was relatively slight. The maps were hand coloured and readily understandable, with a foldout legend visible with all maps. Nominally British, the maps comprised English counties – singly and in pairs – with just two maps for Wales and a folded map of Scotland depicting mainly complex metamorphic and igneous geology, almost as addenda. This editorial

The sketch map (22 × 27.5 cm/8.5 × 11 inches) of England and Wales in Bakewell’s first edition of 1813 shows an extremely approximate division into three categories – primary, transition and secondary. While preceding Smith’s published map of 1815, it lacks the detail that Smith showed even on the schematic map that he had circulated since 1801 for the purpose of raising subscribers for his great map. The maps in Bakewell’s later editions were modified in light of Smith’s 1815 map but still did not incorporate the detail that was then possible and which Phillips and Conybeare showed in their maps of a similar size.

6

ROLE OF MAPS IN PEAK DISTRICT GEOTOURISM

143

Table 3. Selected national geological maps 1815 – 1850 Year 1815 1818

Author/publisher Wm. Smith/John Cary Wm. Phillips

1819 – 1820 G. B. Greenough/GSL* 1820 Wm. Smith/John Cary 1822 Wm. Conybeare and Wm. Phillips 1826 1831

J. Gardiner T. B. Loader

1834 1835

Greenough/Arrowsmith J. and C. Walker and James Knipe (without Knipe in 1836 and 1838) James Wyld (publisher)

c. 1835

1837 – 1881 James Knipe 1837 1839 c1840

John Phillips G. B. Greenough/GSL* R.I. Murchison/SDUK**

1843 – 1852 James Knipe. 1850

James Forbes/Blackwood

1851

T. B. Loader

Description 1 in: 5 miles (1:316 800); 23 legend units. 1 in: 47miles; text foldout, approx. 1:3 000 000; 25 legend units; includes substantial description of Mountain Limestone in Derbyshire 1 in:6 miles (1:380 000); 34 legend units 1 in: 16 miles (1:1013 800); 18 legend units. 1 in: 47 miles; text foldout, approx. 1:3 000 000; 27 legend units 1 in: 17 miles; reduction of Greenough/GSL map 1 in: 9 miles (1:570 200); re-issued in 1833, 1834, 1836 and 1842 1 in: 17 miles (1:1 100 000); 29 legend units. 1 in: 8.5 miles (1:538 600);n26 geological units Atlas page – British Isles, 1 in: ¼ 40 miles (1:2534 400); 21 geological units; later edition c. 1845 England & Wales, 1 in: 8.5 miles (1:538 600); re-issued in 1837 – 42, 1844 – 48, 1850, and 12 further editions 1856 – 1881 1 in: 30 miles (1:1 900 800); 17 legend units. 1 in: 6 miles (1:380 200); 42 legend units. Atlas page, 1 in: 28miles (1:1 774 100); 21 legend elements; re-issued in 1847, 1848, 1856 and 1864. British Isles, 1 in: 12.5 miles (1:792 000); released annually with revisions to 1852, 1854 – 1860 British Isles, two atlas double pages 1 in: 21 miles (1:1 330 000); 25 legend elements 1 in:10 miles (1:633 600); re-issued in 1858

*Geological Society of London. **Society for the Diffusion of Useful Knowledge.

balance reflected the market demographics and possibly the fact that the mainly sedimentary geology was easier to observe and differentiate over most of England. The 1860 edition was published before the Geological Survey was completed and relied on smaller-scale information from commercial sources. In the 1889 edition, the text was expanded, and the maps revised to include substantial geological boundary changes, annotations of geological features and fossil sites and additions to the railway network; on each county map, the sheet lines of the 1-inch geological survey maps were overprinted and there was a small, inset index map which showed the sheet numbering of these maps (see Fig. 6b). Shortly before this edition, in 1885, the OS had introduced the New Series layout and numbering for topographic maps; but geotourists would still have had the ‘Old Series’ geological maps. Edward Stanford (1827–1904) took over Reynolds’s map plates. He retained and updated the railway lines and introduced colour printing and 50 fossil plates, but otherwise kept the map design and Reynolds’s basic organization. Stanford’s first

edition in 1904 expanded the text to include the geology encountered along main railway routes, although none traversed Derbyshire (see Fig. 6). The text included small woodcut illustrations of scenery and geological cross-sections. As the Geological Survey had adopted the New Series layout in 1900 when it began colour printing, Stanford added an inset index map for the New Series, retaining the inset map for the Old Series but dropping the overprinted Old Series sheet lines. Stanford’s 1907 second edition added north and south maps of Ireland with text. The 1914 third edition added the Channel Islands. There was a 1913 Photographic Supplement to Stanford’s Geological Atlas of Great Britain and Ireland bound in matching covers in the same format and intended as a companion volume. It was cross-referenced to the second edition and referred to in the third edition. The photos, which measure 2.5 × 3.5 in (6.35 × 8.75 cm), are remarkably clear and finely printed. They are in stratigraphical order, with photographers’ acknowledgements and references to relevant contemporary memoirs and texts.

144

C. J. HENRY & T. A. HOSE

Table 4. Significant nineteenth-century geology texts Year

Author

Title and comment

1813

Robert Bakewell

An Introduction to Geology, Illustrative of the General Structure of the Earth: Comprising the Elements of the Science, and an Outline of the Geology and Mineral Geography of England, Richard Taylor and Co., London; further editions with revisions in 1815, 1828, 1833 and 1838

1818

William Phillips

A Selection of Facts from the Best Authorities, Arranged so as to Form an Outline of the Geology of England and Wales, William Phillips, London

1822

William Conybeare and William Phillips

Outlines of the Geology of England and Wales, William Phillips, London

1830

Charles Lyell

Principles of Geology; Being an Attempt to Explain the Former Changes of the Earth’s Surface, by reference to causes now in operation, John Murray, London; published with revisions through twelve editions to 1875 (Baldwin 2013, pp. 3 – 4)

1830

Henry De la Beche

Sections and Views Illustrative of Geological Phenomena, Treuttel & Wurtz, London

1831

Geological Manual, Treuttel & Wurtz, London

1835

How to Observe Geology, Charles Knight, London

1838

Charles Lyell

Elements of Geology, John Murray, London; published with revisions through six editions to 1865; third, fourth and fifth editions published as Manual of Elementary Geology.

1838

Robert Bakewell

An Introduction to Geology: Intended to Convey a Practical Knowledge of the Science, and Comprising the Most Important Recent Discoveries: With Explanations of the Facts and Phenomena Which Serve to Confirm or Invalidate Various Geological Theories, Longmans, London

1851

Henry De la Beche

The Geological Observer, Longman et al., London.

1860

James Reynolds

Geological Atlas of Great Britain, Reynolds, London; revised second edition in 1889

1862

A. C. Ramsay

Physical Geology and Geography of Great Britain, Stanford, London; five further editions to 1894

1869

D. Mackintosh

The Scenery of England and Wales, Its Character and Origin: Being an Attempt to Trace the Nature of the Geological Causes, Especially Denudation, by Which the Physical Features of the Country Have Been Produced, Longman et al., London

1876

H. B. Woodward

The Geology of England and Wales: A Concise Account of the Lithographical Characters, Leading Fossils, and Economical Products of the Rocks; with Notes on the Physical Features of the Country, Phillip & Sons, London; with foldout map of Britain, 25 × 21 in (62.5 × 52.5 cm) 1 inch to 69 miles; second edition in 1887

1882

W. J. Harrison

Geology of the Counties of England and of North and South Wales, Kelly, London

1902

Lord Avebury

The Scenery of England and the Causes to Which It Is Due, MacMillan, London; well-illustrated with photos, diagrams and maps

1904

H. B. Whitaker (ed.)

Geological Atlas of Great Britain, Edward Stanford, London

That Reynolds’s and Stanford’s atlases ran to a number of editions (up to 1964) indicates their popularity. They were pocket-sized at 7.5 × 5.25 in (18.75 × 13.13 cm) and intended for use on journeys. They were much used and, because of their low cost binding, few atlases remain intact; however, the individual maps have survived and are

popular, being small, affordable and colourful. Ideally, for the English, they were county-based rather than smaller scale national maps of Wales, Scotland and Ireland. The sequence of Reynolds’s and Stanford’s maps of Derbyshire with the revisions of the geology and railways and addition of supporting map information are shown in Figure 6.

ROLE OF MAPS IN PEAK DISTRICT GEOTOURISM

145

Fig. 6. Derbyshire geological maps from Reynolds’s and Stanfords’ Geological Atlas editions. (a) Reynolds, 1860. Published before Derbyshire was completed by the Geological Survey. Original map size 18.5 × 25 cm (7.25 × 10 in). The legend folded out so that it could be read alongside every map. (b) Reynolds, 1889. Incorporates much geological revision in the north, an inset map showing the Geological Survey map numbers and the sheet lines of the GS maps as a grid over the county. (c) Stanford, 1904. Introduction of basalt (red) and update of railways. Inset maps of Old Series (upper right) and New Series map layouts with removal of sheet lines. (d) Reynolds legend of 1889. The legend was very similar in Stanford’s 1904 edition with two additional igneous units (more reds) and the substitution of Devonian for Lower Silurian and revision of Upper Silurian to simply Silurian, none of which affected Derbyshire.

146

C. J. HENRY & T. A. HOSE

Nineteenth-century geological maps, sections, and guidebooks for Derbyshire Although the Geological Survey maps of Derbyshire did not appear until the mid-1850s there was much other and earlier geological mapping activity in Derbyshire. John Farey (1766–1826) met William Smith in 1801 at Woburn, where Smith displayed his stratigraphical table at agricultural fairs and sought subscribers for his intended national geological map. Farey, then the land steward at Woburn for the duke of Bedford, rapidly absorbed and appreciated Smith’s geological concepts and their potential extrapolation across England. He promoted Smith’s geological ideas widely7 and was employed by his patron, Sir Joseph Banks (president of the Royal Society for some 40 years), to prepare a mineralogical assessment of Banks’s estate at Overton near Ashover in Derbyshire. In 1807, Farey’s work for Banks led to an appointment, by the Board of Agriculture, to prepare a report on Derbyshire’s agriculture and minerals. As part of his published 1811 report, A General View of the Agriculture and Minerals of Derbyshire, Farey produced a geological map of Derbyshire on Smithian principles before Smith’s map was ever published. It was a small foldout map, at a scale of 5.5 miles to the inch (1:348 480) with the geological colouring extended well into the adjacent counties. The base map, with latitude and longitude, appears to be derived from Cary (see Fig. 7). Thomas West’s 1780 Peak District account, in an ‘appendix’ to his Lake District guide, has an itinerary familiar to anybody who has toured Derbyshire’s major caves and limestone crags. In Derbyshire there were several geology works available to the enquiring traveller by the late eighteenth century. Principal amongst these was John Whitehurst’s (1713–1788) An Inquiry into the Original State and Formation of the Earth; Deduced from facts and the laws of Nature of 1778 (Whitehurst 1778). Foldout geological sections at the end followed a conventionally theological text, almost in contradiction of it. It was the region’s first published geological account and went into further editions in 1786 and 1792. 7

Whitehurst, having noticed that fossil shells are much more common than fish, asserted that they were marine organisms created prior to terrestrial animals in accordance with the Bible. In accounting for the exotic rocks, such as supposed ‘Cornish moorstone or granite’ and ‘blue quartzose stone’, he envisaged violent convulsions of the Earth. He described Derbyshire’s geology in chapter 16, employing the term ‘subterraneous geography’; his sections were clearly the result of measurements in the field. He concluded that various beds did not retain an even thickness throughout and that they were fissured and faulted. He noted that mineral veins were distributed along joints and beds and that ‘toadstone’ (basalt) intersected them; further he recognized it as an intrusive rock, rather than lava, with contact metamorphism. His account mentions for the first time in print the term ‘slickenside’.8 John Mawe (1766–1829), a mineral dealer in Castleton, published The Mineralogy of Derbyshire in 1802. He travelled extensively in Britain collecting minerals for the family business – Brown, Son & Mawe, Petrifaction Warehouse. Due to Mawe’s industry, the business flourished to the extent that he could open museums/ shops in London, Cheltenham, Castleton and Matlock Bath. White Watson (1760– 1835), who is best known for his A Delineation of the Strata of Derbyshire of 1811 and A Section of the Strata in the Vicinity of Matlock Bath of 1813, was a close contemporary of Farey (Watson 1813). Delineations has no map but includes complex sections illustrating the faults and intrusions that make the geology of Derbyshire more complicated than the sedimentary sequences of southern England that inspired William Smith. The sections more clearly explained Derbyshire’s geological structure than it was possible to convey on a map. Watson is also known today for his inlaid marble ‘tablets’ with sections of the Derbyshire strata, slabs of polished Ashford Black Marble into which he inlaid the strata using samples of the actual rock types so that they were decorative and informative.9

Based on ‘the meritorious exertions and discoveries of this Gentleman (who was my Master and Instructor in Mineral Surveying’. Farey’s relationship to Smith and to other proto-geologists working in Derbyshire – White Watson, John Whitehurst and Elias Hall – is described in the preface by Hugh Torrens and Trevor Ford in the 1989 reprint of volume 1 of a General View of the Agriculture and Minerals of Derbyshire (Farey 1811, 110). 8 ‘These veins are productive of various Ores of Lead, as the Sulphure of leafy lamellar fracture, compact and spread over, which last is termed Slickensides’ (Watson 1811, pp. 50– 51). The meaning of slickensides morphed from this mineral description to its modern meaning of friction-polished surfaces associated with movement along geological faults and landslips. The lead ores of Derbyshire are associated with faults, and ‘lamellar fracture’ may have been caused by fault movement. 9 White Watson was a sculptor, marble worker and mineral dealer who lived most of his life in Bakewell. He was descended from at least three generations of stone masons and sculptors. He was an acute observer of Derbyshire’s

ROLE OF MAPS IN PEAK DISTRICT GEOTOURISM

Elias Hall (1764–1853) of Castleton in Derbyshire assisted John Farey in his researches for the agriculture and mineral report and, with Farey’s encouragement, advertised his services as a geological guide in addition to the sale of mineral samples which he collected (Torrens & Ford 2011, p. 250). Hall went on to produce models demonstrating geological structures and detailed geological sections through the Peak District and Lancashire. In 1834/35 he published a geological map of the ‘Coalfield of Lancashire and parts of Yorkshire, Cheshire and Derbyshire’ at 1 inch to the mile and coloured in the style of William Smith. As few copies of it survive, it is unclear how widely it circulated amongst geotourists. Another contemporary text was William Martin’s (1767–1810) Petrificata Derbiensia; or, Figures and Descriptions of Petrifactions Collected in Derbyshire, of 1809. He and Farey planned to produce a Derbyshire geological map, but this was thwarted by Martin’s premature death (Torrens 2004). William Adam’s The Gem of the Peak, was a popular guidebook published in six editions from 1838 to 1857; it included a final chapter that provided a ‘brief history of the flour spar’ and ‘remarks on the geology of Derbyshire’, which he also sold separately from his museum in Matlock. He also lectured on geology and guided geotourists (Ford 1973). Henry Moore’s Picturesque Excursions in the High Peak of Derbyshire Forming a New Buxton and Castleton Guide with Descriptive Accounts of the Scenery, Curiosities, and Singular Objects . . . of 1819 was one of the first illustrated guides to the region and the text is liberally sprinkled with geological information. Another and much revised early guidebook of considerable geological interest is J. M. Hedinger’s A Short Description of Castleton in Derbyshire; Its Natural Curiosities and Mineral Productions, which by 1839 was in its 26th edition. Although a slim volume of around 35 pages, it is packed with useful information, and the introduction notes ‘At this period, when Mineralogy forms the study equally

147

of Gentlemen and the Artist, the following short account . . . will I trust prove acceptable to that part of the Public who visit Derbyshire’ (p. 5). The 1871 Journey-Book of England: Derbyshire, part of an intended national set of such guides, is noteworthy for its excursions of some geological interest and the inclusion of a coloured county topographic map. The Rev. J. M. Mello’s Hand-Book to the Geology of Derbyshire, 1876, with its foldout, colour-printed, small-scale geological map and reduced Geological Survey sections and fossil illustrations, was so popular that it went to a revised and expanded second edition in 1891. At the end of the nineteenth century, Elizabeth Dale’s The Scenery and Geology of the Peak District was published. Expanded from a course of lectures she gave in Buxton in 1894, it incorporated photographs, sketches, sections and a foldout 1-inch (1:63 360) geological map centred on the Peak District, based on the Geological Survey (Dale 1900).

Nineteenth-century Derbyshire maps During the nineteenth century the mapping of Derbyshire’s topography was transformed from relatively simple to very advanced maps. Initially, small-scale maps with negligible to limited depiction of physical relief were produced by commercial surveyors, cartographers and publishers and usually displayed as county maps. In 1767, Burdett produced a 1-inch scale encouraged by the Royal Society of the Arts prize. The state, in the form of the OS, intervened to produce medium-scale (1:63 360) accurate topographic maps using sophisticated hachuring, then contouring, as part of a national series of abutting maps. The layout of the OS 1-inch maps for Derbyshire cut across the Peak District. Three Old Series maps were required to cover the core of the district and a further seven for the fringes (see Fig. 8).10 The OS maps of Derbyshire, published between 1838 and 1843, served as base

geology and collected books on geology; he maintained a reading room for friends and had access to the library of Chatsworth House. He sold his ‘tablets’ alongside catalogued mineral collections from his Bakewell museum shop, which was frequented by travellers. A self-educated geologist, he gave lectures on geology and was visited by contemporary eminent geologists and scientists – William Buckland, Andrew Sedgwick, Erasmus Darwin, Joseph Banks and James Sowerby. Alexandre-The´odore Brongniart (1770– 1847), an instructor at the E´cole de Mines in Paris, sent to him for specimens. Watson’s sections, in his book and as souvenir ‘tablets’, and his readiness to discuss the local geology greatly assisted visitors to appreciate the Derbyshire landscape. 10 The British Geological Survey (BGS) miscellaneous series Classical Areas of British Geology at 1:25 000 scale covers much of the core area in seven sheets; ideal for walking and appreciating the geology and landscape, with 25 ft/8 m contour interval and topographic detail reduced from 1:10 000 OS plans. At the end of 2013, all were in print (1970s editions) and could be purchsed for approximately twice the hourly minimum wage, while stocks last. (The adult minimum wage in April, 2013, was £6.19.) The BGS abandoned the printing of maps in 2012, but they are available online.

148

C. J. HENRY & T. A. HOSE

Fig. 7. John Farey’s geological map, 27.5 × 21 cm (11 × 8.25 in) of Derbyshire from his General View of the Agriculture and Minerals of Derbyshire, volume 1, 1811. Volume 1 of 3 deals with geology, whilst the others deal with agriculture. The map has no legend and relies on the text to enable the reader to identify the formations represented by each colour; the numbers on the formations are the clue and relate to the sequence of formations in the text. A copy of the

ROLE OF MAPS IN PEAK DISTRICT GEOTOURISM

149

Fig. 8. The Geological Survey of England and Wales 1-inch-to-the-mile map series cover of Derbyshire is shown on the right in extracts from the Old Series (a) and New Series (b) sheet lines. The New Series was introduced in 1885 by the OS to introduce one standard rectilinear sheet size, resolve problems of the original map projections and renumber the maps. It is still in use at the time of this writing, more than a century later. In both series the coverage of Derbyshire requires several sheets. Sheet 81 northeast, shown here, covers most of the far north of Derbyshire in the High Peak area; Buxton is located just south of the southeast corner. It was first published in 1840 as a topographic base map and in 1852 as a geological map. This was one of the smallest sheet sizes of the Old Series and covered an area of 11.5 by 14.5 miles (18.4 × 23.2 km). Hachuring had reached its highest standard of topographic representation in the OS Old Series and shows clearly through the light transparent water-colouring of this particular sheet. The red shows igneous intrusions into the blue limestones. The coal measures and millstone grits are not distinguished on this early issue. Less clear, due to reduction, are the fine gold ink lines running approximately east–west that signal metalliferous veins. John Phillips, William Smith’s nephew, geologically surveyed this sheet. (Courtesy of the Geological Society of London)

maps for the first edition of the Geological Survey maps published between 1852 and 1868. Official geological surveying began in the 1830s within the OS, as the Ordnance Geological Survey, and independently, as the Geological Survey of England and Wales from 1845, using the OS 1-inch topographic series as base maps as they became available (see Table 5). Non-Geological Survey geological maps adapted available commercial county and national maps as base maps and were aimed at a popular market. Commercial publishers, Reynolds followed by Stanford, recognized the popular interest and need for more accessible and affordable geological information and combined the county atlas format with revised geological and railway information.

Maps in popular Derbyshire guidebooks Until the late nineteenth century, when it became the norm, only a few of the available guidebooks included topographic maps. Geological images were popular (see Fig. 4). Rhodes in his popular, The Derbyshire Tourist’s Guide and Travelling Companion (1837) typically dispensed with maps, although he had incorporated a set in his 1824 Peak Scenery; or, The Derbyshire Tourist, because ‘The Road Sketches prefixed to this Volume are intended to supply the place of a Map, and it is hoped that they will be found equally useful . . . as instructions for their Excursions in Derbyshire.’ (p. x) As maps became more familiar, consumers recognized their utility, and guidebook publishers

Fig. 7. (Continued) map, annotated by G. B. Greenough in his copy of Farey’s report (held in the Library of the Geological Society), lists the formations as: (1) gravel, (2) lias, (3) red marl, (4) magnesian limestone, (5 and 6) coal measures, (7) grit [illegible text] shale, (8), limestone and toadstone (basalt), and (9) fourth limestone. The sequence is William Smith’s, which does not extend far into Derbyshire. Farey expressed the wish that Smith (Farey 1811, p. 116) would complete his map so that he could apply it to Derbyshire, but produced a schematic section (Farey 1811, p. 129) of the sequence of limestones, shales, grits and toadstones that sufficed. (Courtesy of Duncan Hawley)

150

C. J. HENRY & T. A. HOSE

Table 5. First Edition OS & Geological Survey Maps of the Peak District Sheet

71NW 72NE 81NW 81NE 81SW 81SE 82NW 82SW 88SE

First edition OS base

Geology

1839 1838 1842 1840 1842 1840 1840 1840 1843

1855 1852 1864 1852 1864 1852 1852 1852 1868

Cover

Place

Geologist(s)

Fringe Fringe Fringe Core Fringe Core Fringe Core Fringe

Wirksworth Ashbourn Stockport Chapel en le Frith Macclesfield Buxton Sheffield Matlock Penistone

W. W. Smyth, A. C. Ramsay, E. Hull W. W. Smyth, A. C. Ramsay, E. Hull A. C. Ramsay, E. Hull John Phillips A. H. Green, A. C. Ramsay J. Phillips and W. W. Smyth J. Phillips and W.W. Smyth J. Phillips and W. W. Smyth A. H. Green, J. R. Dakyns, J. C. Ward

responded positively. Maps for guidebooks had to cover the area in question and show sufficient detail. This was a challenge met by various solutions, including foldout maps and/or additional plans with detail. Burdett’s 1-inch (1:63 360) map can be recognized, by the hachuring, in James Pilkington’s 1789 A View of the Present State of Derbyshire with an account of its most remarkable Antiquities (in two volumes), although it was reduced to 4 miles to the inch (1:253 440). The Rev. R. Ward noted in his early guidebook that ‘the Peak of Derbyshire has often supplied a favourite theme to the tourist, and is visited by a multitude of strangers on very different accounts; by the invalid, in hopes of deriving benefit from its salutary waters; by the admirer of the beauties of nature, to view its delightful dales; by the botanist, to inspect the numerous indigenous plants with which its varied surface is clothed; by the mineralogist to investigate its minerals and fossils; and by the geologist to contemplate its diversified features, examine its strata, and explore its caverns and mines. The antiquary also is here gratified with sight of various objects peculiarly suited to his taste’ (Ward 1827, pp. 9–10). Ward describes the geology very much as Farey did and refers the reader to Mawe and White when discussing mines. His maps are five basic road ‘sketches’ with tables of distances, but without an overall map; however, he made no attempt to use Farey’s geological map. Rival guidebooks in similar small pocketable formats and often with red covers appeared from the 1870s. In the Ward Lock Guides,11 11

Bartholomew’s maps at ‘quarter-inch’ or 4-miles-to the-inch scale (1:253 440) appear first with hachures alone, then combined with contours. Contours provided actual heights, but in later editions when the hachures were dropped the visual potential of contours was not realized, because there were two contour intervals; on higher and steeper ground the wider interval meant that the visual impression of closely spaced contours was absent. The introduction of colour layering by Bartholomew in 1880 improved the perceptibility of high ground but not the detail. In Derbyshire the first colour-layered editions of Bartholomew’s maps appeared in Thorough Guides in 1891. By the 1880s, Bartholomew had moved away from supplying maps to other publishers and began to publish maps in their own name. Their ‘half-inch’ (1:126 720) maps with colour layering became a popular standard map series covering all of the UK, generally in blue but also in brown card folders. They began publishing a series for Scotland of individual folded half-inch maps in 1875 (Gardiner 1976, p.108). The separate England and Wales series was published in 37 sheets between 1897 and 1903. Two sheets were required to cover the Peak District, with Buxton on the boundary.12 Bartholomew’s maps developed over the decades, with improvements like coloured contours and scaled borders. The half-inch or 2-miles-tothe-inch maps were especially popular with tourists and cyclists. Their coloured contour system was more intuitive than the OS’s black-and-white 1-inch maps. They retained the required accuracy,

The Ward Lock Guides began, according to the company history, in 1854 with Ebenezer Ward (1819–1902) and George Lock (1832– 1891) in London (www.wardlockredguides.co.uk). No connection has been confirmed between the Rev. R. Ward, his printer William Ward already in the travel guide business in the 1820s, and Ebenezer Ward of Ward Lock & Co. 12 Bartholomew sheets 9 – Sheffield and 13 – Derby & Notts covered the Peak District. Buxton was located at the northern edge of sheet 9 and in an ‘extrusion’ into the bottom margin of sheet 13. It was not until 1932 that Bartholomew published a special Peak District map.

ROLE OF MAPS IN PEAK DISTRICT GEOTOURISM

151

Table 6. Cost of maps – prices converted from £.s.d. to current decimal £ 6a: Large national geological maps Author/map

Year

Wm. Smith/England and Wales G. B. Greenough/England and Wales

1815 1819

J. A. Knipe/England and Wales J. A. Knipe/British Isles

1843 1848– 1852 1858 1871

A. C. Ramsay/England and Wales

1859

A. Geikie/England and Wales

1897

Cost Wall map on rollers £7.00; as separate sheets £5.50* Wall map on rollers £6.00; to members of GSL £5.00† £2.10‡ £4.30 in case £3.00 in case £3.75 as coloured sheets; £4.20 as wall map on rollers. Folded colour print £1.25* (Ramsay was the new head of the Geological Survey.) Folded colour print £0.37* (Geikie was head of the Geological Survey.)

6b: One-inch OS maps – full folio sheets unless quarter-sheets specified Year

Cost

1805 1809 1811 1813

£3.75 for four £6.50 for eight £3.75 for five £2.18 for four

1810 – 16

£4.20 for five

1816 – 17 1818 – 20

£1.85 for two £2.175 for five

1837 – 48 1848 – 66

£0.35– 0.40 each £0.10

1866 – 1920

£0.125

2014

£7.99

Comment Essex only as set of four sewn together along west edge of maps Devon only as set of eight sewn together along west edge of maps Dorset only as set of five sewn together along west edge of maps Cornwall only as set of four sewn together along west edge of maps Sussex and Isle of Wight as set of five sewn together along west edge of maps; Isle of Wight available alone at £0.80 Surrey: individual sheets at £0.925 Pembrokeshire only as set of 5 sewn together along west edge of maps Quarter-sheets becoming available at a quarter of the costs Electrotyping of plates reduced production costs; quarter-sheets at £0.025 Quarter-sheets at £0.05 From 1885 OS (topographic) New Series maps began to replace Old Series map. The New Series had a uniform format the same or similar to the quarter-sheets of the Old Series. Standard equivalent scale 1:50 000 covers 40 × 40 km.

6c: Cost of 1-inch Geological Survey maps Year

Cost

1878

£0.15 uniform price for quarter-sheets – three times the cost of the OS base map Old Series £0.14– 0.48/quarter-sheet depending on amount and complexity of colouring; New Series £0.07 uniform price where available Old Series £0.18– 0.37/quarter-sheet depending on amount and complexity of colouring; New Series £0.10 uniform price where available £12.00; standard equivalent scale 1:50 000 still New Series, covering approximately 19 × 29 km

1909 1930 2014

*From contemporary advertisement’s in CJH’s collection. † Winchester 2001. ‡ Toland et al. 2013.

because they were based on OS mapping, but their reduced scale gave a greater coverage for the tourer than the 1-inch12; indeed, the OS was publishing competing colour road maps at the same ‘half-inch’

scale by 1914. The organization that initially did the most to promote road improvements and accommodation for cyclists, the Cyclists’ Touring Club (CTC), was acknowledged on Bartholomew’s maps

152

C. J. HENRY & T. A. HOSE

from the early twentieth century. Its logo appeared at the bottom of Bartholomew’s maps from around 1904 to 1924. Seemingly before 1904 ‘good’ and ‘inferior’ roads were not distinguished in their map keys; this information was added as the CTC’s members informed Bartholomew through a formal arrangement from 1910 to 1928. Bartholomew’s maps were often sold by map specialists or booksellers, such as W.H. Smith, in their own in-house covers; a practice which was later continued by petrol companies such as Shell and BP for their touring maps.

The availability and cost of maps and guidebooks Good quality topographic maps and many tourist guidebooks were fairly expensive right up to the mid-twentieth century. Table 6 shows the original prices of the maps discussed. To gauge their affordability some note of earnings and prices (McGrandle 1973, pp. 80– 82; Vaughan 1974, pp. 87–9; Priestley 1979, pp. 21–23) is helpful. Labourers’ wages in the mid-eighteenth century were (in modern decimal currency) 5–7.5p per day and after the 1820s averaged less than 50p per week. In the 1860s to the 1870s working-class people ‘in service’ (such as a butler) earned around £50– 80 a year and a footman £14–28, but this was not always in cash, due to deductions for essentials; concomitantly, factory workers earned £44 –88 and farm labourers around £31 a year. White-collar workers such as clerks earned £26 –60 a year. Around the same time, renting a couple of rooms (as was the normal domestic arrangement for many people) cost £6.50 a year. In the late 1880s, £2– 2.50 a week would keep a young, married, lower middle-class couple in some comfort. In the same decade, a coachman earned 80p to £1.25 and grooms 50–90p weekly. By the late 1890s a bank clerk, government clerk or teacher was earning some £150 a year. The first 1-inch OS maps, published in the 1800s, cost three guineas (£3.30). This was the weekly wage of one of its engravers and 20 days of wages for a craftsman (Hewitt 2010, pp. 166–167); it was more than a month’s wages for a farm labourer (Hobley 1964, p. 24). While such a price restricted maps to the social elite, their availability was somewhat wider, since they could be borrowed from the developing public and commercial libraries and from Mechanics Institutes to which the OS donated copies. Those same institutions and the emerging 13

natural science societies literally afforded workingclass and lower middle-class readers the opportunity to view the latest maps and read the latest geology texts for free or a small fee. Late in the nineteenth century, J. Bartholomew anticipated the tourist demand for easily comprehended maps that graphically and attractively depicted topography. Hence, he adapted the accurate OS series into a more useful and popular half-inch-to-the-mile (1:126 720) national map series employing colour layering. Bartholomew’s ‘half-inch’ maps, as they were popularly named, were very affordable. In the 1870s they cost 12.5p on paper and 17.5p on cloth; they were reduced in 1890 to 5p and 10p, respectively. Thus they were the equivalent of about a tenth to a twentieth of the working-class weekly wage, and less than a fiftieth of a middle-class weekly wage. By 1912, the price of the paper version was 7.5p, cloth-backed remaining at 10p and dissected at 12.5p.13 By late 1919 the price was updated to 7.5p paper, 15p on cloth, and 20p dissected. They were especially liked, initially by cyclists and then by motorists, because of the large area covered by each sheet. Elite travellers’ guidebooks were expensive for their time although they only cost from 5p to 12.5p in the 1780s. This was the equivalent of up to 2 week’s wages for a farm or building labourer and a day’s pay for a skilled craftsman; however, for a curate on £30 a year they were just about affordable. Rhodes’s The Derbyshire Tourist’s Guide and Travelling Companion (1837) probably had a wider readership for its 50p price than the earlier and larger volume it replaced, despite its relatively high price; the contemporary The Northern Tourist’s Guide to the Lakes of Cumberland, Westmorland, and Lancashire in its 1836 sixth edition cost 15p or, with six engravings, 27.5p. In the 1860s Abel Heywood published his Penny Guides (0.4p) with that for Buxton appearing in 1866 and Half-holiday Trips around Manchester in 1870. The low price, appealing to the lower middle-class public travelling on the railways, was achieved by extensive advertising and low-quality printing and binding. Ward Lock’s Illustrated Guides, which included the Bartholomew topographic maps, when first published in the 1880s cost 5p, and their Tourist Guides cost 12.5p; they were thus as affordable as the Bartholomew’s maps. As an interesting comparison, prices for populist published matter such as novels varied in the l860s from 17.5p to £1.57 and by the 1890s cost between 17.5p to 30p – somewhat dearer than the affordable maps and guidebooks of the period.

Dissected refers to the practice of cutting a map into uniform small rectangles which were mounted on linen so that the map could be folded along the linen with much reduced wear and tear. The dissected maps were often folded into card covers and could be shelved like a book.

ROLE OF MAPS IN PEAK DISTRICT GEOTOURISM

Conclusion In summary, at the beginning of our study period, 1780–1930, the state of mapping in England had progressed to the stage of facilitating navigation of the road network confidently but was in it is infancy in terms of assisting appreciation of the physical landscape. Nationally, cartographers such as John Cary and geologists such as William Smith, and later James Knipe, produced maps that greatly aided the budding geotourist until they were overtaken by better-resourced state institutions, the OS and the Geological Survey of England and Wales, later of Great Britain. In Derbyshire, with its rugged terrain and history of mining, there was much local talent. John Farey, mapping the Derbyshire estate at Ashover of Sir Joseph Banks, his mentor and president of the Royal Society, met such talents of the Dales as Elias Hall, White Watson and William Martin during the course of this work and his subsequent investigations for his agricultural report (see Fig. 7). He encouraged them to write and make maps, sections and models for geotourists, and to guide those same geotourists when they arrived in the Peak District. Over time the availability and affordability of maps and guides improved such that, by the opening of the twentieth century, even the geotourist of the most modest means could go into the field suitably supplied. This brief account demonstrates the truism that ‘Throughout the ages much research has been directed towards trying to understand the ways of the earth, but not until men . . . went out literally into the highways and hedges to see for themselves was there built a foundation on which others after them could work’ (Stebbing 1943, p. 4). Equally, such perambulations both led to, and gained from, improvements in topographic mapping and geological description. Some of these developments have been highlighted herein, as they pertain to the birthplace of geotourism (Hose 2008, p. 43) – the Peak District. The preparation of this chapter was aided by various anonymous contributors to digital book and map archives and online sources that helped to unearth scarce tourist guidebooks, geology texts and maps. CJH would particularly like to thank Ros Westwood and Anna Rhodes of the Buxton Museum, Duncan Hawley for advice on geological map history, and Richard Oliver for map price information. TAH would particularly like to acknowledge the assistance of the staff of The Geological Society Library, especially Wendy Cawthorne and Michael McKimm.

References Andrews, J. H. 2009. Maps in Those Days; Cartographic Methods before 1850. Four Courts Press, Dublin.

153

Baldwin, S. A. 2013. A Brief Bibliography of Sir Charles Lyell, FRS, Bt Geologist. Baldwin’s Scientific Books, Wickham Bishops. Burdett, P. P. [1791] 1975. Survey of Derbyshire. Reprint with an introduction by J. B. Harley, D. V. Fowkes & J. C. Harvey. Derbyshire Archaeological Society, Derby. Cary, J. 1794. Cary’s New Map of England and Wales with part of Scotland. On which are carefully laid down All the Direct and Principal Cross Roads, the Course of the rivers and Navigable Canals, Cities, Market and Borough Towns, Parishes, and most considerable Hamlets, Parks, Forests Etc, Etc. Delineated from Actual Surveys and materially assisted from Authentic Documents Liberally supplied by the right Honourable the Most Masters General. J. Cary, London. Dale, E. 1900. The Scenery and Geology of the Peak of Derbyshire. C. F. Wardley, Buxton. Davies, A. G. 1952. William Smith’s geological atlas and the later history of the plates. Journal of the Society for the Bibliography of Natural History, 2, 388– 395. De la Beche, H. T. 1832. A Geological Manual. 2nd edn. Treuttel & Wurtz, London. Farey, J. [1811] 1989. A General View of the Agriculture and Minerals of Derbyshire. Vol. 1. Reprint with introduction by T.D. Ford & H.S. Torrens. Peak District Mines Historical Society, Matlock. Ford, T. D. 1973. William Adam (1794?–1873) and Gem of the Peak. In: Adam, W. [1851] 1973. The Gem of the Peak; or, Matlock Bath and its Vicinity. An Account of Derby; A Tour from Derby to Matlock; Excursions to Chatsworth, Haddon, Monsal Dale, Dovedale, Ilam, Alton towers, Hardwick, Wingfield, Newstead Abbey, Ashbourne, Buxton, and Castleton. Historical and Geological; Brief History of the Fluor Spar, from the Earliest Period Down to the Present Time; A Review of the Geology of Derbyshire; Catalogue of Minerals and Rocks, and of the Flora of the High and Low Peak, Illustrated with Maps and Numerous Engravings, 5th edn, J. and C. Mozley, Derby. Moorland Reprints, Buxton. Fordham, H. G. [1925] 1976. John Cary; Map Engraver, Chart and Print-Seller and Globe-Maker, Cambridge University Press, Cambridge. Martino Publishing, Mansfield, CT. Gardiner, L. 1976. Bartholomew, 150 Years. John Bartholomew, Edinburgh. Handford, C. C. 1971. Some Maps of the County of Derby 1577–1850. Derbyshire Archaeological Society, Derby. Harbeson, J., Watts, N. & Harley, J. B. 1963. The Society of Arts and the surveys of English counties 1759– 1809. Journal of the Royal Society of Arts, 112, 43–46. Harrison, W. J. 1882. Geology of the Counties of England and of North and South Wales. Kelly & Co/Simpkin, Marshall & Company, London. Hedinger, J. M. 1839. A Short Description of Castleton, in Derbyshire; Its Natural Curiosities and Mineral Productions. 26th edn. S. Dodge, Stockport. Hewitt, R. 2010. Map of a Nation: A Biography of the Ordnance Survey. Granta, London. Hobley, L. F. 1964. Living and Working: A Social and Economic History of England and Wales from 1760– 1960. Oxford University Press, London.

154

C. J. HENRY & T. A. HOSE

Hose, T. A. 2008. Towards a history of Geotourism: definitions, antecedents and the future. In: Burek, C. V. & Prosser, C. D. (eds) The History of Geoconservation. Geological Society, London, Special Publications, 30, 37–60, http://doi.org/10.1144/SP300.5 Knight, C. 1871. The Journey-Book of England: Derbyshire. Charles Knight & Co., London. Knipe, J. 1843. Geological Map of the British Isles and Part of France, Showing Also the Inland Navigation by Means of Rivers and Canals, the Railways and Principal Roads, and Sites of the Minerals, under the Patronage of His Royal Highness Prince Albert, to whom this Map is by Express Permission Humbly Dedicated. H. Ballie`re, London. Martin, W. 1809. Petrificata Derbinesia; or, Figures and Descriptions of Petrifactions Collected in Derbyshire. D. Lyon, Wigan. Mawe, J. 1802. Mineralogy of Derbyshire: with a Description of the Most Interesting Mines in the Noth of England, in Scotland, and in Wales; and an Analysis of Mr. Williams’s Work, Intitled “The Mineral Kingdom.” Subjoined is a Glossary of the Terms and Phrases used by Miners in Derbyshire. William Phillips, London. McGrandle, L. 1973. The Cost of Living in Britain. Wayland, London. Mello, J. M. 1876. Handbook to the Geology of Derbyshire. Bemrose, London and Derby. Moule, T. 1837. The English Counties Delineated: or, A Topographical Description of England, Illustrated by a Map of London and a Complete Series of County Maps. George Virtue, London. Ogilby, J. [1675] 1971. Britannia, Volume the First: or An Illustration of the Kingdom of England and the Dominion of Wales: by a Geographical and Historical Description of the Principal Roads thereof. Actually admeasured and Delineated in a Century of Whole-Sheet Copper-Sculps. Accommodated with the Ichonography of the Several Cities and Capital Towns; and Completed by an Accurate Account of the More Remarkable of Passages of Antiquity, Together with a Novel Discourse of the Present State, Ogilby, London. Osprey, London. Phillips, J. 1844. Memoirs of William Smith, LL.D., Author of the “Map of the Strata of England and Wales”. John Murray, London. Pilkington, J. 1789. A View of the Present State of Derbyshire; with and Account of its Most Remarkable Antiquities; Illustrated by an Accurate Map and Plates. J. Drewry, London. Priestley, H. 1979. The What It Cost the Day before Yesterday Book. Kenneth Mason, Havant. Reynolds, J. 1860. Reynolds’s Geological Atlas of Great Britain. James Reynolds and Sons, London. Reynolds, J. 1889. Reynolds’s Geological Atlas of Great Britain. James Reynolds and Sons, London. Rhodes, E. 1824. Peak Scenery; or, The Derbyshire Tourist. Longman, Hurst, Rees, Orme, Brown & Green, London. Rhodes, E. 1837. The Derbyshire Tourist’s Guide and Travelling Companion Including An Account of the Various Places Generally Visited by Strangers in the County of Derby: to Which is Added the Detail of an Excursion from Dove-Dale to

Ilam Hall, and Alton Towers. R. Groombridge, London. Smith, W. 1815. A Delineation of the Strata of England and Wales with Part of Scotland; Exhibiting the Collieries and Mines, the Marshes and Fen Lands Originally Overflowed by the Sea, and the Varieties of Soil According to the Variations in the Substrata, Illustrated by the Most Descriptive Names. J. Cary, London. Simpson, S. 1746. The Agreeable Historian, or Compleat English Traveller. R. Walker, London. Stanford, E. 1904. Stanford’s Geological Atlas of Great Britain. Edward Stanford, London. Stebbing, W. P. D. 1943. Some early references to geology from the sixteenth century onwards. Proceedings of the Geologists’ Association, 54, 49–63. Toland, C., Ryder, K. & Torrens, H. 2013. The life and works of James Alexander Knipe (?1803–1882), British itinerant geological map maker. Earth Sciences History, 32, 279–312. Torrens, H. S. 2004. Oxford Dictionary of National Biography. Retrieved from www.oxforddnb.com [accessed 25 September 2013] Torrens, H. S. & Ford, T. D. 2011. Elias Hall, pioneer mineral surveyor and geologist in the Midlands and Lancashire. Mercian Geologist, 17, 249 –261. Toulmin Smith, L. 1907– 1910. The Itinerary of John Leland in or about the Years 1535– 1543, 5. George Bell & Sons, London. Vaughan, J. 1974. The English Guide Book c. 1780– 1870. David & Charles, Newton Abbot. Ward, R. 1827. A Guide to the Peak of Derbyshire Containing a Concise Account of Buxton, Matlock, and Castleton and Other Remarkable Places and Objects Chiefly in the Northerly Parts of That Very Interesting County. 7th edn. William Ward, Birmingham. Watson, W. 1811. A Delineation of the Strata of Derbyshire, Forming the Surface from Bolsover in the East to Buxton in the West. W. Todd, Sheffield. Watson, W. 1813. A Section of the Strata in the Vicinity of Matlock Bath. Derbyshire, Chesterfield. Whitaker, H. B. (ed.) 1904. Geological Atlas of Great Britain. Edward Stanford, London. Whitehurst, J. 1778. An Inquiry into the Original State and Formation of the Earth; Deduced from Facts and the Laws of Nature. To which is added an Appendix, containing some General Observations on the Strata in Derbyshire. With Sections of them, representing their Arrangement, Affinities, and the Mutations they have Suffered at Different Periods of Time. Intended to illustrate the preceding inquiries, and as a Specimen of Subterraneous Geography. Cooper, London. Wigley, P. (ed.) 2015. www.strata-smith.com. United Kingdom Offshore Geophysical Library [accessed 23 March 2015]. Winchester, S. 2001. The Map That Changed the World: The Tale of William Smith and the Birth of a Science. Viking, London. Woodward, H. B. 1876. The Geology of England and Wales. Phillip & Sons, London. Woodward, H. B. 1894. Geology in the field. Proceedings of the Geological Society of London, 13, 247– 273.

ROLE OF MAPS IN PEAK DISTRICT GEOTOURISM Woodward, H. B. 1907. The History of the Geological Society of London. Geological Society, London.

Further Reading Ambroziac, R. M. & Ambroziac, J. R. 1999. Infinite Perspectives; Two Thousand Years of Three-Dimensional Mapmaking. Princeton Architectural, New York. Andrews, J. H. 2006. A Paper Landscape: The Ordnance Survey in Nineteenth Century Ireland. 2nd edn. Oxford University Press, Oxford. Bates, D. G. 2010. Sir Henry Thomas De la Beche and the founding of the British Geological Survey. Mercian Geologist, 17, 149–165. Beresiner, Y. 1983. British County Maps: Reference and Price Guide. Antique Collectors’ Club, Woodbridge. Close, C. [1926] 1969.The Early Years of the Ordnance Survey. Institution of the Royal Engineers. David & Charles, Newton Abbot. Harley, J. B. 1975. Ordnance Survey Maps: A Descriptive Manual. Ordnance Survey, Southampton. Hellyer, R. 1999. Ordnance Survey Small-Scale Maps Indexes: 1801– 1998. David Archer, Kerry, Wales. Hewitt, R. 2010. Map of a Nation: A Biography of the Ordnance Survey. Granta, London. Knight, D. M. 1989. Natural Science Books in English 1600–1900. Portman, London. Nicholson, T. 1983. Wheels on the Road: Maps of Britain for the Cyclist and Motorist 1870– 1940. Geo Books, Norwich. Nicholson, N. & Hawkyard, A. 1988. The Counties of Britain; a Tudor Atlas by John Speed. Pavilion/

155

Michael Joseph in association with the British Library, London. Owen, J. & Bowen, E. 1720. Britannia Depicta or Ogilby Impov’d; Being a Correct Coppy of Mr. Ogilby’s Actual Survey of all y Direct & Principal Cross Roads in England and Wales: Wherein are exactly Delineated & Engraven, All y cities, Towns, Villages, Churches, Seats & scituate on or near the Roads with their respective Distances in Measured and Computed Miles. And to render this work universally usefull & agreeable, [beyond any of its kind] are added in a clear & most compendious method. A full & particular description & account of all the cities, borough-towns, towns-corporate & their arms, antiquity, charters, privileges, trade, rarities, & with suitable remarks on all places of note drawn from the best historians and antiquaries. Thomas Bowles, London. Owen, T. & Pilbeam, E. 1992. Ordnance Survey; Map Makers to Britain since 1791. Ordnance Survey, Southampton. Parker, M. 2013. Mapping the Roads; Building Modern Britain. AA Publishing, Basingstoke. Seymour, W. A. (ed.) 1980. A History of the Ordnance Survey. Dawson, Folkestone. Smith, D. 1985. Victorian Maps of the British Isles. Batsford, London. Smith, D. 1988. Maps and Plans for the Local Historian and Collector. Batsford, London. Trench, R. 1990. Travellers in Britain: Three Centuries of Discovery. Arum, London. Worms, L. & Baynton-Williams, A. 2011. British Map Engravers. Rare Book Society, London.

Three centuries of open access to the caves in Stoney Middleton Dale Site of Special Scientific Interest, Derbyshire MARK A. COPE AECOM, Royal Court, Basil Close, Chesterfield, Derbyshire, S41 7SL, England, UK (e-mail: [email protected]) Abstract: The limestone caves of Stoney Middleton Dale are a Site of Special Scientific Interest (SSSI) for features of geological interest, but have been open access since at least the eighteenth century and have a documented history of geotourism. Using the examples of Carlswark Cavern and Merlin Cavern, this chapter sets out evidence that 300 years of open access has resulted in significant disturbance to the features of geological interest. In particular, the integrity of the overall site for scientific research was affected by historical removal of speleothem formations long before it became a SSSI, with damage still occurring today. As most cave SSSIs in Derbyshire have historically had similar open-access arrangements, the examples presented highlight that there is potential for the integrity of less well historically documented caves elsewhere to have been disturbed. This article highlights the importance of establishing a baseline of historical disturbance to recognize whether the integrity of cave SSSIs have been affected and help monitor if disturbance is still occurring. For cave SSSIs to be useful for geological research, it is suggested that there is a need for the historical legacy of open access to individual caves elsewhere to be better understood.

Stoney Middleton Dale (hereafter referred to as ‘the Dale’) is located in the Derbyshire ‘White’ (Carboniferous Limestone) Peak District (Fig. 1), in the vicinity of the settlements of Stoney Middleton and Eyam. A large part of the hydrological catchment of the Dale is designated a statutory protected Site of Special Scientific Interest (SSSI), notified under Section 28 of the Wildlife and Countryside Act 1981 (as amended). The SSSI designation relates to features noted for geological and biological interest. A network of public rights of way passes though the Dale, making it popular with recreational users of the Peak District National Park. The A623, which is a busy route across the Peak District for cars and heavy goods vehicles, passes through the Dale, and a number of active and disused quarries are present near to and within the SSSI. Recreational and industrial uses of the Dale place great pressure on the SSSI designation, mainly associated with erosion, littering, noise and pollution of air and water. Derbyshire is recognized as the birthplace of geotourism (Hose 2008) and the settlements of Stoney Middleton and Eyam have historically attracted geotourists. Notably, Stoney Middleton was a stagecoach horse posting station on the route between Manchester and Sheffield (Cowen 1910). Local attractions were first recorded in travellers’ accounts (Bray 1783; Pilkington 1789), and later in dedicated tourist guidebooks (Rhodes 1824; Adam 1838). An account exists suggesting that geotourism was taking place as early as in 1734, with visitors being led into caves to view speleothems (Short 1734). Where the historical accounts

relate specifically to the caves of the Dale, these form a historical evidence base presented in this chapter. Caves may be considered hidden physical landscapes. Their location underground means that special trips have to be made to view them, and tourists are unlikely to realize that they even exist without their inclusion in guidebooks. Despite the caves of Derbyshire being mostly located out of sight and on private land, historically there are few physical restrictions on access in place. The caves of Derbyshire are in effect open access, and the caves of the Dale are no exception. This is in contrast to the Derbyshire show caves at Castleton, Buxton and Matlock. Access to the show caves has historically been restricted to guided tours for visitors arranged by the proprietor of the cave. It is recognized that visitors to caves cause disturbance to speleothems and sediments (i.e. Murphy & Chamberlain 2008), which is managed by the show caves out of necessity to protect their commercial interests. While the show caves together with many of the non-show caves in Derbyshire are afforded protection as SSSIs, it has only been since the 2000 Countryside and Rights of Way Act that it has been an offence for a third party (such as a visitor to a cave) to recklessly or intentionally damage a SSSI (Gray 2004). For a synopsis of the politics that have led to SSSI cave sites remaining largely open access, refer to Hardwick & Gunn (1996). It was identified by the author when undertaking a conservation audit for the caves of the Dale SSSI

From: Hose, T. A. (ed.) 2016. Appreciating Physical Landscapes: Three Hundred Years of Geotourism. Geological Society, London, Special Publications, 417, 157–169. First published online October 30, 2014, http://dx.doi.org/10.1144/SP417.3 # The Geological Society of London 2016. Publishing disclaimer: www.geolsoc.org.uk/pub_ethics

158

M. A. COPE

Fig. 1. Location of Stoney Middleton Dale.

(Cope 2014) that some descriptions of speleothems in historical accounts did not match the features as they are now found underground. While it is likely that the early writers exaggerated their descriptions (as will become clear in the language used in the historical accounts presented here), the distinct lack of speleothems in Carlswark Cavern and Merlin Cavern today is too stark to account for the differences alone. It is hypothesized that such differences represent a change in condition of the caves over the past three centuries, and therefore presents an opportunity to investigate such change. Where differences are found between the historical accounts and modern observations, there is potential for insight into the effect of open access on the caves.

Features of interest in Stoney Middleton Dale SSSI There has been much modern research interest in the caves and mines of the Dale (Jefferson 1961; Beck 1975, 1980; Ford 1977; Shaw 1984; Gunn 2002). Fifty-five specific cave-related features of interest have been identified by the author in and around the Dale, details of which are described in full in Cope (2014). It is not the intention to reiterate this detailed information here. This chapter concerns only the historically well-documented features of interest in Carlswark Cavern and Merlin Cavern. General features are described in the following sections, in particular features of geological and biological interest which are afforded statutory

THREE CENTURIES OF CAVE OPEN ACCESS

protection by the SSSI designation (refer to Natural England 1990). The caves are however also considered important for recreational activities, education and industrial heritage, and so features of interest relating to these uses are briefly considered.

Features of geological interest Of geological interest are approximately 5 km of actively forming and now inactive cave passage, from which four distinct stages of cave development have been identified (Beck 1975, 1980). These stages have been related to regional lowering of the groundwater table, with the lowest stage representing the modern groundwater level. The cave development strongly reflects the structural and stratigraphical geology of the area (Natural England 1990). For a review of the geological features relating to the SSSI citation, refer to Waltham et al. (1997). Sediment sequences have been noted by the author within the cave systems, which may be of geological research interest in the future (Cope 2014).

Features of biological interest Ancient woodland exists within the Dale, generally comprising ash (Fraxinus excelsior), wych elm (Ulmus glabra) and the introduced species sycamore (Acer pseudoplatanus), with buckthorn (Rhamnus catharticus) scattered throughout. The limestone crags are also of interest with a number of locally important species such as rock whitebeam (Sorbus rupicola), yew (Taxus baccata), field maple (Acer campestre) and mountain currant (Ribes alpinum). Dogwood (Cornus sanguinea) is present high in the Dale, a species now uncommon in Derbyshire. Important communities of mosses and liverworts occur in the swallets on the north-western edge of the Dale, in particular the nationally rare Spruce’s Leskea (Platydictya jungermannioides), Craven Featherwort (Pedinophyllum interruptum) and the liverwort Cololejeunea rosettiana. The Dale SSSI is also designated for grassland interest. The European cave spider (Meta menardi) is found in most cave entrances in the Dale. The author has seen bats (species undetermined) roosting in a few of the caves of the Dale during summertime, although these have only been noted in locations which appear to be seldom visited by anyone else.

Features of recreational interest The Dale is an important area for recreational caving. There is a chapter on caving in the Dale in the local caving guidebook Caves of the Peak District (Barker & Beck 2010) and caves of the Dale are also described in the national caving

159

guidebook Selected Caves of Britain and Ireland (Marshall & Rust 1997). A wide range of difficulty of the cave systems makes the caves of the Dale suitable for use by cavers of a wide range of abilities. Many of the caves can be visited within a few hours or less, and are located only 30 minutes by car from the populations of Sheffield and Chesterfield. Easy access to the caves mainly encourages evening time caving in the Dale, but outdoor instructors may be seen leading groups on most days of the week.

Features of educational interest The caves of the Dale are of general interest for education, due to the relatively easy access for novice cavers and closeness to the large populations of Sheffield and Chesterfield. In particular, the Gin Terrace and the cave passages between the Eyam Dale Shaft and Gin Entrance of Carlswark are two of the most accessible parts of the Dale for educational visits, which are frequently used by local outdoor education centres as an introduction to recreational caving. The Carlswark Eyam Dale Shaft and the Upper Shaft Entrance of Merlin Cavern are similarly used by the outdoor education centres to teach abseiling.

Features of industrial heritage interest Relating to the legacy of mining, there are a number of features common throughout the caves and mines of the Dale of industrial heritage interest. They demonstrate that miners gained access to many of the natural caverns known today, and provide evidence of associated mining technologies that can be dated. For example, shot holes drilled by historical miners are often the first indication that miners have entered a natural cavern, but which indicate a relatively late mining technology. Evidence of early mining techniques is largely absent from the Dale, except for one mine in Eyam Dale named Fireset Mine on account of the fire technique of mining that was employed there (J. Barnatt, pers. comm. 2010). Sough drainage is also of interest, as the lowering of local and regional groundwater tables was important in facilitating mining in and around the mining liberties of Eyam and Stoney Middleton. Records exist of numerous ventures to ‘free’ lead veins by the driving of soughs to artificially draw down perched water tables (refer to Rieuwerts 2007 for a detailed description of records relating to Sough ventures). Steam engines were also employed later for mine dewatering, for example at Watergrove Mine to the west of the Dale. Today only the most extensive soughs (but none of the engine shafts) remain easily accessible, which continue to impact greatly on the hydrogeology of

160

M. A. COPE

the Dale. For example, for most of the year water from Watergrove Sough to the west of the Dale is understood to drain east through Streaks Pot, the Nickergrove Mine Streamway, the Merlin Cavern Streamway, the Lower Carlswark Cavern Streamway, the ‘Boil-Up’ in Glebe Mine and eventually into Moorwood Sough to the east of the Dale (Fig. 2). In this instance sough drainage has greatly modified the apparent natural drainage route downdip along the Dale syncline.

Cave exploration and exploitation The Dale has a long history of exploration and exploitation, both above and below ground, dating back to Roman times (Cowen 1910). The village of Stoney Middleton is thought to have been named from being the middle town between a Roman settlement at Chesterfield and the Roman Station of Brough, but also perhaps with reference to the limestone cliffs that have been worked for stone there throughout the ages. It is likely that some underground exploration took place as workings in the cliffs enlarged the entrances to caves on the Gin Terrace; however, it has only been within the past two centuries that mining records indicate systematic mining for lead locally (Rieuwerts 2007). Little is known about historical tourist interest in the caves of the Dale, apart from what can be gleaned from the historical tourist guide books. Systematic modern exploration of the caves in the Dale has only taken place since the 1950s. Exploration was initially undertaken by members of the British Speleological Association, and the most significant discoveries were made following exploration in the 1960s and 1970s (Beck 1980). The frequency of modern discoveries has declined somewhat since then, but even at the time of writing discoveries continue to be made (J. Beck, pers. comm. 2012). The caves of the Dale are still visited by recreational cavers, but the author has also observed school groups visiting the more accessible parts of the caves with outdoor education professionals. Geologists are known to have used the caves of the Dale for research, with sediment and speloethem samples having been collected as part of research projects (Ford 1977; Shaw 1984).

Historical accounts and modern observations of Carlswark Cavern and Merlin Cavern This section provides a discussion of Carlswark Cavern and Merlin Mine from historical accounts, written prior to the re-entry of these caves by modern cave explorers in the 1950s. A comparison

is made with modern-day observations of features of interest as identified by the author, with reference to Cope (2014).

Historical accounts of Carlswark Cavern From the travellers’ accounts and tourist guidebooks noted in the opening text, geotourists are understood to have visited the cave system known today as Carlswark Cavern since at least the eighteenth century. Historically, the most significant part of Carlswark Cavern referred to was the ‘Wonder Cavern’ or ‘The Wonders’, so-called due to the speleothem formations it once contained. The Wonder Cavern was purportedly accessed from ‘Bamforth Hole’, located to the west of ‘Charleswork’. Charleswork is another historic name for Carlswark Cavern, and is interpreted to be the entrance known today as the Lower Resurgence Entrance (Fig. 2). The earliest recorded visit to Charleswork was made before 1734 by Dr Short, a Sheffield clinician on one of his ‘pleasure rides’. Short (1734) gave a detailed account of his visit in a footnote to the Natural, Experimental and Medicinal History of Mineral Waters, which is discussed in detail below. Later descriptions were also provided by Bray (1783), Pilkington (1789) and Rhodes (1824), but there is no evidence to suggest these later authors actually visited the caves as they only refer to Dr Short’s original account. The Lower Resurgence entrance to Carlswark Cavern is today entered though a small opening, approximately one square metre in size, at the foot of the cliff between the minor mineral veins Nut Scrin and Wonder Scrin. Short (1734, p. 95) describes entry to the cavern as ‘six yards high and eight wide’, which is considerably larger than the narrow opening forming the entrance today. The reason for the closing up of the entrance is recorded by Ebenezer Rhodes: ‘the entrance into it is nearly closed up by the falling of a mass of rubbish from above’ (Rhodes 1824, pp. 29– 30). Indeed, old bottles can be found near the entrance today, suggesting it has been used as a tip in the past. A spring often issues from the Lower Resurgence entrance during winter months, which can prevent access to the sumped passages beyond. This conforms with Dr Short’s description of ‘an unpassable deep stagnant Lake’. The Lower Resurgence entrance chamber and Lower Sump beyond were noted by the author to contain a silty deposit that has a faint odour of hydrocarbons. It appears that by the early nineteenth century the entrance was unpopular for this reason, as Rhodes refers to it as ‘a dreary hole’ and ‘a gloomy cavern’ which ‘even when more easy of access, it was but rarely visited’ (Rhodes 1824, p. 29). An abandoned

THREE CENTURIES OF CAVE OPEN ACCESS 161

Fig. 2. Plan of the caves and mines of Stoney Middleton Dale.

162

M. A. COPE

modern explorative dig in a silted tube is present in the entrance chamber today, with unstable mine workings on Wonder Scrin, possibly connecting with other mine workings above. Very rarely have water levels in the Lower Sump been known to be low enough to make passage though the sump possible. Approximately halfway through the Lower Sump a small chamber is met which has been found to contain shot holes, indicating that water levels in the past were low enough for miners to penetrate the sump. Short mentions that ‘this Cave reaches quite through the mountains and opens into Eynedale [Eyam Dale], which is above half a mile’, suggesting that a through trip was possible in the early eighteenth century. Such a trip is still possible but only during drought conditions, with exit to Eyam Dale possible via the Eyam Dale Shaft (Fig. 2). An alternative through trip is possible today between the Gin Entrance and the Eyam Dale Shaft, but it appears this was not possible at the time of Short’s visit. The modern Gin and Eyam Dale Shaft entrances to Carlswark Cavern are perhaps the two most frequently used cave entrances in the Dale, probably owing to the through trip that is possible between them. The Gin Entrance lies on the Wonder Scrin, almost directly above the Lower Resurgence entrance. Some mine workings are present inside and above the Gin Entrance, indicating that historical miners may have entered from within the Eyam Passage at this locality. A small tube once thought to connect naturally with the Eyam Passage is located on the surface further down the hillside, although there is no physical or documentary evidence that this or the modern Gin Entrance were historically used as an entrance to Carlswark Cavern. It is understood from Short (1734) that a mine level located a little to the west of the Carlswark Resurgence entrance, called ‘Bamforth Hole’, is the former historic entrance to Carlswark Cavern (Smith 1971). This mine level was investigated in detail by the author, including its possible continuation for 15 m along Wonder Scrin beyond a significant collapse of unconsolidated material from above. The likely location of its intersection with Carlswark Cavern at the other side of the collapse was also investigated. Physical connection could not be made given the scale of the blockage, so speculation regarding this route as an entrance remains. It is nevertheless reasonable to assume that, given the location of the mine level on Wonder Scrin, it would at one time have provided access to the Wonder Cavern. Indeed, Short’s detailed description almost exactly matches the morphology of the present-day route to Oyster Chamber, albeit the blockage on Wonder Scrin prevents access from the mined level.

Dr Short also describes Bamforth Hole as giving access to ‘a large Cave’ and describes in some detail the ‘Stalactitious Petrefactions’ that he observed. The detailed description of the cave also seems to match the morphology seen today in Oyster Chamber and Eyam Passage, although the fine speleothem formations no longer exist: ‘. . . your Toil is rewarded in seeing great Variety of Stalactitious Petrefactions. But leaving the Cave behind you are introduced into a most magnificent subterranean state Room nine Yards wide, and two high; the most statly, and awful Dome I ever saw, its Roof Floor and Sides are all shining with endless Numbers and Variety of beautiful transparent Statues, with several regular Ranks of fine Pyramids, and other curious Figures, some upon Pedestals, others reaching the Roof, others wishing to support this Reproach of Art, as afraid the Time might spoil the Glory of this solemn Dome, reach from the roof to the Floor. In the Middle of this Room is a Bason three Yards long and two wide, on each Side of it is a stately Pillar of Stalactites, one fine polished Marble and another in the Middle upon a Pedestal, through the Bottom of this is a very small Passage a few Feet down into another Entry to several other Caves still lower . . .’ (Short 1734, p. 95)

Comparison of Short’s detailed description with modern-day observations provides a useful inventory of the speleothem formations that have subsequently been lost, and demonstrates the historical removal of speleothems. Little evidence of Short’s ‘Stalactitious Petrefactions’ remain, apart from the bases of what would have been very large stalagmites (Fig. 3). These stalagmites in the passage appear to have been initially removed by miners prior to the twentieth century, as ‘shot holes’ (drill marks for the use of explosives) in the remaining bases are clearly evident (Fig. 4). Mining of speleothems is known to have taken place in Derbyshire to furnish ‘grottos’ in country estates, such as that in nearby Chatsworth House. In an early account of a tour of Derbyshire and Yorkshire on which William Bray passed through the Dale to Eyam, there is a hint at a local speolothem trade at that time: ‘Where the road turns off to Eyam . . . Mr. Longstone has . . . made a grotto with the spars, &c. found in the neighbourhood. One Benneson earns a livelihood here by collecting them, and has a number of specimens in his house.’ (Bray 1783, p. 178)

The removal of speleothems may also have been the result of tourists collecting their own specimens, as evidenced by William Wood in his History and Antiquities of Eyam, dated 1842: ‘Very near the Caelswark [Carlswark] is the cavern called the Wonders, which is explored by numberless strangers every year. It was once richly adorned with stalactites of innumerable forms, which have been taken therefrom to the cabinets of the curious.’ (Wood 1842, p. 159)

THREE CENTURIES OF CAVE OPEN ACCESS

163

Fig. 3. Oyster Chamber in Carlswark Cavern. Despite the morphology of the cave passage being almost identical to the description of Short’s Bamforth Hole, no evidence of the ‘Stalactitious Petrefactions’ remain apart from a few rounded bases of speleothems, such as that seen in the foreground of the photograph.

A brief untitled article written by the Derby Correspondent appeared in several national and regional newspapers a few years later in 1845, which reports that access to the Wonder Cavern was lost by a large-scale collapse. It reports that ‘The capacious cavern in Stoney Middleton Dale known as the Wonder Cavern is now closed up from visitants, probably forever’, by the collapse of ‘many tons of broken or blasted limestone’, and that it is fortunate that a party was not ‘entombed’. The report also refers to ‘numbers visiting the cavern’, suggesting it was popular with tourists at this time. This is an important detail, which places a likely earlier date range on the removal of the speleothems from Carlswark Cavern. It stands to reason that the cavern would not be so popular if all the speleothems had already been removed. There is also in a description in a 1950s cave guidebook with reference to Carlswark: ‘. . . interesting stalagmite formations and fossils in the upper passages to be seen . . .’ (Porteous 1950, p. 81), suggesting that there were at least some formations remaining into the mid-twentieth century.

Short’s reference to the ‘Pillar of Stalactites’ and ‘Pedestal, through the Bottom of this is a very small Passage a few Feet down into another Entry to several other Caves still lower’ is interesting. There is a narrow passage connecting between a hole in the floor of Oyster Chamber and a small chamber near to Lower Sump. Before the Carlswark Gin Entrance was discovered by the modern cave explorers in the early 1970s (Smith 1971), this section of passage was the only known means of entry to the upper passages of Carlswark Cavern. It would therefore appear that access between the caves was possible even before the removal of speleothem formations from Oyster Chamber. Although not mentioned in Dr Short’s description, it can be speculated that access to the Oyster Chamber was possible via the Lower Resurgence entrance (Smith 1971), even after the apparent blockage of the passage in Bamforth Hole.

Modern observations of Carlswark Cavern Loss of speleothem formations in Carlswark Cavern is known to continue today, occasionally observed

164

M. A. COPE

Fig. 4. Two miner’s shot holes in one of the speleothem bases located in Oyster Chamber provides evidence that historical speleothem mining has taken place in Carlswark Cavern. From Short’s description it would appear that this particular speleothem was a large column.

as new damage or vandalism particularly in the most heavily used Eyam Passage (Cope 2014). Disturbance from ‘wear and tear’ may also be attributed to the large numbers of cavers visiting Carlswark Cavern, since easy access became possible in the 1970s with the reopening of the Gin Entrance (J. Beck, pers. comm. 2008). The remaining speleothem bases in Oyster Chamber and in the Eyam Passage are very rounded today, which is probably a result of modern abrasion from cavers’ muddied oversuits due to the constricted nature of the cave passage. Indeed, polishing and scratching on exposed limestone surfaces is present to such a high degree in places that areas that were once deeply scalloped appear smooth and flat today (J. Beck, pers. comm. 2008). Disturbance of the calcite floor in the Eyam Passage has occurred, which has reduced in size quite considerably from that indicated on a cave survey plan of Carlswark Cavern from the early 1980s (Midgley 1983). Static pools of muddy water appear to be actively growing as a result of regular disturbance by cavers. It would appear that, as cavers pass through

the pools of water, ‘bow-waves’ from their movement flush sediments from beneath the weakening calcite floor (Fig. 5). Graffiti in the form of painted yellow arrows and green paint splashes in the passages between the Eyam Dale Shaft Entrance and the Gin Entrance have been observed several times, although each time was immediately removed by the author. A few abandoned digs remain in small tubes off the main passage, with discarded sediment fill strewn across the passage. The SSSI designated ‘unnamed tube’ is one such example. A large unsightly pile of boulders is also present, strewn across the main passage close to the Gin Entrance. This is believed to originate from the opening of Gin Entrance by cavers in the 1970s. A large amount of abandoned and decomposing cave exploration equipment was also noted. Other parts of Carlswark Cavern do appear much less disturbed, given that the difficult access contributes to the relative degree of preservation. There are no historical visitor accounts for comparison. Nevertheless, Aladdin’s Crawl, which is located in

THREE CENTURIES OF CAVE OPEN ACCESS

165

Fig. 5. A rounded speleothem base is all that remains of what would have been a large area of calcite flowstone in the Eyam Passage in Carlswark Cavern. Undercutting of the remaining flowstone is evident, as it is apparent that sediments have been flushed out from underneath the calcite flowstone leaving it overhanging. This large pool is shown as having a much smaller extent in a 1983 survey of the cave.

a difficult-to-access location directly above Rift Sump, is understood to have formerly contained fine speleothem formations (Barker & Beck 2010), indicating that even the more remote parts of the cave system are at risk of disturbance.

Historical accounts of Merlin Cavern The Merlin Cavern is located in an area historically known as the ‘Rock Gardens’ on the west side of Eyam Dale, close to the junction of Stoney Middleton Dale and Eyam Dale (Fig. 2). An overview of the Rock Garden as a tourist attraction is provided by William Wood: ‘The Rock-garden was once the greatest object of attraction in this romantic dell: it was the repository of all kinds of fossils, found in the Peak; and their dispersion has been greatly regretted by the inhabitants of Eyam. In this very interesting place is the Merlin – a cavern abounding, with wonder; but it is not so often visited on account of its being at times almost filled with water, which appears to rise from some subterraneous cavity . . . The pristine grandeur of this wonderful dale has been destroyed by the burning of lime, which is now carried on there to a great extent.’ (Wood 1842, p. 159)

The Merlin Cavern has a similar early history of being visited by tourists as Carlswark Cavern, albeit recorded in much less detail. On his tour of

Derbyshire and Yorkshire before 1777, William Bray noted: ‘Where the road turns off to Eyam, Mr. Longstone has . . . made a grotto with the spars, &c. found in the neighbourhood.’ (Bray 1783, p. 178)

In a similar description James Pilkington reported that: ‘Mr. Longsdon has raised a beautiful plantation, and in the midst of it formed a grotto, which he has furnished with some of the most elegant fossils collected in that part of the county.’ (Pilkington 1789, vol. 1, p. 23)

Later in the early nineteenth century John Farey refers specifically to ‘Merlin’s Cave’: ‘beautiful Stalactites well preserved, by the care of the late William Longsdon’. (Farey 1811, p. 294)

Jewett (1817, p. 93) differentiates between ‘the grotto’ and beneath it ‘the entrance of Merlin’s Cave’. And finally, in the early twentieth century in his history of Stoney Middleton Thomas Cowen describes the ‘Merlin Cavern’ as: ‘The cavern which is rich in stalactites and stalagmites, was re-opened a few years ago, but the roof has now fallen in places.’ (Cowen 1910, p. 29)

As with Carlswark Cavern, no speleothems of any great size or number remain accessible in the Merlin Cavern today. There is a description in a 1950s cave guidebook, ‘In Eyam Dale is beautiful

166

M. A. COPE

Merlin Cave’ (Porteous 1950, p. 81), suggesting that there were at least some formations remaining into the mid-twentieth century. However, a photograph also provided in the guidebook with the caption ‘Merlin Cave, Eyam Dale. Showing Early Stalactite Formation’ shows some small ‘straw’ speleothems, suggesting very minor regrowth of speleothems.

Modern observations of Merlin Cavern The main entrance to Merlin Cavern is an open adit level, so can be easily entered from the Dale. Inside the adit entrance level is a good example of bedding plane anastomosis, whereby initial cave formation is evidenced by the interconnection of small formerly active conduits within a bedding plane. This feature together with solution cavities elsewhere in Merlin Cavern indicate that, prior to historical mining activity, a naturally formed cave existed there. The Merlin Cavern is of heritage interest, namely in respect of its use for a time as a grotto (Bray 1783; Pilkington 1789; Farey 1811; Cowen 1910; Wood 1842; Rieuwerts 1960; J. Rieuwerts, pers. comm. 2012), although no physical evidence of this former use remains today. Surviving features relate to former lead mining activity including shot holes, stacked ‘deads’ (mining waste), stopes (open mine workings), cross-cut levels, winzes (internal shafts) and false floors (Barnatt & Penny 2004; J. Barnatt, pers. comm. 2010). Additionally, there is a surface shaft high in the roof of the mine that has been blocked from the surface and a run-in internal shaft at the north-western end of the pipe vein, which could have given past access to the Merlin Streamway. The author noted the remains of a wooden staircase on the floor of the adit level, which would have been used to gain access to a small solution cavity above (J. Rieuwerts, pers. comm. 2012). Instability was noted by the author in the passage on the intersection of the Merlin Pipe with Stub Scrin, in the shafts to lower levels on Stub Scrin (two of which are run-in) and in the floor below the second shaft on Stub Scrin. The second shaft presently gives access to the Merlin Streamway passages below. It is possible however that the instability on Stub Scrin conceals the precise location of the grotto referred to above, as there is no evidence that the Merlin Streamway passages were used as a grotto. There is no evidence that Speleothems existed in any great number in the Merlin Streamway. Furthermore, despite the statement by Wood (1842, p. 159) that ‘the Merlin . . . is not so often visited on account of its being at times almost filled with water’, the Merlin Streamway is accessed via an internal shaft and is an active stream route. It would seem very unlikely

that the Merlin Streamway could have been used as a grotto, visited by eighteenth century tourists. The Merlin Cavern Streamway comprises a series of enlarged collapse chambers aligned with the Merlin Pipe and interconnected by smaller tubes. Given the hydrological activity of the Merlin Streamway much of the cave passage is often flooded, preventing access without specialist diving equipment (except at times of drought). During the very dry summer of 2011 the author was able to reach Sump 8, close to the furthest point reached by modern exploration. Prior to the author’s visit, a group of cavers had made an earlier visit and had left their names inscribed in the freshly exposed sediments banks. As with the example of disturbance in Aladdin’s Crawl in Carlswark Cavern, the disturbance in Sump 8 of the Merlin Streamway demonstrates that even the most remote parts of the cave system are vulnerable to modern disturbance. Pristine natural cave passages named ‘Gimli’s Dream’ were discovered in the late 1970s from the passages close to the Eyam Dale Shaft in Carlswark Cavern, and were subsequently found to connect with passages in the Merlin Cavern Streamway (Beck 1980; J. Beck, pers. comm. 2008). Gimli’s Dream is of similar morphological character to the Merlin Cavern Streamway, but no longer contains an active stream. Features within the passages that evidence the sequence of cave formation, stagnation and reactivation of the stream are clearly present. The sequence begins with evidence of a stretch of relict phreatic passage that was affected by ‘breakdown’ (roof collapse) instability while the passage was still hydrologically active. The collapsed blocks are enveloped in sediment that consists of several fining-upwards sequences capped by a stalagmite layer, indicating gradual stagnation of the passage as a drainage route over time. The sediments and stalagmite layer are incised in places, indicating a later phase of vadose drainage route reactivation (Beck 1980; Fig. 6). There are few cave sites in Derbyshire where such a complete sequence of cave formation, stagnation and reactivation can be so clearly demonstrated, making Gimli’s Dream an important geomorphological site. Shortly after the discovery of Gimli’s Dream, vandalism and removal of speleothems took place (J. Beck, pers. comm. 2008). To prevent any further damage occurring as a result of caving through trips between Merlin Cavern and Carlswark Cavern, one of the local caving groups blocked the connection between Gimli’s Dream and Carlswark Cavern with concrete. For many years access to Gimli’s Dream was only possible from Merlin Cavern by descending the shaft on Stub Scrin. Curiously, however, nibbling-loss damage is still understood to have occurred (J. Beck, pers. comm. 2008). Many

THREE CENTURIES OF CAVE OPEN ACCESS

167

Fig. 6. A large block enveloped in sediments that are capped with calcite flowstone, and incised on the left-hand side. Although the flowstone, stalagmites and sediments appear to have been disturbed, the sequence of cave formation, stagnation and reactivation as interpreted by Beck (1980) in Gimli’s Dream is still identifiable.

Fig. 7. (a) A large stalagmite in Gimli’s Dream a short time after its discovery by Dr John Beck (pictured) in 1973 (photograph courtesy of J. Beck). (b) Thirty-five years later in 2008, Dr John Beck poses again at the same location although the large stalactite is no longer present.

168

M. A. COPE

of the remaining speleothem formations referred to in the SSSI designation have been muddied over or removed, and there is evidence of explorative activities (digs) in places which has resulted in further disturbance of sediment and speleothem features (Fig. 7). When Gimli’s Dream was discovered, the passages were found to be in a pristine natural condition with no evidence of being entered by historical miners/geotourists (J. Beck, pers. comm. 2008); all of the damage and loss of speleothems observed there today can therefore be attributed to modern open access.

Conclusions The caves of the Dale have a documented history of geotourism since before 1734 at least, with Dr Short’s recorded visit to Carlswark Cavern. Dr Short’s account indicates that speleothems were present in Carlswark Cavern at the time of his visit, and later accounts indicate that geotourists were attracted to this cave for a further 110 years until collapse of the entrance passages in 1845 prevented easy access. Similar accounts indicate that speleothems were also present in the Merlin Cavern in the late eighteenth and early nineteenth centuries when under the protection of Mr Longsdon. It would appear that the formerly more accessible parts of the Carlswark Cavern and Merlin Cavern had mostly been cleared of speleothems prior to access by modern cave explorers. History repeated itself after an extension linking these caves was discovered in the late 1970s, resulting in loss of speleothems in one of the most important cave geomorphology sites in Derbyshire. These examples of losses of speleothems, together with other examples of disturbance to the SSSI designated features of interest in Carlswark Cavern and Merlin Cavern described here, highlight the effect of long-term open-access arrangements to a sensitive cave system. As most cave SSSIs in Derbyshire have historically had similar openaccess arrangements to Carlswark Cavern and Merlin Cavern, the examples presented here suggest that there may also be historical and modern disturbance at less-well-documented caves. This chapter highlights the importance of establishing a baseline of historical disturbance to openaccess cave SSSIs where historical records are available, in order to determine if the integrity of a cave as a SSSI has already been affected and if disturbance to features of geological interest is still occurring. Without such a baseline, the integrity of such sites for geological research is unclear. Despite the increase in protection against thirdparty damage to SSSIs afforded by the Countryside

and Rights of Way Act 2000, the author is unaware of any other review of historical impact on cave sites. The nature of caves as hidden physical landscapes perhaps encourages an attitude of ‘out of sight, out of mind’. For cave SSSIs to be protected for geological research as they were designated, it is however suggested that there is a need for the historical legacy of open access to individual caves elsewhere to be better understood. I thank J. Beck for the wealth of information he provided on the caves of Stoney Middleton Dale, including the cave survey data and photograph used in this paper. I also thank J. Rieuwerts and J. Barnatt for the additional information provided regarding the Merlin Cavern. I dedicate this paper to the memory of John Salisbury Beck, who passed away before this paper went to print.

References Adam, W. 1838. The Gem of the Peak; or, Matlock Bath and its Vicinity, 4th edn. Longman & Co., London. Barker, I. & Beck, J. S. 2010. Caves of the Peak District, 7th edn. Hucklow Publishing, Great Hucklow. Barnatt, J. & Penny, R. 2004. The Lead Legacy. The Prospects for the Peak District’s Lead Mining Legacy. Peak District National Park Authority, Bakewell. Beck, J. S. 1975. The caves of the Foolow-Eyam-Stoney Middleton Area, Derbyshire, and their genesis. Transactions of the British Cave Research Association, 2, 1– 11. Beck, J. S. 1980. Aspects of Speleogenesis in the Carboniferous Limestone of North Derbyshire. PhD thesis, University of Leicester. Bray, W. 1783. Sketch of a tour into Derbyshire and Yorkshire, Including Part of Buckingham, Warwick, Leicester, Nottingham, Northampton, Bedford, and Hertford-Shires, 2nd edn. B. White, London. Cope, M. A. 2014. Stoney Middleton Dale SSSI Cave Audit and Conservation Management Plan, 2nd edn. Derbyshire Caving Association, Great Hucklow. Cowen, T. E. 1910. History of the village of Stoney Middleton. The Derbyshire Times, Chesterfield. Farey, J. 1811. General View of the Agriculture and Minerals of Derbyshire: With Observations on the Means of Their Improvement. Drawn Up for the Consideration of the Board of Agriculture and Internal Improvement. Sherwood, Neely and Jones, London, 1, 294. Ford, T. D. (ed.) 1977. Limestones and Caves of the Peak District. Geo Books, Norwich. Gray, M. 2004. Geodiversity Valuing and Conserving Abiotic Nature. Wiley, Chichester. Gunn, J. 2002. Location of Special Interest Features in Cave SSSI in the Peak District. Limestone Research Group Report 04/2002. Hardwick, P. & Gunn, J. 1996. The conservation of Britain’s limestone cave resource. Environmental Geology, 28, 121 –127. Hose, T. A. 2008. Towards a history of geotourism: definitions, antecedents and the future. In: Burek, C. V. & Prosser, C. D. (eds) The History of

THREE CENTURIES OF CAVE OPEN ACCESS Geoconservation. Geological Society, London, Special Publications, 300, 37– 60. Jefferson, D. P. 1961. The development of Middleton Dale Derbyshire. Bulletin of the Peak District Mines Historical Society, Matlock Bath, 1, 37–43. Jewett, A. 1817. The Northern Star, or Yorkshire Magazine, a monthly register of arts, biography, statistics, manufactures, & c. Baldwin, Cradock, and Joy, London, 1, 93. Marshall, D. & Rust, D. 1997. Selected Caves of Britain and Ireland. Cordee, Leicester. Midgley, M. F. 1983. Carlswark Cavern. Composite plan based on original surveys carried out by members of Eldon Pothole Club and Technical Speleological Group. Unpublished Survey Plan. Murphy, P. J. & Chamberlain, A. T. 2008. Cavers and geoconservation: the history of cave exploration and its contribution to speleology in the Yorkshire Dales. In: Burek, C. V. & Prosser, C. D. (eds) The History of Geoconservation. Geological Society, London, Special Publications, 300, 207– 215. Natural England. 1990. Stoney Middleton Dale Site of Special Scientific Interest Citation. World Wide Web Address: http://www.sssi.naturalengland.org.uk/cita tion/citation_photo/1004059.pdf Pilkington, J. 1789. A View of the Present State of Derbyshire. Drewry, Derby, 1, 23. Porteous, C. 1950. Caves and Caverns of Peakland. Derbyshire Countryside Ltd., Derby.

169

Rhodes, E. 1824. Peak Scenery or The Derbyshire Tourist. Longman, Rees, Orme, Brown & Green, London. Rieuwerts, J. H. 1960. The Merlin Mine Eyam, Derbyshire. Mining History, 1, 3 –6. Rieuwerts, J. H. 2007. Lead Mining in Derbyshire, History Development & Drainage. Vol. 1: Castleton to the River Wye. Landmark Publishing, Ashbourne, pp. 120 –148. Shaw, R. P. 1984. Karstic sediments, residual and alluvial ore deposits of the Peak District of Derbyshire. PhD thesis, University of Leicester. Short, T. 1734. The Natural Experimental and Medicinal History of the mineral Waters of Derbyshire, Lincolnshire and Yorkshire, Particularly those of Scarborough, Together with the Natural History of the Earths, Minerals and Fossils Through Which the Chief of them Pass. Published for the Author, London & Sheffield. Smith, M. E. 1971. Bamforth Hole, Stoney Middleton, Derbyshire. Comments on Dr Thomas Short’s Description of 1734. Bulletin of the Peak District Mines Historical Society, Matlock Bath, 4, 370 –374. Waltham, A. C., Simms, M. J., Farrant, A. R. & Goldie, H. S. 1997. Karst and Caves of Great Britain. Chapman and Hall, London, Geological Conservation Review Series, 12, 361–363. Wood, W. 1842. The History and Antiquities of Eyam: With a Full and Particular Account of Great Plague which Desolated that Village A.D. 1666. Thomas Miller, London.

Geology and landscape in SW England in the late eighteenth century, as recorded in the travel journals of William George Maton (1774 – 1840) JOHN D. MATHER Department of Earth Sciences, Royal Holloway, University of London, Egham, Surrey, TW20 0EX (e-mail: [email protected]) Abstract: In the summer of 1794 William George Maton, together with his friends Charles Hatchett and Thomas Rackett, embarked on a tour into Cornwall visiting the more southerly parts of Dorset and Devon en route. Rackett and Maton completed a second tour two years later, covering the rest of Dorset and Devon together with Somerset. An account of the tours was subsequently published by Maton, providing a contemporary description of SW England during the latter part of the eighteenth century. This was perhaps the first description of the region by scientifically aware travellers. They explored valleys, descended mines, visited smelters and collected minerals and must be regarded as among the earliest geotourists. Many sites which they visited, such as Roche Rock in Cornwall, Kent’s Cavern in Devon and Wookey Hole in Somerset, became major attractions for geoscientists in the following centuries. Discussions in the text suggest that the travellers looked at the rocks with neptunist eyes. Maton summarized the geological and mineralogical references on a map which used shading with lines rather than colour to differentiate individual strata. Although rudimentary and inaccurate, the map is of considerable historic importance.

SW England has been a popular destination for travellers at least since the early part of the sixteenth century when John Leland, in receipt of a commission from King Henry VIII, travelled through the West Country into Cornwall compiling a list of works in monastery and college libraries (Chandler 1996). His Itinerary contains mainly historical and archaeological information, with considerable detail on the religious houses he visited (Chope 1918). However, there is some information concerning natural history, including mention of creeks, rivers and quarries as well as a description of the hydrology of Loe Pool in Cornwell: ‘Lo [Loe] pool is a 2 miles in length, and betwixt it and the main sea is but a bar of sand. And once in 3 or 4 years what by the weight of the fresh water and rage of the sea it breaketh out, and then the fresh and salt water meeting maketh a wonderful noise. But soon the mouth is barred again with sand. At other times the superfluity of the water of Lo pool draineth out through the sandy bar into the sea’ (included in Gray 2000a, p. 12). The final sentence suggests some appreciation of the concept of hydraulic head. As with Leland, the accounts of subsequent travellers are influenced by their backgrounds and by the reasons for their journeys. Although Celia Fiennes, one of the best-known travellers of the late seventeenth century, commented on the mines and quarries she encountered, she spent much of her time visiting her relatives and her account is a valuable source of social information (Morris 1947). The abundant metal mines were also noted

by other well-known eighteenth century travellers such as Daniel Defoe and Benjamin Martin. Of greater interest are the accounts of travels made by Dr Richard Pococke in 1750 and the Reverend Stebbing Shaw in 1788 (Chope 1918; Brayshay 1996). These describe soils and building stones as well as giving limited information on the geology. Pococke describes the stone most used for building in Exeter as ‘a red crumbling free-stone with pebbles in it, dug out of the quarry at Hevitree [Heavitree] a mile from the town’ (Gray 2000b, p. 55) and Scott notes the red soils of Exeter which contrast with the ‘barren flinty common’ of Haldon to the south of the city (Gray 2000b, p. 69). It was not until 1797 that the first account of travels which can be considered primarily scientific was published (Maton 1797a). This was the result of two tours made in the summers of 1794 and 1796 involving three scientific friends, all of whom were destined to become Fellows of the Royal Society. The main instigators seem to have been Charles Hatchett (1765–1847), a mineral chemist, and Thomas Rackett (1755– 1840), a Church of England clergyman and antiquary. In 1794 they invited William George Maton (1774– 1835), who had known Rackett since he was nine years old (Paris 1838) and had recently graduated from Oxford University, to accompany them on a tour into Cornwall visiting the more southerly parts of Dorset and Devon en route. Two years later, Rackett and Maton (this time without Hatchett) completed a second tour, which covered the

From: Hose, T. A. (ed.) 2016. Appreciating Physical Landscapes: Three Hundred Years of Geotourism. Geological Society, London, Special Publications, 417, 171–185. First published online November 13, 2014, http://dx.doi.org/10.1144/SP417.8 # The Geological Society of London 2016. Publishing disclaimer: www.geolsoc.org.uk/pub_ethics

172

J. D. MATHER

northern parts of Dorset and Devon together with Somerset. During these tours Maton kept a detailed journal, noting the mineralogy, geology, antiquities and natural histories of the districts visited, and Rackett drew a series of sketches illustrating some of the scenery and antiquities. On completion of their tours, Maton felt that the information he had recorded was worthy of publication with ‘the hopes of directing public attention to a district hitherto very imperfectly described, and with the wish to assist the researches of those who may visit it with views and pursuits similar to my own’ (Maton 1797a 1, p. viii). His view was supported by the President of the Royal Society, Sir Joseph Banks (Paris 1838). Rackett placed his drawings at his disposal and 16 of these were rendered as aquatints by the London engraver, Samuel Alken (1756–1815). Details of the two journeys were published in separate volumes (Fig. 1), with a dedication to Rackett dated 22 April 1797. Although a second edition was foreseen ‘if the work should be thought worthy’ (Maton 1797a 1, p. viii), this was never to appear. The two volumes contain abundant references to the botany, soils and geology of the region and to the mines which the travellers encountered. These geological and mineralogical references were summarized ‘in the manner of a map, by which a general idea of the several transitions of substances may be obtained at one view’ (Maton 1797a 2, p. 201). This was described by Maton’s biographer (Paris 1838, p. 16) as ‘the first attempt in England to construct a geological map’. Maton’s map has been briefly described by historians of geology (e.g. Butcher 1968, 1983; Challinor 1971; Boud 1975; Mather 2013a) but the wealth of information scattered throughout the text on geological localities has been largely ignored, apart from a brief summary by Challinor (1971). For example, De la Beche records that Maton’s description of his tours ‘contains a geological map and numerous geological observations’ (De la Beche 1839, p. xxiii), but no further reference to it was made. The objective of the present paper is to highlight some of these geological references, which provide a view of localities and practices long since lost, through the eyes of scientifically aware observers. In the following discussion, references to Maton’s Observations are given by volume number and page(s). For example (1, 33) refers to Maton 1797a, volume 1, page 33. Maton’s spelling is followed but is clarified, when appropriate, by the modern equivalent in square brackets.

The travellers Thomas Rackett, the oldest member of the party, was a graduate of University College, Oxford

(BA, 1777; MA 1780) and rector of the Parish of Spettisbury with Charlton Marshall, located on the banks of the River Stour about 5 km west of Blandford Forum in Dorset (Hutchins 1868; Dewar 1965). The living, which he held for almost 60 years, provided him with a substantial income and he was able to pursue a range of outside interests. A distinguished antiquary and naturalist, he was an avid collector and his collections contributed substantially towards the formation of the Dorset County Museum in 1846, six years after his death. He was an active member of the Linnaean, Antiquarian and Royal societies and a constant presence at lectures of the Royal Institution, spending lengthy periods in London. This led to an accusation in the House of Lords in 1829 that he was neglecting his parochial duties, although this was later withdrawn. He corresponded with a wide range of people on a diverse range of topics and an editor of his papers considered him ‘One of the best examples of the type of country clergy who graced the 18th and early 19th century with their learning and accomplishments’ (Dewar 1965). Charles Hatchett was the son of a prominent coachbuilder who went into his father’s business. He was a wealthy man and equipped a private chemical laboratory at his home. A self-taught chemist and mineralogist, he established a reputation as a mineral chemist and, in 1800/1, discovered the element now called niobium (Griffith & Morris 2003). He also examined marine shells such as oysters, fish bones, fossils from Gibraltar and shark’s teeth, showing that the important constituent of bones and teeth was calcium phosphate rather than calcium carbonate, whereas shells contained little or no phosphate (Hatchett 1799). He was an early member of the Geological Society, elected on 6 January 1809 (membership number 130). Most of his scientific work was carried out between 1796 and 1806 after which, following his father’s death, he had to pay more attention to the family business. He was a prominent member of London society, a friend of William Hyde Wollaston, Joseph Banks and William Herschel, and frequently sat on public commissions when scientific expertise was needed. He was, for example, one of the Commissioners appointed to inquire into the prevention of the forgery of bank notes in 1818 (Anon 1819). The youngest of the travellers and the writer of the journals, William George Maton (Fig. 2), was the son of a Salisbury wine merchant. He went up to Queen’s College, Oxford in 1790, where he graduated BA in 1794 and MA in 1797 (Paris 1838; Rolleston 1942). Originally intended for the church, he continued his academic studies before transferring to medicine and gaining his MD in 1801. Shortly afterwards he was elected physician to the Westminster Hospital where his first few years of

TRAVEL JOURNALS OF WILLIAM GEORGE MATON

173

Fig. 1. Title page from Maton (1797a) volume 2 which describes the tour made by himself and Thomas Rackett in 1796.

practice were unproductive, with few patients. ‘His pecuniary resources being inadequate to sustain a long state of expectancy . . ..’ (Paris 1838, p. 19),

he began to reside at Weymouth during the season, a resort made fashionable in 1789 by the patronage of King George III and his family. Here, with little

174

J. D. MATHER

undergraduate (Maton 1792). Shortly after his return towards the end of 1794, he delivered a paper to the Linnean Society on a new species of Tellina [a bivalve mollusc] found in chalky parts of the bed of the River Avon (Maton 1797b). He was to follow these with a series of papers in the Transactions of the Linnean Society, including a long paper co-authored with Rackett on testaceological writers (Maton & Rackett 1804). This paper occupied 125 pages and was read over four sessions between February and June 1803. (Testaceological writers are those who write on the zoology of animals having a shell, especially a hard calcareous, unarticulated shell.) In addition he published numerous articles in antiquarian and medical journals (e.g. Maton 1800, 1815). He was elected to Fellowship of the Royal Society and served on its Council and as a Vice-President from 1831 to 1833.

The journeys

Fig. 2. Reproduction of an engraving of William George Maton, reproduced as a frontispiece in Paris (1838).

or no practice but much time on his hands, he was able to pursue his interests in botany. Through this he was introduced to the royal family and, with their patronage, his practice advanced rapidly, so much so that he retired from his position at Westminster Hospital in 1809 to concentrate on his private patients. He became Physician-Extraordinary to Her Majesty Queen Charlotte in 1816 and was chosen to attend the Duke of Kent when he was seriously ill in Sidmouth, Devon in 1820. Although unable to save his life, Maton was subsequently appointed Physician-in-Ordinary to the Duchess of Kent and her infant daughter, who was to become Queen Victoria. His practice increased and became one of the largest in London during the first third of the nineteenth century (Rolleston 1942). By 1827 he was able to pay off debts incurred by his father, for which he was given the freedom of Salisbury by a grateful corporation. Looking forward to retirement in 1834, he bought a country seat, Redlynch House, near Downton in Wiltshire but died within six months (Paris 1838). By the time of Maton’s first tour of SW England, he had already published a brief note on the origin of a seal discovered in Salisbury as a precocious

Both journeys began and finished at Salisbury with a total length of 729 miles (1173 km) for the first tour and 461 miles (742 km) for the second (Fig. 3). Although no indication of the mode of transport is given, there are references in the text and in Rackett’s letters to his wife (Gray 2000a, b) to walking their horses and it seems likely that much, if not all, of their journeys were completed on horseback. Away from towns, West Country roads were narrow and generally unsuitable for wheeled vehicles. William Marshall, author of the Rural Economy of the West of England published in 1796, noted that he did not meet a pair of wheels while travelling between Bideford and Swimbridge in North Devon, a distance of about 18 miles (30 km) (Chope 1918). Similarly, the Reverend Shaw, travelling in 1788 around Lydford to the west of Dartmoor, noted that the common vehicles of the country were horses with panniers and his party did not meet a single carriage all day (Chope 1918). As well as his comments on matters of scientific interest, Maton provides abundant topographical detail. Touring Somerset he noted that: ‘Few spots have a more lovely landscape than the hill above Banwell. We had an opportunity of contemplating it under the advantage of a beautiful setting sun, which, when sinking behind the Welsh mountains, gave a fullness to their outline, and displayed to us an infinite number of magnificent eminences swelling one above the other with an effect inconceivably sublime’ (2, 119). However, he did not look at everything through rose-tinted spectacles; St Just in Cornwall was described as ‘a sad dismal place, situated in a most inhospitable and cheerless corner of the county’. He noted that in the surrounding countryside ‘Scarcely a shrub appears to diversify

TRAVEL JOURNALS OF WILLIAM GEORGE MATON

Fig. 3. Routes taken by the travellers through the southwest in 1794 (red lines) and 1796 (blue lines). 175

176

J. D. MATHER

the prospect, and the only living beings that inhabit the mountainous parts are the goats which browse their scanty herbage’ (1, 221/2). He had an opinion on most towns and villages through which the travellers passed. In Dorset, Kimmeridge was ‘a miserable village’, whereas Lyme Regis was ‘a very pleasant town’. Torquay in Devon ‘Far exceeded our expectation in every respect instead of the uncomfortable village that we had imaged’. In contrast, Plymouth lived down to expectations as ‘an ill-built, disagreeable place infected with all the filthiness so frequent in sea-ports’ and St Michael [now Mitchell] in Cornwall was ‘a sad, mean place, and did not offer anything worthy of notice’. A favourite town was Taunton, in Somerset, ‘so large, well-built and handsome, that it may vie with most cities in the Kingdom as a place of trade and industry, perhaps exceeds them all, except for London and Bristol’. Little information is provided about the travellers themselves, the inns and houses at which they stayed, or how they coped with the conditions. Occasionally Maton bemoans the lack of suitable accommodation; for example, Bow in Devon was ‘a most wretched place, unable to afford the smallest accommodation of a decent kind’. A letter from Rackett to his wife (Gray 2000b, p. 81) sent from Exeter in August 1794 notes that they stayed one night in Lyme Regis, ‘it being too hot to travel for more than a few miles in one day’. Their route seems to have been decided as they went along as, in the same letter, Rackett is unable to tell his wife where to direct letters to him ‘as we are never sure where we are to stop’. It was from Exeter that he told his wife that ‘Mr Maton has lost his heart since he has been here, to a young lady whose name he does not know but whom he saw on the public walks’. These walks were probably those described by Maton near Rougemont Castle as ‘a very pleasant terrace, much frequented by the beau-monde of Exeter’ (1, 89).

Geological references The first journey Throughout his account of the journeys, Maton describes the local building stones and soils which the travellers encountered. The following review of his observations passes over these routine descriptions, concentrating on visits to specific localities where information of more particular interest is given. The first journey took the travellers south from Salisbury to the Isle of Purbeck (Fig. 3). On the way they passed the ball clay pits south of Wareham, noting the very fine white clay (much of which ended up in Staffordshire for the manufacture of earthenware; 1, 11). Travelling westwards

they visited the fossil (or stone) coal workings in the cliff close to Kimmeridge. Although nothing more than an ‘argillaceous slate in a high degree of impregnation with bitumen’, a demonstration soon showed ‘that it burned very freely and gave out a strong degree of heat’ (1, 33). Further west at Lulworth Cove, the shelly limestones close to the sea were observed to be similar to those at Peveral (Peverill) Point near Swanwich (Swanage) some 21 km to the east and, between the two localities, the whole range was seen to ‘make the same angle (about 45 degrees) with the horizon, or nearly so, pitching or dipping in general to the north’ (1, 46). At Portland the travellers were impressed by the freestone quarries. The stone, described as a calcareous grit, contained moulds or ‘larvae’ of various shells and in some places immense ammonites and regular veins of chert (1, 55–6). They proceeded westwards along Chesil Beach, a ridge of pebbles sometimes so loose that horse’s legs sank almost knee deep at every step. Most of the pebbles were of ‘white calcareous spar’ and they noted how their size gradually diminished as they travelled north-eastwards along the beach. Close to Portland, pebbles were from one to three inches (2.5–7.5 cm) in diameter whereas at Abbotsbury, some 9 miles (15 km) away they were ‘little larger than horse beans’ (1, 67). (The horse bean is a leguminous plant grown as food for cattle. The seed has a tougher outer shell than a broad bean with a maximum dimension of 1.0–2.0 cm.) Continuing onto the Dorset/Devon border close to Lyme Regis, they stopped at Charmouth to buy fine specimens of chalcedony from the ‘Curi-man’, noting the cliffs of ‘indurated marl’ abounding in ‘skeletons of fishes and other animals in a fossil state’ (1, 75). Some 10 miles (16 km) west of Lyme they crossed into Devon, noting how rich the soil was compared to the ‘frequent wastes, silent with desolation’ of west Dorset (1, 80). It was then on to Exeter where they visited the manganese mine at Upton Pyne, some 3.5 miles (6 km) north of the city. They found an opencast working where ore was dug from a ‘deep red viscid clay’. The ‘ore was in nodules of various dimensions, and generally crystallised in the inside’. The workings were about 20 ft (6 m) deep and the ore diminished ‘so much in richness in proportion to its depth’ that there was no need for a shaft (1, 93–5). The mine was very productive and the black oxide obtained was used to take away the yellow, green or blue tinge from glass to produce a clear white product. However, by the time of their second visit two years later, the ore had been exhausted and the mine filled up, with another opened up at Newton St Cyres some 3 km to the west (2, 47). According to Dines (1956), exploitation continued only until 1815. Some 4 km to the north of Upton Pyne, Maton

TRAVEL JOURNALS OF WILLIAM GEORGE MATON

was fascinated by the stone quarried at Thorverton, which he described as ‘one of the most curious substances, with respect to its composition, I have ever seen’ (1, 96). Since identified as an altered olivinedolerite, as a cautious mineralogist Maton declined to speculate upon its origin. Leaving Exeter for the southern coast of Devon, a diversion was made to see ‘some curious coalpits’ near Bovey-Heathfield. They were surprised to find the coal in ‘alternate strata with a whitish clay’ (1, 106 –8). They noted that the clay supported a pottery close by but did not relate it to the white clay they had seen previously near Wareham. The coal retained a vegetable structure and he had no doubt that it had once been wood. However, he was unable to suggest how ‘such an arrangement of strata was occasioned’. This was despite the fact that decayed roots and whole trees were seen in a large turf bog west of the pits, declaring that these did not ‘bear the least resemblance to Bovey Coal’ (1, 108). Travelling down to Torquay, the travellers visited Kent’s Hole (Kent’s Cavern) where they were guided by ‘two ancient females’ who were not ‘the most comely of their years’, aided by candles stuck in pieces of ‘slitted stick’ (1, 119– 21). The lights ‘when viewed at a distance, gleaming through the gloomy vaults, and reflected by the pendant crystals’ made them imagine that they were in the abode of some magician. The wet stone floor was difficult and dangerous and the roof was often so low that they had to crawl. As they rode from Torquay to Dartmouth, unlike many travellers they do not seen to have visited the famous ebb and flow well at Brixham (Atwell 1732; Mather 2013b). Travelling westwards to Ivybridge, a footnote (1, 128) describes the granite as ‘a dead whitish colour, and composed of a very large proportion of fel spar (which appears for the most part in long narrow crystals), pellucid quartz, some schorl, and a few scarcely discernable specks of mica’. By-passing Plymouth to the north, they crossed the Tamar into Cornwall and proceeded westwards, staying for some days with Hatchett’s friend Philip Rashleigh (1729– 1811) MP for Fowey. They admired Rashleigh’s mineral collection and grotto built with marble and serpentine together with brilliant crystals, pebbles and shells, which contained a table of 34 species of granite, all collected in Cornwall (1, 151). From his house at Menabilly they made excursions to the Poth tin streaming works, the Polgooth mine and Roche Rock. At Poth, about 4 miles (6.5 km) from Fowey, the rounded pebbles from which tin is extracted were found embedded in a bluish marl, mixed with sand (1, 152 –4). The principal stratum yielding tin was at a depth of about 20 ft (6 m) with a thickness of 6–7 ft (1.8– 2.1 m). Part of it had been worked previously before iron tools had come into use. Maton

177

thought that the tin-bearing sediments had been formed partly from the sea and partly from soil and fragments washed down from the surrounding mountains. At this time Polgooth, some 2 miles (3.2 km) southwest of St Austle (Austell), was one of the richest and largest tin mines in the county if not in the world (1, 155–65). The main ore lode, about 6 ft (1.8 m) thick, ran from east to west, dipping to the north. The ore was disseminated through a matrix of ‘caple’ (capel), a miner’s term for the rock occurring in the walls of a lode which, at Polgooth, was black and consisted of ‘a heavy kind of quartz which is perfectly opake [opaque], close textured and contains a large proportion of argill’. Maton had trouble obtaining definitions of mining terms and, for example, could never get a satisfactory definition of ‘elvan stone’, a term which he found to encompass a range of different intrusive rocks (1, 157). (According to the Oxford English Dictionary, the word ‘elvan’ is derived from the Cornish word elven, meaning a spark. An ‘elvan stone’ was therefore a rock that was so hard as to produce sparks when hit by a hammer.) It seems that only Maton and Hatchett descended into this and other mines as, in a letter dated 29 August 1794 from Truro, Rackett told his wife that he did ‘not intend to descend any of them as I am not fond of going down ladders’ (Gray 2000a, p. 76). Visiting Roche Rock (Fig. 4) just off the Bodmin–Truro road, they found it to consist of ‘a white sparry quartz, mixed with schoerl [tourmaline] which appears in innumerable needle-like crystals’ (1, 167). On their way to Falmouth, they noted the site at Grampound where Cornish china-stone was procured. They identified this as ‘merely a decomposed granite, the feldspar being become a soft lithomarga’. At Truro, this was made into retorts and crucibles of excellent quality and contained so little iron that porcelain made from it in Worcestershire and Staffordshire was ‘very little discoloured’ (1, 170). Rough seas meant that they were unable to cross the Helford River and had to make a detour by land to explore the Lizard Peninsula, where two guides had to be recruited and the weather was poor. Maton found the serpentine of Kynance Cove ‘extremely interesting to a mineralogist’ with its varying colour reminiscent of the undulating marks on a serpent’s back (1, 187). Proceeding to Penzance via St Michael’s Mount they visited the Wherry Mine, the descent into which was through the sea (Russell 1949). The upper part of the shaft resembled ‘an immense iron chimney, elevated about 12 ft (3.5 m) above the level of the sea and a narrow platform led to it from the beach’. The descent was by a rope tied round the thighs and ‘you were let down in a manner exactly the same as a bucket is

178

J. D. MATHER

Fig. 4. View of ‘Roche Rocks’ near Bodmin, Cornwall from Maton (1797a) volume 1, opposite page 168. This is a quartz-tourmaline rock formed late in the history of the Cornubian granite batholith. The chapel, close to the summit and dedicated to St Michael, has been a ruin since the early eighteenth century.

into a well; – a well indeed it is, for the water is more than knee-deep in many parts of the mine’ (1, 209). Near St Buryan they were taken to see the ‘Loggen-stone . . . an immense mass of granite, perhaps more than ninety tons in weight, and so exactly poised on the top of one of the highest rocks, that a child might move it’ (1, 215). At Land’s End they admired the granite scenery and learnt how to split the rock ‘by applying several wedges to holes cut (or pooled) in the surface of the stone at the distance of three or 4 inches (7.6– 10.1 cm) from each other’ (1, 219). Leaving Land’s End, they began their return journey by visiting the mining country along the north Cornish coast from St Just to St Ives. They passed numerous pits and Maton imagined that ‘ancient miners must have opened the ground to obtain tin in the same manner as we do stone quarries’ (1, 221). From St Ives they visited the large smelting house at Hale (Hayle) and the neighbouring copper mills. Maton was appalled by the appearance of the workmen in the smelting house, which was extremely hot with air impregnated with arsenic. ‘Some of the poor wretches who were lading the liquid metal from the furnaces to the molds

looked more like walking corpses than living beings’ (1, 233). Maton also questioned some of the miners about their health and it has been suggested that he was one of the earliest to show any interest in the field of industrial disease in this country (Harvey 1970). Maton’s account goes on to describe visits to a number of mines as the travellers moved through Cornwall from Hayle to Bodmin (1, 234–55), detailing the lodes mined, minerals found and technology used (1, 234–55). They visited Huel Mexico, the only silver mine in Cornwall, where silver occurred in a matrix of ‘ochraceous iron-ore’. Maton found it very dangerous to descend ‘on account of the ladders continuing quite strait to the bottom, and there being no resting place except a niche cut on one side in the earth’ (1, 253). Leaving the mining area he made the following general observations on what they had seen (1, 251–2): (1)

The tin appeared most commonly in the ‘calciform’ state otherwise called the ‘state of calx’. (These were terms used by early chemists for a powder or friable substance produced by

TRAVEL JOURNALS OF WILLIAM GEORGE MATON

burning or roasting. The calx was the essential substance of the mineral, usually the metallic oxide, after all the volatiles had been dispelled. The implication was that the tin was in the form of the oxide, cassiterite, although he never uses the term.) (2) The most prevalent matrix for the ore was either an ‘argillaceous’ or a ‘siliceous’ substance, or a stone made up of both which the miners called ‘caple’. (3) There were no calcareous minerals associated with the ore except for fluorspar. (4) Oxides of iron and arsenic were those with which the tin was most frequently associated. (5) Copper lodes lay deeper than those bearing tin and the richer lodes were generally accompanied by gossen (the leached and oxidized near-surface part of a vein). (6) Copper ores were mostly of the ‘pyritous and sulphureted kind’ and had a variety of matrices. (7) Tin and copper lodes appeared most frequently to have granite as the country rock and to make an angle of 60 –708 with the horizontal. Continuing north-eastwards they passed the Denyball (Delabole) Slate Quarries, noting that the best slate came from a depth of 30 fathoms (55 m) or more with the upper part of the quarries producing nothing ‘that is good for much’ (1, 258). They then followed the River Tamar south to Plymouth, passing en route the mines around Callington and St Kitt’s Hill where they identified wolfram in ‘great quantity’ (1, 272). In the lead mines at Bere Alston they collected ‘some fluor in octahedral crystals, which is a very rare mineral, and at no other place in England, I believe, is to be found’ (1, 276). In Plymouth they distinguished two distinct limestones, one hard and capable of taking a good polish and a second, finely bedded with a brownish-red colour resulting from iron oxide impurities (1, 290). Travelling north to Tavistock and then Okehampton they visited more copper and tin mines and were captivated by the spectacular waterfall at Lidford (Lydford) (1, 301– 3]. Proceeding quickly through Exeter, Honiton and Axminster, they soon re-entered Dorset and its bleak chalk downs. The onward journey to Salisbury was geologically uneventful, and there are no further observations of significance.

The second journey Maton and Rackett commenced their survey of the northern part of Dorset from Shaftsbury, a town whose situation was so high that it was acutely short of water. The engine that had once pumped water 300 ft (91 m) perpendicularly to supply the

179

town had been neglected and water was being carried from a distance either on heads or by horseback (2, 3). Further south, observations made during well digging around Sherborne provided them with information on the ‘succession of the strata’, and changes in vegetation showed them when they passed into chalk (2, 14 –5). On the way to Ilchester in Somerset, they passed through the village of Broad Marston and searched for a ‘congeries of ammonitae’ which had been discovered in 1778 at the opening of a marl pit (2, 21). They found this jumble of ammonites, individually between 0.25 and 1 inch (0.6–2.5 cm) in diameter, in a bed about 8–9 inches (0.02–0.23 cm) in thickness and about 8 ft (2.5 m) below the surface of the ground. As they journeyed westwards they admired the local building stones; Ham Stone, a freestone which contained a multitude of shells, differed from Portland Stone only in being tinged with iron oxide and ‘blue lyas’ (lias), of which most of the cottages were built. As they approached Taunton, Maton noted rounded pebbles at various depths, supposing these ‘sure testimonies of a soil formed by deposition from some permanent current, or bed of water’ (2, 37). Travelling south westwards into Devon they passed through Exeter for the third time, noting a fertile red soil running eastwards from Bow but only 3 miles (5 km) at the most in breadth. Subsequently heading north through Torrington and Bideford, they reached Hartland and the cliffs between Hartland Point and Quay. Here the cliffs presented a ‘black slatey front towards the sea’ (2, 64). The spectacular folding was not interpreted as such, but Maton noted the ‘singular position of the broad laminae [beds of sandstone] projecting from this promontory’ which ‘meet each other (in some places) like the timbers on the roof of a house, and diverge from a common line like the down on a quill’, a romantic description of the plunging folds. Retracing their steps eastwards through Bideford and Barnstaple, the travellers headed for Combe Martin, a village surrounded by mineral lodes. As well as the iron and lead which they expected they found the galena to be rich in silver, yielding 20 – 168 ounces (567 –4760 g) of silver per ton. The east –west-striking veins appeared close to the surface and had been worked ‘at a trifling expense’ (2, 78). However, the shallow ores had been worked out and the villagers were ‘anxious for some spirited exertions being made by the opulent’ to invest in deeper workings (2, 78). Travelling eastwards through rugged terrain towards Linton (Lynton), they suddenly came to a wild valley ‘bounded by large naked rocks, or rather fragments of rocks, piled one upon another’ (2, 87). This was the ‘Valley of Stones’ (Valley of

180

J. D. MATHER

the Rocks; Fig. 5).They decided that the scenery was the work of nature rather than man and speculated that the valley must have been the course of a vast and violent torrent which poured itself into the Severn at its western extremity. As a botanist, Maton was keen to look at the relationship between mineralogy and vegetation, commenting that ‘Sudden variations there in the composition of rocks may often be discovered at merely a glance by becoming acquainted with their more obvious vegetable inhabitants’ (2, 90). Proceeding further along the north coast of Devon into Somerset, they reached Watchett where they recognized that rocks cropped out which were similar to those in Glamorganshire, on the northern side of the Bristol Channel (2, 107). Ammonite-rich limestones also cropped out to the east, similar to those they had seen previously at Lyme Regis. They were intrigued by the alabaster along the Watchett cliffs where ‘calcareous matter seems to have oozed as it were, from the marl’ (2, 107). Travelling north into the Mendip hills, they established their base for a while at the village of Cheddar where they admired ‘the Alps of Somersetshire’ (2, 126), noting strong similarities with the

rocks of the Peak in Derbyshire. Both abounded with ‘veins of lead and calamine’, both contained ‘vast caverns and subterraneous vaults’ and both consisted of ‘a similar species of stone’ (2, 126). Maton describes the mining of calamine (an obsolete name for ZnCO3) and lead, in the form of galena, in workings known as grooves, the miners being known as groovers. Maton went into galleries where working conditions were so confined and air so foul that he nearly suffocated within a quarter of an hour. Stealing tools or ore from others was punished by ‘the burning of the hill’, where the offending thief was shut up in a small hut on the hill, around which straw and dry wood was placed, the whole then being set alight. Allowed to make his escape as best he could, the offender was invariably singed and half suffocated and was never allowed to dig on the hills again (2, 135). The travellers visited Wookey Hole, more spacious than Kent’s Cavern, near Torquay. They were conducted through a network of caverns about 600 ft (183 m) in length, with ‘sparry projections from the roof and sides’ (2, 137). Many of these caverns had been given names ‘by the vulgar’ such as ‘witch’s brewhouse’, ‘furnace’ and ‘parlour’.

Fig. 5. View of ‘Valley of Stones’ near Lynton, North Devon from Maton (1797a) volume 2, opposite page 84. The view shows the Devil’s Cheesewring, with Castle Rock in the left background and Rugged Jack to the right. The crags are formed from the Lower Devonian Lynton Beds, a group of marine mudstones, siltstones and fine-grained sandstones with occasional thins beds of fossiliferous limestone.

TRAVEL JOURNALS OF WILLIAM GEORGE MATON

After visiting Glastonbury and Wells they headed towards Bath and took the opportunity to learn something of the mines of the Somerset Coalfield. They were able to work out a local succession of strata overlying the coal (2, 161– 2) and described workings at Radstock, Camerton and around Midsummer Norton. During the summer of 1796, William Smith was working on the Somerset Coal Canal (Torrens 2001), although it seems unlikely that the travellers encountered him. Maton described the Bath Stone as a ‘sort of oolithus occasionally mixed with spar and shells’ (2, 166), noting that the city was wholly surrounded by limestone hills. Towards Keysham he was fascinated by the giant ammonites which were carefully picked out and polished for sale by the quarrymen, who called them snake stones (2, 167). Re-entering Dorset on their way back to Salisbury, Maton commented on the ‘linchets’ (lynchets) which were common on the chalk downs. Now interpreted as ledges formed on the downhill side of cultivated plots by ploughing in ancient times, he thought they were natural terraces formed by ‘subsidences of the ground in a state of solution’ (2, 187). He visualized waves of mud sliding down the hillsides and felt that this accounted for the roundness of the brows of chalk hills ‘because if their upper strata had not being originally fluid, a sharpness of edge must have been left’ (2, 188).

The mineralogical map Maton concluded the account of his travels by summarizing the many observations on soils, minerals and rocks, scattered throughout the two volumes, in the form of a mineralogical map (Fig. 6). His observations were fragmentary and he recognized that all he could do was to show ‘the grand stretch of the different strata, and the most prevalent substance in the composition of each’ (2, 202). He decided to do this using shading with lines, on the basis that differences and relationships between strata could be better shown by different directions and combinations of lines than by using colours. The key to the map shows that four strata types are represented by simple straight lines, horizontal lines for clay, vertical for chalk, diagonal lines proceeding downwards from the right for quartz and from the left for serpentine. Combinations of these lines are then used to represent the other strata and show the affinities between them. This is illustrated in Figure 7 which is a segment of the main map showing strata to the north of Exmouth. To the east, the vertical straight lines of chalk are undulated to represent limestone. This has no affinity with the clay to the west and the junction is sharp and well defined. However, the clay and strata to the west of it all have what Maton termed an ‘argill’

181

component. In consequence, the horizontal lines representing clay are continued laterally into the argillaceous gritstone, where they are combined with diagonal lines indicating the ‘quartzose’ component of these strata, and then into the argillaceous slate where the lines are undulated to show the schistose (or slatey) nature of these rocks. The connection of lines across boundaries indicates that the different strata have the ‘same substance in their composition’ (2, 204), in this case the ‘argill’ component. Taking this further, the diagonal lines in the gritstone continue across the junction with the granite, indicating that both have a ‘quartzose’ component. The opposing diagonal lines in the granite shading are used separately for the magnesium-rich mineral serpentine, but are also used by Maton to represent the ‘magnesian’ component in the granite which is present in mica (2, 205). A more detailed discussion of the principles behind the shading techniques used is given by Mather (2013a). Maton claimed that the use of what he termed ‘mineralogical signs’ to characterize strata could be varied to infinity by doubling, undulating or interrupting lines, ‘a convenience that cannot be obtained by colours’ (2, 206). A ruler and a pen or pencil were all the materials necessary for a traveller to carry with him in order to delineate the strata on a map and he ‘hoped to see very shortly a complete picture of the mineral face of the whole island’ (2, 208). Unfortunately his map is over-generalized, boundaries are inaccurate and what is on the map is often contradicted by statements in his journals. For example, all the rocks to the east of the belt of north –south-trending clay (Fig. 6), which includes nearly all of Dorset and East Somerset, are shown as limestone or chalk, totally ignoring the Wareham clays, the marls he saw around Lyme Regis and the Somerset Coalfield. In Devon and Cornwall the granite is shown as a continuous outcrop down the spine of the peninsula, rather than a series of discrete outcrops. His map would have been of questionable value to travellers and its significance lies in the fact that it was the first attempt to produce a regional geological map ‘of any large part of the country’ (Challinor 1971, p. 79) rather than in its intrinsic value.

Discussion Maton was primarily a botanist and it was Charles Hatchett who was the most knowledgeable mineralogist among the travellers. He did not take part in the second journey and this is reflected in the more detailed geological commentary provided in volume 1 of the travel journals in comparison to volume 2. For example, there are accurate descriptions of the tin and copper ores of Cornwall with references to chemical analyses made by the

182 J. D. MATHER

Fig. 6. ‘Mineralogical Map of the Western Counties of England’, annexed to Maton (1797a) volume 2.

TRAVEL JOURNALS OF WILLIAM GEORGE MATON

Fig. 7. Detail of Maton’s mineralogical map, showing strata to the north of Exmouth. An explanation of the line shading is given in Figure 6.

German chemist Klaproth (Martin Heinrich Klaproth 1743– 1817) (1, 155; 1, 243) and the Danish mineralogist Brunnich (Morten Thrane Bru¨nnich 1737–1827) (1, 154). There is also speculation on the formation of selenite and gypsum on the Isle of Purbeck, with the suggestion that the sulphuric acid concerned was derived from the pyrite common in interbedded limestones (1, 21). The minerals of the Dartmoor granite are correctly identified (1, 128), as are those of Roche Rock, a distinctive quartz/tourmaline rock formed late in the history of the Cornubian batholith. A footnote comments that, according to the French naturalist M. d’Aubenton (Louis-Jean-Marie Daubenton, 1716–1799), the latter rock is a granitello, ‘a name which Mr Kirwan proposes, with great propriety, to apply to all binary aggregates of the granitic kind’ (1, 168). Richard Kirwan (1733– 1812) was an Irish eccentric and prominent neptunist, with an implicit belief that all rocks formed from crystallization of minerals in the Earth’s oceans. Maton was to show granitello, along with granite, on his mineralogical map. There are also references in volume 1 to the work of Abraham Gottlob Werner (1749–1817), the founding father of neptunism. A bluish rock found at Ivybridge was equated by Hatchett to a rock type defined by Werner (1, 129) and a calcareous mineral, termed schiefer spar by Werner, was found in one of the tin lodes of the Polgooth Mine (1, 251). Werner’s terminology was again used in defining rocks found in Huel Friendship, near Tavistock in Devon (1, 296/7). The text references to the work of prominent neptunists strongly suggest that Hatchett and, through his influence, Maton looked at the rocks

183

they encountered through neptunist eyes. This is supported by some of the terminology used by Maton, who speaks of searching for a stratum of granite, and by his discussion of the origin of the stone found in quarries near Thorverton, north of Exeter. The origin of this rock puzzled Maton who recognized that a ‘vulcanist perhaps will pronounce it the effect of fusion’ but a ‘cautious mineralogist will set it down among amygdaloidal earths’ (1, 96). A few days later he was walking along the beach at Teignmouth where he noted partly rounded pebbles mostly of an amygdaloidal composition very nearly allied to the Thorverton stone. He ‘could not help fancying that I saw nature in the very act of forming the latter, and bringing together materials for constituting future quarries’ (1, 101). Here he was leaning towards a sedimentary (neptunist) origin for the altered dolerites of Thorverton. Throughout the trip Maton had problems with the terms used by the local miners. He misinterpreted words such as elvan and capel, thinking they were the names of specific rock types rather than words describing the position the rock occupied in relation to a mineral lode or its host rock. He was also confused by the word killas, a miner’s term for the sedimentary rocks of Cornwall, particularly the slaty mudstones and associated coarsergrained rocks. Variations in rock type were characterized by descriptions such as ‘soft shaly killas’ and ‘hard sandy killas’ (Dines 1956) and the terms were used whether or not the sediments were within the metamorphic aureole of the granite. However, Maton associated the term killas with the contact between granite and sediments and his map (Fig. 6) shows the granite, over much of its outcrop, bordered by a relatively narrow zone of killas. Maton makes a number of references to the relationship between mineralogy and botany, particularly in the second volume. During his visit to the Valley of the Rocks (2, 89) he reflected on how lichens in particular could aid the mineralogist (or, more correctly, petrologist). Lichen geographicus and Byssus saxatilis were partial to limestone and a limestone boulder occurring among others could be distinguished immediately. L. caesius and L. cupularis were common only on slate mountains and L. furfuraceus seemed to prefer granite. Visiting the area south of Sherborne he was immediately aware that they had passed into chalk by changes in the vegetation. By the presence of one particular plant, such as Hippocrepis comosa (horse-shoe vetch), the botanist could tell ‘that the soil can be no other that a cretaceous one’ (2, 16). Maton felt that sudden variations in the composition of rocks could often be discovered by merely a glance at their more ‘obvious vegetable inhabitants’ (2, 90),

184

J. D. MATHER

and he can be regarded as one of the first to propose the use of geobotanical methods for geological mapping. At the same time that Maton and Rackett were undertaking their second journey in the summer of 1796, Hatchett independently embarked on a tour through England and Scotland (Raistrick 1967). He published his first scientific paper in the Philosophical Transactions the following year, and this achievement was probably the main reason for his election to Fellowship of the Royal Society on 19 March 1797. Maton was elected three years later on 26 June 1800. According to the citation this was largely on the basis of his authorship of Observations on the Western Counties of England. Both Hatchett and Philip Rashleigh were among his list of proposers. Although the eldest of the trio, Rackett had to wait until 17 February 1803 before he was elected, probably on the basis of his expertise in conchology, when both Hatchett and Maton were among his proposers. Following the publication of his Observations, Maton does not appear to have shown any further interest in mineralogy or geology although he would have known many of the distinguished geologists of the time (in 1832, William Buckland, George Greenough and Roderick Murchison were fellow Royal Society council members). Hatchett pursued his interest in mineralogy and carried out experiments on sulphides from Cornwall (Hatchett 1804a) and on the Bovey Coal (Hatchett 1804b), localities he had first visited in 1794.

determined to see and experience as much as possible on their journeys and must be regarded as among the earliest dedicated geotourists. Maton attempted to summarize the multitude of observations in the form of a mineralogical map for which he used shading with lines to distinguish individual strata. This map is a disappointment as it is over-generalized and the positions of strata on the map often bear no relationship to statements in his journals. No attempt was made to group the rocks into any sort of stratigraphical sequence, not even into the main groups of Primitive, Secondary and Tertiary, which were becoming accepted among scientists by the end of the eighteenth century and which infer relative ages (for further discussion, see Geikie 1905). Maton’s account has been described as the first ‘regional geology’ of any large part of the country (Challinor 1971). Although this seems rather an ambitious claim, it is the first description of SW England written by someone with an interest in science and knowledge of geology and is based on first-hand observations. The fact that it was written at all, and that an attempt was made to distil geological observations into a regional map, marks it out as of considerable historic importance. Sincere thanks to D. Williams of GeoEd in Cornwall who entrusted me with his copy of Maton’s Observations and to my wife J. Bennett for her help with the illustrations.

References Conclusions Although geology was only one of their interests, Maton’s description of the journeys he made in the company of Hatchett and Rackett contains a wealth of geological observations. These are more detailed in the report of the 1794 journey, reflecting the presence of the mineral chemist Hatchett who was absent in 1796. His Observations provide early scientific descriptions of sites and features which were to become well-known tourist attractions in the following centuries, before they were developed or commercialized. For example, his description of Kent’s Cavern was before the major excavations of the eighteenth century, when the accessible extent of the caves was much less than today. They saw Rashleigh’s remarkable grotto and mineral collection, the latter now held by the Royal Cornish Museum with portions at the Natural History Museum, as well as collecting minerals themselves. They clambered down mineshafts, crawled along tunnels, explored caves and visited tin and copper smelters, experiencing the appalling working conditions. They appear to have been

Anon 1819. Report of the Commissioners Appointed for Inquiring into the Mode of Preventing the Forgery of Bank Notes. House of Commons, London. Atwell, J. 1732. Conjectures upon the nature of intermitting and reciprocating springs. Philosophical Transactions, 37, 301– 316. Boud, R. C. 1975. Early development of British geological maps. Imago Mundi, 27, 73–96. Brayshay, M. 1996. Introduction: the development of topographical writing in the South West. In: Brayshay, M. (ed.) Topographical Writers in South-West England. University of Exeter Press, Exeter, 1– 33. Butcher, N. E. 1968. W. G. Maton and the geological map of S.W. England. Proceedings of the Ussher Society, 2, 14. Butcher, N. E. 1983. The advent of colour-printed geological maps in Britain. Proceedings of the Royal Institution of Great Britain, 55, 149– 161. Challinor, J. 1971. The History of British Geology. A Bibliographic Study. David & Charles, Newton Abbot. Chandler, J. 1996. John Leland in the West Country. In: Brayshay, M. (ed.) Topographical Writers in SouthWest England. University of Exeter Press, Exeter, 34–49. Chope, R. P. (ed.) 1918. Early Tours in Devon and Cornwall. Reprinted with an introduction by A. Gibson 1967. David and Charles, Newton Abbot.

TRAVEL JOURNALS OF WILLIAM GEORGE MATON De la Beche, H. T. 1839. Report on the Geology of Cornwall, Devon and West Somerset. Longman, Orme, Brown, Green and Longmans, London. Dewar, H. S. L. (ed.) 1965. The Thomas Rackett Papers. 17th to 19th Centuries. Dorset Record Society Publications, Dorchester, 3. Dines, H. G. 1956. The Metalliferous Mining Region of South-West England. HMSO, London, 2. Geikie, A. 1905. The Founders of Geology. 2nd edn. Reprinted Dover Publications, New York, 1962. Gray, T. 2000a. Cornwall. The Traveller’s Tales, 1. The Mint Press, Exeter. Gray, T. 2000b. Exeter. The Traveller’s Tales, 1. The Mint Press, Exeter. Griffith, W. P. & Morris, P. J. T. 2003. Charles Hatchett FRS (1765– 1847), chemist and discoverer of niobium. Notes and Records of the Royal Society of London, 57, 299–316. Harvey, B. 1970. Who started it all? British Journal of Industrial Medicine, 27, 81– 83. Hatchett, C. 1799. Experiments and observations on shell and bone. Philosophical Transactions of the Royal Society of London, 89, 315–334. Hatchett, C. 1804a. Analysis of a triple sulphuret of lead, antimony and copper, from Cornwall. Philosophical Transactions of the Royal Society of London, 94, 63–69. Hatchett, C. 1804b. Observations on the change of some of the proximate principles of vegetables into bitumen; with analytical experiments on a peculiar substance which is found with the Bovey Coal. Philosophical Transactions of the Royal Society of London, 94, 385–410. Hutchins, J. 1868. The history and antiquities of the county of Dorset: compiled from the best and most ancient historians, inquisitions post mortem, and other valuable records and Mss. Etc. 3rd edn. Corrected, augmented & improved by W. Shipp & J. W. Hodson. J. B. Nichols & Sons, Westminster. Maton, W. G. 1792. Mr Maton’s seal, found at Salisbury. Gentleman’s Magazine, 62, 410. Maton, W. G. 1797a. Observations Relative Chiefly to the Natural History, Picturesque Scenery, and Antiquities of the Western Counties of England, Made in the Years 1794 and 1796, 2 volumes. Easton, Salisbury.

185

Maton, W. G. 1797b. On a species of Tellina, not defined by Linnaeus. Transactions of the Linnean Society, 3, 44–45. Maton, W. G. 1800. Account of the fall of some of the stones of Stonehenge, in a letter from William George Maton, M. B. to Aylmer Bourke Lambert, Esq. F.R.S. and F.A.S. Archaeologica, 13, 103– 106. Maton, W. G. 1815. Case of chorea, in an aged person, cured by musk. Medical Transactions of the Royal College of Physicians of London, 5, 149 –165. Maton, W. G. & Rackett, T. 1804. An historical account of testaceological writers. Transactions of the Linnean Society, 7, 119–244. Mather, J. D. 2013a. William George Maton (1774– 1840) and his mineralogical map of the Western Counties of England. Geoscience in South-West England, 13, 159 –164. Mather, J. D. 2013b. The history and hydrogeology of Laywell, a celebrated ebb and flow spring at Brixham, Devon. Report and Transactions of the Devonshire Association for the Advancement of Science, Literature and the Arts, 145, 133–154. Morris, C. (ed.) 1947. The Journeys of Celia Fiennes. The Cresset Press, London. Paris, J. A. 1838. A Biographical Sketch of the Late William George Maton, M.D. Fellow of the Royal and Antiquarian Societies, Vice-President of the Linnean Society, and Fellow of the Royal College of Physicians. Richard and John Taylor, London. Raistrick, A. (ed.) 1967. The Hatchett Diary: A Tour Through the Counties of England and Scotland in 1796 Visiting their Mines and Manufactories. Bradford Barton, Truro. Rolleston, H. 1942. William George Maton (1774– 1835). Annals of Medical History, 4, 18– 24. Russell, A. 1949. The Wherry mine, Penzance, its history and its mineral productions. Mineralogical Magazine, 28, 517 –533. Torrens, H. S. 2001. Timeless Order: William Smith (1769– 1839) and the search for raw materials. In: Lewis, C. L. E. & Knell, S. J. (eds) The age of the Earth: from 4004 BC to AD 2002. Geological Society, London, Special Publications, 190, 61–83.

The role of Carclaze tin mine in eighteenth and nineteenth century geotourism COLIN BRISTOW Cornwall Geoconservation Group/China Clay Historical Society (e-mail: [email protected]) Abstract: Carclaze tin mine was an open pit operation which exploited a massive cassiteritebearing greisen-bordered quartz-tourmaline vein stockwork, straddling the granite margin near St Austell. It became a ‘must-see’ location for visitors to Cornwall from all over Europe in the late eighteenth and early nineteenth centuries, particularly those interested in the then-fashionable pursuits of geology and mineralogy. Intellectually, the early scientific interest in Carclaze can be seen as part of the Enlightenment, but in the nineteenth century the influence of Romanticism can also be detected. Much of the attraction was due to the openness and accessibility of Carclaze pit, which allowed the geology to be easily appreciated. This resulted in the development of the mine being particularly well documented by a large number of contemporary accounts and illustrations, which has also enabled an early, partly underground, canal to be rediscovered. The earliest account was by the Frenchman M. Jars who visited the pit in 1765; this was followed by accounts by other French geotourists from the early to mid-nineteenth century. The Germans Von Oeynhausen and Von Dechen provided the first geological map and cross-section of the pit in 1829. Accounts by local Cornish authors emphasize that Carclaze was a significant ‘sight’ for visitors. The earliest account of the pit by an English geologist was by Adam Sedgwick in 1822; in later publications he speculated on the formation of parallel vein swarms and schorl rock, partly based on his observations in Carclaze Old Tin Pit. De la Beche provided a pen-and-ink sketch of the south face of the pit in 1839. There are also many accounts by non-scientific visitors throughout the nineteenth century and, together with published lithographs, these are particularly helpful in describing and showing the methods of mining. Tin extraction from the Old Tin Pit had practically ceased by the mid-nineteenth century as production switched to china clay from new pits to the north. The historic south face of the Old Tin Pit, as illustrated by De La Beche, has survived into the twenty-first century and has been designated a County Geology Site by the Cornwall Geoconservation Group, although it is now threatened by housing and industrial development proposals.

One of the most fascinating mining locations in the St Austell area, from both a historical and geological standpoint, is the Old Tin Pit at Carclaze. Partly because of its physical accessibility, it was a ‘must-see’ location for many travellers (some of whom were actual geotourists) to Cornwall in the late eighteenth and nineteenth centuries; this is reflected in the large number of published accounts as well as many paintings and lithographs from the period. Most of the early accounts are scientific and factual and were aimed at a relatively small circle of intellectual readers who were potentially dedicated geotourists (Hose 2008). However, some of the earliest accounts were little more than intelligence reports prepared for foreign governments. Later in the nineteenth century, a romantic element enters into some of the publications (Murray 1851, 1859; White 1855; Leifchild 1857) which were clearly designed to appeal to the burgeoning numbers of discerning tourists. As the number of tourists visiting Cornwall dramatically rose in the twentieth century, mining and geology (as geoheritage) became of growing interest to a significant number of such visitors. Consequently, a number of mining-themed

visitor attractions have been created and there is a substantial body of popular literature devoted to local mining- and geology-related subjects. The St Austell mining district contains a number of large and important disused tin and copper mines; the tin mines include Polgooth, Charlestown United, Wheal Eliza and Carclaze, as well as many other smaller mines (see Fig. 1). In medieval times, the district formed part of the Blackmoor Stannary (Beare 1586) that was mainly concerned with tin streaming the extensive alluvial deposits, some of which became sophisticated mining operations in the eighteenth century, such as Happy Union (Anon/Raspe 1790; Steinmetz & Clarke 2014). Underground tin production in this area had practically ceased by the end of the nineteenth century but alluvial operations in the Goss Moor and Bugle areas, using floating dredges, continued into the first decade of the twentieth century. A substantial account of Cornwall’s nineteenth century tin mining and smelting is provided by Barton (1969). The mining tradition has continued to the present day with the development of the china clay (kaolin) industry. Centred on the St Austell granite, in terms

From: Hose, T. A. (ed.) 2016. Appreciating Physical Landscapes: Three Hundred Years of Geotourism. Geological Society, London, Special Publications, 417, 187–197. First published online February 11, 2015, http://dx.doi.org/10.1144/SP417.11 # 2016 The Author(s). Published by The Geological Society of London. All rights reserved. For permissions: http://www.geolsoc.org.uk/permissions. Publishing disclaimer: www.geolsoc.org.uk/pub_ethics

188

C. BRISTOW

Fig. 1. Map showing part of the Mining District of St Austell. The tin mine names are in italics.

of tonnage and value china clay production dwarfs the whole of the metal production from Cornwall. Since the beginning of the twentieth century it has been one of the main economic drivers for the Cornish economy. Just under one million tons of china clay was produced in 2012. A distinct culture has developed in the ‘clay country’ involving brass bands, choirs and non-conformist chapels.

Situation and geology Carclaze Old Tin Pit is situated about 3 km NNE of St Austell at an altitude of 210 m, in what would originally have been rather bleak moorland with a dramatic view over St Austell Bay. Worked as an open pit, this tin mine much impressed late-eighteenth- and early-nineteenth-century visitors due to its size; however, it is now dwarfed by

CARCLAZE TIN MINE: GEOTOURIST DESTINATION

the large pits of the present-day china clay industry. Water played an important part in the mining operation, but the supply for power and mineral processing was a major challenge due to the situation of the mine on the watershed. This was partially overcome by the construction of a series of ponds to trap winter rains and the building of a lengthy water supply leat (Bristow 2011). The geology of Carclaze Old Tin Pit has been described by Bristow (2008) and Collins (1878). This tin occurrence lies adjacent to the southern boundary of the St Austell granite and the open pit exploited a massive stockwork composed of a network of cassiterite-bearing greisenbordered quartz-tourmaline veins in kaolinized granite, generally dipping steeply south parallel to the granite/aureole contact, which lies along the top of the southern face of the pit. Early accounts suggest there was a massive development of schorl rock, now quarried away, perhaps accompanied by the development of tourmaline breccia.

Early pit development Tin extraction in the Carclaze area is reputed to date back to Tudor times (Drew & Hitchens 1824; Collins 1912), although tin streaming in nearby valleys probably dates back to Bronze Age times (Penhallurick 1986). Gilbert (1817) suggested that the earliest working was by means of an underground mine. This was followed by the development of an open pit where the ore was shovelled up to the surface by the labour-intensive process of ‘shambling’ (Carne 1822). The earliest first-hand account of a visit to the pit is by the French mining engineer Jars who visited in 1764–1765 (Jars 1781). He described how the pit was being worked and reported that the ore-bearing veinstone was dropped through a hatch in the bottom of the pit, into a chain of tub boats floating in an underground canal beneath. The tub boats were then pulled out of the underground canal and conveyed in an above-ground canal along the contour to a point where the barges could be discharged by a crane into carts. These carts took the ore down to the bottom of the St Austell river valley where there was sufficient water power to drive the stamps which pulverized the ore; by this means the cassiterite was liberated and recovered. The softer kaolinized granite was washed away as a slurry through a second, gently inclined, adit. Later in the 1760s, a deeper canal was constructed which allowed the pit to be deepened and the original upper canal then became a leat conveying water into the pit. The supply of water was augmented by being connected by a leat to a series of springs on the south side of Hensbarrow, 3.5 km away to the northwest. This allowed the stamps to

189

be moved into the pit and Bonnard reported in 1803 that ‘the great excavation’ was 60 fathoms long, 30 wide and 20 deep, with no less than ten sets of stamps in operation. Further details of the mining technology can be found in Bristow (2011). Mining techniques developed in Carclaze Old Tin Pit in the eighteenth century later proved invaluable to the early development of the china clay industry in the nineteenth century.

Visiting Carclaze in the eighteenth and nineteenth centuries Visiting Cornwall in the early eighteenth century and earlier was difficult as it was a remote area, especially in terms of transport infrastructure, from the rest of England. The most practical way of travelling to Cornwall then was by coastal voyage; movement by sea was aided by the numerous navigable estuaries around the Cornish coast leading into the interior. Land-based transport within Cornwall was severely hampered by a distinct lack of roads. A good description of the problems facing late seventeenth century road travellers in Cornwall is found in Celia Fiennes’ ‘Through England on a side saddle’ (1695). Kanefsky (1999) reported that there were few roads serving the long-distance traveller in the early eighteenth century, and those that did exist were little more than bridleways. Pack horses were used for the transport of goods and riding horses for long-distance journeys. The visitor therefore had to be prepared to surmount these difficulties of getting around Cornwall which, for many English visitors, seemed like the ‘Wild West’. This is emphasized by Howard’s (1999) article on early tourist destinations. His map of the locations of south-western landscape paintings from 1800–1865 shows few west of the Tamar, except around Mount’s Bay which could easily be accessed by sea. Because of the lack of roads there were very few wheeled vehicles in the early eighteenth century. The roads improved after 1750 when various Acts set up turnpike trusts, and horse-drawn ‘vans’ provided a form of public transport between the main towns. Further improvements took place in the nineteenth century, culminating in the main railway line through Cornwall which opened in 1859. However, as late as 1863, Bradshaw’s Descriptive Railway Hand-Book (section 2, p. 34) stated that Cornwall was one of the least inviting of English counties to visit and describes it as ‘A ridge of bare and rugged hills, intermixed with bleak moors runs through its whole length, and exhibits the appearance of a dreary waste. The most important objects in the history of the county are its numerous mines . . .’. In his book on West Country Railway

190

C. BRISTOW

History, Thomas (1973) relates that before 1890 most traffic on the Cornish railways was related to mining and agriculture/fishing (mainly for the London markets); rail tourists to Cornwall were remarkably few because the county initially had no established resorts. Passenger numbers to Cornwall rapidly increased after 1890 with faster trains, enhanced timetables and cheaper fares, especially following promotion by the Great Western Railway (Wilson 1987).

Early descriptions of Carclaze by foreign travellers The earliest known account of the pit is by Gabriel Jars and is contained in Voyages Me´tallurgiques, published posthumously in three volumes by his brother Antoine Gabriel in 1774–1781. Jars visited many mining regions in Europe and reported on them. He appears to have visited Cornwall in 1764 or 1765 and places visited include streamworks around St Austell, Pednandrea mine in Redruth and Godolphin mine near Helston. Carclaze was not named although his description leaves little doubt that it was Carclaze; the paper contains a detailed description of the pit and the manner of working as well as the partly underground canal, which corresponds with later descriptions by Gilbert (1817) and Drew & Hitchens (1824). Jars only mentions the production of tin ore from Carclaze, as this was before china clay production from the St Austell deposits had become significant. At first sight it would seem that Jars was an innocent scientific tourist but, on closer examination, we find that during his journeys around Europe Jars was actually being paid by the French state; on his return he was appointed Inspecteur Ge´ne´ral des Manufactures. He was appointed correspondent of L’Acade´mie des Sciences in 1761 (before his visit to England) and became a full member in 1768. Rather than an innocent tourist, perhaps he should be regarded as an early form of what would now be called an ‘industrial spy’. Jars’ visit in 1765 was followed by a visit in 1803 by another Frenchman, A.H. Bonnard, who later became Inspecteur Ge´ne´ral des Mines in Paris. Bonnard’s account, published in the Journal des Mines, contains a detailed description of the mining and mineral processing techniques in use at Carclaze; this appears to have been yet another information-gathering exercise on behalf of the French State. Whether the local people in St Austell realized what was going on is not known. However, without these accounts by Jars and Bonnard we would have little knowledge of the early development of Carclaze Old Tin Pit in the late eighteenth century together with its associated mineral processing facility, as there are

no accounts by British writers dating back to this period. In 1814 the landscape painter Joseph Farington produced a lithograph of ‘Curclaze tin mine’ (Fig. 2) which formed part of ‘Views in Cornwall’ intended for the revision of Britannia Depicta by T. Cadell and W. Davies. This is the earliest published image of the mine and it shows the pit facing westwards with a stream of water gushing from an adit in the south face, powering three sets of stamps complete with thatched roofs. No doubt in 1814 this must have been a dramatic and unusual scene, forming part of a unique industrial landscape. The next publication of note is Gilbert’s 1817 substantial two-volume Historical Survey of the County of Cornwall; a whole page (p. 248) in this refers to ‘Crelaize’ open mine and provides a helpful description of the manner of operation of the canal and the chains of tub boats.

Some famous early-nineteenth-century geologists’ and mining engineers’ visits to Carclaze In 1822 Adam Sedgwick paid the first of several visits to ‘Carglaze’; he appears to have been impressed by being able to visit and see the geology in an open pit and recorded: ‘The traveller may there see the operation of the miner carried on in the light of day, without being compelled to descend a hundred fathoms below the surface of the earth and then to crawl into a dirty dripping cavern’. It was this facility to see a tin mine in the light of day that made Carclaze such an effective visitor attraction. Sedgwick appears to have again visited the St Austell granite in the company of Professor Whewell in 1828 and, in his retiring address as President of the Geological Society (1831), he refers to the schorl masses as ‘veins of segregation’. In a paper in the Transactions of the Geological Society in 1835, he returned to the same theme and was clearly puzzled by the relationship of the schorl rock to the granite and the veins. It is a debate that still continues to this day. The notable Cornish antiquary and mineralogist Philip Rashleigh maintained a lively correspondence with other scientists and antiquarians, which included a description of Carclaze in 1804. With the establishment of the Royal Geological Society of Cornwall in 1814 Carclaze is mentioned briefly in the Transactions by Carne and Hawkins in 1822, although the scientific content of their observations is minimal. A more substantial description of Carclaze is found in the substantial two-volume History of Cornwall by Drew & Hitchens, published in 1824 (in volume 2, pp. 61 –62). The pit was then described as covering 12 acres (4.8 ha) and was

CARCLAZE TIN MINE: GEOTOURIST DESTINATION

191

Fig. 2. Farington’s (1814) lithograph of Curclaze tin mine looking west, showing water being discharged from an adit on the left and three sets of waterwheel driven stamps. Published with permission from an original print in Cornwall Records Office, Truro.

referred to as ‘an astonishing excavation’ of ‘singular form’. China clay production was by then becoming significant and tin production was declining in importance. French authors now re-enter the fray. Mining engineers Dufre´noy and de Beaumont produced three papers (in 1824, 1826 and 1827) which provide detailed factual descriptions of the geology and working methods at ‘Carclase’. Two German authors, Von Oeynhausen and Von Dechen, published a paper in 1829 on the junction of the granite with the killas in various parts of Cornwall; they included the first map and section of Carclaze tin pit. Later, the Frenchman Daubre´e produced two papers in 1841. The first was published in Annales des Mines (Daubre´e 1841a) and is on the origin of tin deposits with a description of the mineralogy of Carclaze. The second was in the Bulletin Societe´ Geologique de France (Daubre´e 1841b) and has some mentions of Carclaze within a discussion on the geochemistry of the tin-deposit-forming processes. These various Franco-German papers all

appear to be simply concerned with reporting on the geology, mining methods and geochemistry, and do not appear to be catering in any specific way for the casual (Hose 2008) geotourist. As the author has pointed out (Bristow 2008) however, it is clear that the approach of these continental authors was far more professional than the more dilettante Cornish ‘gentleman geologists’ (or perhaps dedicated geotourists) such as Carne and Hawkins. This deficiency was partially remedied in the first Memoir to be published by what was to become the Geological Survey: De la Beche’s 1839 Report on the geology of Cornwall, Devon and West Somerset. This contains a detailed description of the geology of ‘Carglaze’ on pages 164 and 346–347, as well as a pen-and-ink sketch of the south face of the pit, probably by the great man himself (Fig. 3). The deficiency was further remedied by Smith’s detailed account in the Gentlemen’s Magazine in 1840 of the mineral processing methods in use at Carclaze, and a paper by the Frenchman Moissenet published in Annales des Mines in 1858

192

C. BRISTOW

Fig. 3. Pen-and-ink sketch of ‘Carglaze’ pit from De la Beche (1839), showing the south face of the pit. From an original copy in the author’s possession.

provides interesting data on tin recovery at Carclaze and the costs involved. A lithograph by Thomas Allom published in 1831 (p. 35), and paintings by Mitchell in 1841 (Fig. 4) and Smyth in 1845 (watercolour of Carclaze mine in the collection of the Royal Institution of Cornwall), provide superb and detailed images of the pit. The fact that these artists chose to depict Carclaze pit emphasizes the perception that Carclaze was considered one of the ‘sights’ in Cornwall at the time the artists were at work. Before the railway network was established, these images would also no doubt have encouraged the few tourists reaching Cornwall to visit Carclaze.

The arrival of the railway and modern tourism The national railway network reached Cornwall in 1859, making Cornwall far more accessible. More general and less technical accounts published around this time were clearly aimed at potential tourists rather than intellectuals although, as pointed out above, the numbers visiting the county before 1890 were low compared to the present day. Notable publications which mention Carclaze include Bradshaw’s Descriptive Railway Hand-Book in 1863

(section 2, p 34), Murray’s Handbook for Devon and Cornwall published in 1859, and Walter White’s (1855) A Londoners walk to the Land’s End and a trip to the Scilly Isles. There are also more mining-specific books such as Leifchild’s (1857) Cornwall: its mines and miners. Murray wrote: ‘It requires, indeed, no great stretch of the imagination to fancy Carclaze a work of enchantment, and a chasm which has been opened by some potent magician in a mountain of silver’. White waxes equally lyrical: ‘The pale grey rock, almost white in places, and full of crystalline grains, it sparkles everywhere in the sunlight, and with such brightness, that a stranger might fancy to be of the richest metal . . .’. Leifchild summarizes the attraction of Cornwall to the visitor: ‘Although the chief subject of the book is mining, yet the antiquities, natural and national, lingual, local and mineral; the wild Cornish granite moors, with their weird scenery, and innumerable blocks and pillars; the bold coasts, and high bluff headlands; the boundless ocean, . . . will prove as interesting to the reader as they have been to the writer’. It is now clear that we are entering an era in which the tourist needs to be encouraged by some purple prose! A field excursion report by Michell et al. concerning a trip organized by the Miners’ Association of Cornwall and Devon in 1870 and an excursion to

CARCLAZE TIN MINE: GEOTOURIST DESTINATION

193

Fig. 4. Philip Michell’s 1841 lithograph of Carclaze, looking east, showing four sets of stamps driven by water issuing from adits on the right. Published with permission from an original coloured print in the collection of the Royal Institution of Cornwall, Truro,

Cornwall made by the Geologist’s Association in 1887 (Thomas 1889) show that Carclaze was becoming a favoured locality for geotourists coming to Cornwall in the late nineteenth century. Important papers by Symons (1878– 1881) and Collins (1878, 1892) in the final decades of the nineteenth century complete the roll of scientific papers on Carclaze, after which the Old Tin Pit was abandoned due to falling tin yields. China clay production from the Baal pits to the north of the Old Tin Pit took over as the main source of economic activity at Carclaze. Thereafter, any scientific and technical publications mentioning Carclaze simply reiterate information from earlier papers.

Post-industrial interest in the mining landscape In the first half of the twentieth century a number of excellent technical reviews of the china clay industry (Collins 1912; Howe 1914, p. 25) were published, but there is little evidence of tourist interest in Carclaze or, for that matter, the very extensive

network of china clay workings in the St Austell granite. This seems to reflect a perception at that time that this was a purely industrial area which had little to offer the tourist. However, after the Second World War, writers, poets and artists began to take an interest in the china clay area. This interest has steadily gathered pace, such that there is now a veritable torrent of art and literature of varying quality being produced. The cultural traditions of the area, centred on choirs, brass bands and non-conformist chapels, have also become a matter of considerable public interest. The Wheal Martyn China Clay Museum was established at Carthew in 1974 and has become a focus for interest in the china clay industry and a significant tourist attraction. Landscaping of the disused pits and tips has created unique and biodiverse areas of heathland and woodland (Bristow 2006, pp. 35 –43). With its futuristic biomes, the Eden Project was built in a former china clay pit at Bodelva in 2000 and has become a major tourist attraction and a focus for plant conservation. The China Clay History Society was established in 2001 and is actively concerned with maintaining a large archive of documents,

194

C. BRISTOW

photographs and other artefacts concerned with the industry. It has refrained from active campaigning to conserve sites of geological or industrial archaeological significance however, partly because its archive is sited in a building provided by the main china-clay-producing company Imerys. Elsewhere in Cornwall and Devon, important tourist attractions have been created around the remains of former metalliferous mines such as Geevor/Levant near St Just, the Heartlands Centre in Redruth, King Edward Mine near Camborne, Poldark Mine near Helston and Morwellham and the Tamar Valley trails on the banks of the River Tamar. Walking and cycling trails have been established throughout the china clay area, as the postindustrial landscape of worked-out flooded pits and tips is re-instated and made available for public access. This has created a dramatic and unusual landscape which has considerable visitor appeal. A disused railway line, formerly used to convey china clay from St Austell to the small port of Pentewan, is a particularly popular trail much used by local people and visitors, with associated facilities such as cycle hire. Another railway line from

St Austell station to Wheal Martyn is also a popular trail. The Cornwall and West Devon Mining Landscape World Heritage Site was set up in 2006 and further emphasizes the historical importance of mining in Cornwall, although Carclaze Old Tin Pit did not form part of the World Heritage bid. One problem which Carclaze faces in this context is the fact that the World Heritage team has been largely composed of archaeologists who are mainly interested in structures such as beam engine houses and other built structures, and not in the geology. While the archaeological team associated with the World Heritage Site is quite generously funded, geoconservation has practically no funding. Much the same applies in a comparison with wildlife funding. Carclaze Old Tin Pit was largely forgotten until the 1990s, when a paper by the author (2008) set out its historical significance. The associated, partly underground, canal was the earliest canal in Cornwall and might be the earliest in Britain to use chains of tub boats (Bristow 2011). A County Geology Site (formerly a RIGS site) was set up in

Fig. 5. The present-day south face of Carclaze Old Tin Pit, now designated as a County Geology Site. This is probably the same face as that sketched by De la Beche in the 1830s (see Fig. 3.) Photo by the author 2011.

CARCLAZE TIN MINE: GEOTOURIST DESTINATION

January 2009 to try to preserve the historic south face of Carclaze Old Tin Pit (Fig. 5), as sketched by De la Beche in 1839 (Fig. 3). Most of the china clay industry, including Carclaze, is now in the hands of the French company Imerys which, in view of the considerable French interest in the site in the late eighteenth and early nineteenth centuries, is rather appropriate. However, there have been conflicting development proposals for the area of the pit, including an ‘Ecotown’ proposed by Imerys/Ecobos and a large industrial building which has just been erected nearby, that greatly detracts from the general historic character of the area. Although several walking and cycling trails pass around the perimeter of Carclaze/Baal pits, it is currently not possible to descend into the pit to view the historic south face. Imerys/Ecobos control entry to the site and it is difficult to obtain permission to visit; this is largely because of health and safety considerations that do not permit visitors to freely wander around an abandoned and flooded china clay working. What the future will hold for this historic site is anybody’s guess.

Conclusion This study shows how the tourist (in the broadest sense) visiting Carclaze evolved from the late eighteenth century up to the twentieth century, reflecting the change in their motivation from intellectual enquiry to tourism as a leisure activity. In summary: (1) Carclaze Old Tin Pit attracted many visitors in the eighteenth and nineteenth centuries, mainly because it was an open pit in an accessible position where the geology and mining methods could be easily appreciated, especially by casual geotourists. (2) Early French visitors in the eighteenth and early nineteenth centuries, sponsored by the French government, were principally interested in learning about the innovative mining and mineral processing methods in use at Carclaze. (3) Early scientific accounts of the mine’s geology in the first half of the nineteenth century, aimed at an intellectual readership, were provided by some of the leading British geologists of the day such as Sedgwick and De la Beche, as well as by leading French and German investigators. (4) The flow of both dedicated and casual geotourists in the eighteenth and early nineteenth centuries was restricted by difficult travelling conditions within Cornwall. Nevertheless, the number of lithographs and paintings

(5)

(6)

(7)

195

from this time suggest that Carclaze was one of the recognized ‘sights’ and a locality which should be visited by the more determined traveller. In the late nineteenth century travelling conditions improved with the arrival of the railway in 1859, and Carclaze became an established site on the tourist trail around Cornwall. Geological descriptions in a number of guide books and itineraries indicate that geology was perceived as an important part of natural history. Some of the accounts include ‘romantic’ descriptions of the pit. Through most of the twentieth century little interest was shown in the Old Tin Pit, but in the last ten years there has been much interest in the pit and its associated canal. The historic south face, as sketched by De la Beche, is still extant and has been designated a County Geology Site (¼RIGS). However, access is restricted at present due to health and safety considerations and there are conflicting development plans for the area. The future of the pit as a site for geotourism is still very uncertain, despite increasing interest in the history and environment of the china clay area in general. A particular problem is the lack of funding for geoconservation in comparison with the wildlife and archaeological sectors and the small number of people interested in geoconservation. This disparity is in spite of tourism being the most important ‘industry’ in Cornwall; most tourists are attracted to Cornwall because of its landscape and scenery which are, of course, underpinned by the geology.

I thank the anonymous referee and the Volume Editor, T. Hose, for both constructive comments and suggesting additional sources that were most helpful in finalizing the paper. W. Cawthorne at the Geological Society Library and the staff of the Cornish Studies Library at Redruth are thanked for their assistance in tracking down some of the more obscure references. The assistance given by C. Thurlow in suggesting some of the sources is also gratefully acknowledged.

References Allom, T. 1831. Engraving of ‘Carclase tin mine, near St. Austle’. In: Cornwall Illustrated in a Series of Views. Fisher & Son, London, opposite 35. Anon (in the style of Raspe, R.E.) 1790. Ausgang aus bem Reifejournale eines Deutfchen. Reife von London in die Grafschaft Kornwall. Bergmannifches Journal, Dritter Jahrgang, Zwenter Band. Siebendes Stu¨ck. Julius, 141–153. Barton, D. B. 1969. A History of Tin Mining and Smelting in Cornwall, 2nd edn. Cornwall Books, Exeter.

196

C. BRISTOW

Beare, T. 1586. The Bailiff of Blackmoor. [Transcribed and edited by Buckley, J.A. in 1994]. Penhellenick Publications, Camborne. Bonnard, A. H. 1803. Sur le gisement, l’explotation et le traitement de l’E˙tain, dans le duche´ de Cornouailles. Journal des Mines, 14, 443– 454. Bradshaw, G. 1863. Bradshaw’s Descriptive Railway Hand-Book of Great Britain and Ireland. Henry Blacklock & Co. Ltd., London. Bristow, C. M. 2006. China Clay – A Geologist’s View. Cornish Hillside Publications, St Austell. Bristow, C. M. 2008. Late 18th and early 19th century forays into economic geology – some little known Franco-German papers describing Carclaze Old Tin Pit, near St. Austell, Cornwall. Geoscience in SouthWest England, 12, 1 –8. Bristow, C. M. 2011. The rediscovery of the 18th century Carclaze-Scredda canal system near St. Austell, Cornwall. Journal of the Trevithick Society, 38, 3– 23. Carne, J. 1822. On the mineral production and the geology of the Parish of St. Just. Transactions of the Royal Geological Society of Cornwall, 2, 345. Collins, J. H. 1878. The Hensbarrow Granite District. Lake & Lake, Truro, 36–37. Collins, J. H. 1892. On the origin and development of ore deposits in the west of England (second part). Journal of the Royal Institution of Cornwall, 11, 126. Collins, J. H. 1912. Observations on the west of England mining region. Transactions of the Royal Geological Society of Cornwall, 14, 59–60 & 437. Daubre´e, M. 1841a. Sur le gisement, la constitution et l’origine des amass de minerai d’e´tain. Annales des Mines, Troisieme Se´ries, Tome, 20, 90–91. Daubre´e, M. 1841b. Extrait d’un me´moire sur le gisement, la consititution et l’origene des amass de minerai d’e´tain. Bulletin De La Societe Geologique De France, 12, 393– 401. De la Beche, H. T. 1839. Memoir of the Geological Survey: Report on the geology of Cornwall, Devon and West Somerset. Longman, Orme, Brown, Green & Longmans, London. Drew, S. & Hitchens, F. 1824. History of Cornwall. William Penaluna, London, two volumes. Dufre˙noy, P. A. & de Beaumont, E. 1824. Sur le gisement, l’exploitation et le traitement des minerais d’e´tain et de cuivre du Cornouailles. Annales des Mines, 9, 905– 908. Dufre˙noy, P. A. & de Beaumont, E. 1826. Sur la constitution ge´ognostique et les gites me´tallife˙res du Cornouailles et du Devonshire. Annales des Sciences Naturelles, 7, 220–221. Dufre˙noy, P. A. & de Beaumont, E. 1827. Voyage me´tallurgique en Angleterre, ou recueil de me´moire sur le gisement, l’exploitation et le traitement des minerais d’e´tain et de cuivre, de plomb, de zinc, et de fer, dans le Grande Bretagne, etc. Bachelier, Paris. Farington, J. 1814. Lithograph of ‘Curclaze Tin Mine’. ‘Views in Cornwall’, chapter. In: Cadell, T. & Davies, W. (revs) Britannia Depicta. 4 edn. Fiennes, C. 1695. Through England on a Side Saddle. Leadenhall Press, London (1888). Gilbert, G. S. 1817. An Historical Survey of the County of Cornwall. Longman, Hurst, Orme and Brown, London, 2 vols.

Hawkins, J. 1822. On the stratified deposits of tin-stone called tin-floors, and on the diffusion of tin-stone through the mass of some primitive rocks. Transactions of the Royal Geological Society of Cornwall, 2, 31. Hose, T. A. 2008. Towards a history of geotourism: definitions, antecedents and the future. In: Burek, C. V. & Prosser, C. D. (eds) The History of Geoconservation. The Geological Society, London, Special Publications, 300, 37– 60. Howard, P. 1999. Chapter 56, ‘Early tourist destinations, the influence of artists’ changing landscape preferences’. In: Kain, R. & Ravenhill, W. (eds) Historical Atlas of South-west England. University of Exeter Press, Exeter, 450– 452. Howe, J. A. 1914. A Handbook to the Collection of Kaolin, China Clay and China Stone in the Museum of Practical Geology. HMSO, London. Jars, G. 1781. Voyages me´tallurgiques, ou recherches et observations sur les mines de cuivre, celles de calamine, & la fabrication du laiton; les mines d’e´tain, les monnoies; les mines et fabriques d’alun; celles de soufre & de vitriol; les mines de sel, & les salines, les poteries, les pipes, les briques & les tuiles, faites en 1758, 1765, jusques et compris 1769, en Allemagne, en Sue˙de, Angleterre, Norve˙ge, Tirol, Liege, & en Hollande. P.F. DIDOT, L. CELLOT AND L.A. JOMBERT, PARIS, Volume III. Kanefsky, J. 1999. Chapter 45, ‘Turnpike Roads’. In: Kain, R. & Ravenhill, W. (eds) Historical Atlas of South-west England. University of Exeter Press, Exeter, 357–363. Leifchild, J. R. 1857. Cornwall: Its Mines and Miners. Longman, Brown, Green, Longmans and Roberts, London. Michell, S., Argall, W. H. & Eudey, J. 1870. Account of an excursion to Carclaze Pit on 2nd August 1870. Report of Miners’ Association of Cornwall and Devon Moissenet, L. 1858. Excursion dans le Cornwall 1857. Pre´paration Me´chanique du Minerai d’e´tain. Annales des Mines, 14, 77– 216 (partially translated from the French by A.J. Clarke and published by the Trevithick Society, 2010). Murray, J. 1851. Handbook for Travellers in Devon and Cornwall. John Murray, London. Murray, J. 1859. Handbook for Travellers in Devon and Cornwall. John Murray, London (available as a 1971 reprint by David & Charles, Newton Abbot). Penhallurick, R. D. 1986. Tin in Antiquity. The Institute of Metals, London. Sedgwick, A. 1822. On the geology of Cornwall, etc. Transactions of the Cambridge Philosophical Society, 1, 108. Sedgwick, A. 1831. Address to the Geological Society delivered on the evening of the 18th of February by the Rev. Professor Sedgwick, MA, FRS on retiring from the President’s chair 1831. Proceedings of the Geological Society, 1, 281–316. Sedgwick, A. 1835. Remarks on the structure of large mineral masses, and especially on the chemical changes produced in the aggregation of stratified rocks during different periods after their deposition. Transactions of the Geological Society of London, 3, 3rd Series, 461–486.

CARCLAZE TIN MINE: GEOTOURIST DESTINATION Smith, B. P. 1840. Trip to the Far West. The Gentleman’s Magazine, and Historical Chronicle, 167, 164–173. Steinmetz, E. & Clarke, A. 2014. An anonymous late 18th century account of the Happy Union Alluvial Tin Mining operation near Pentuan; in the style of Rudolph Erich Raspe. Translated by Steinmetz, E. & Clarke, T. with an introduction by Bristow, C. Journal of the Trevithick Society, 41, 73– 95. Symons, R. 1878– 1881. On Carclaze tin and china-clay pit. Journal of the Royal Institution of Cornwall, 6, 140–143.

197

Thomas, D. St J. 1973. West Country Railway History. David & Charles, Newton Abbot. Thomas, W. 1889. Report on an excursion to Cornwall, August 8th to 13th 1887. Proceedings of the Geologist’s Association, 10, 198–199. Von Oeynhausen, C. & Von Dechen, H. 1829. On the junction of the granite and the killas rocks of Cornwall. Philosophical Magazine, Second Series, 5, 161– 170 & 241– 247. White, W. 1855. A Londoners Walk to the Land’s End and a Trip to the Scilly Isles. Chapman & Hall, London. Wilson, R. B. 1987. Go Great Western: A History of GWR Publicity, 2nd edn. David & Charles, Newton Abbot.

From tourism to geotourism: a few historical cases from the French Alpine foreland NATHALIE CAYLA*, CHRISTOPHE GAUCHON & FABIEN HOBLE´A Laboratoire EDYTEM UMR CNRS-Universite´ de Savoie, Poˆle Montagne, Technolac, FR – 73 376 Le Bourget du Lac, Savoie, France *Corresponding author (e-mail: [email protected]) Abstract: This paper traces the touristic trajectories of three spectacular gorges located in the Alpine foreland and the southern Jura: the gorges of the upper Rhoˆne (Ain/Haute-Savoie), the Sierroz (Savoie) and the Fier (Haute-Savoie). All three are located within a distance of 50 km from each other. The upper Rhoˆne gorge, already famous at the end of the eighteenth century, was drowned under the floodwaters of the Ge´nissiat dam in 1948; only a significant iconography remains of two centuries of (geo)tourism. The Sierroz gorge, close to the spa resort of Aix-lesBains, became famous after the dramatic and tragic death in 1810 of a young noblewoman. Following that event many tourists staying on the shore of the lake Bourget visited the gorge until 1970 when it was closed to the public. Since then, the gorge has gradually become a touristic wasteland. The Fier gorge near Annecy became a tourist attraction in 1869 with the opening of the nearby railway station of Lovagny; since then, visitors have been attracted to it in increasing numbers. The history of these three gorges illustrates how tourism and heritage are in constant interaction; however, the development of the one will not always ensure the protection of the other. Today, geoheritage assessment is based upon criteria that are as objective as is possible. The intrinsic geological and geomorphological characteristics are the initial geoheritage values, to which can be added the cultural value elements. Associated with the development of geotourism and geoparks, this new approach should ensure a better and sustainable use of these sites in the long term.

In 1741, English traveller William Windham (1717– 1761) discovered the Mer de Glace during his extended visit to Chamonix during 1741 –1742. Windham had embarked on an extensive Grand Tour of Europe in 1737, living in Geneva from 1740 to 1741. In Geneva he joined a group of British expatriates known as the ‘The Common Room’ and, in the company of some of its members, he explored the area around Chamonix in 1741 (Rowlinson 1998); seemingly they were the region’s first recorded pleasure travellers, the term ‘tourist’ not being in popular use at that time. Having scaled Montenvers with the aid of local guides, they christened the glacier below the Mer de Glace. Returning to England in 1742 Windham published a pamphlet (Windham & Martel 1744) about his Chamonix exploits in which he described the glaciers as: ‘You must imagine your lake put in agitation by a strong wind, and frozen all at once, perhaps even that would not produce the same appearance’. From that time, traveller’s guide books, alpine literature and all the associated iconography has attracted an increasing number of tourists to the Mont-Blanc massif (Reichler 2002; Deline et al. 2012). Beyond the valley of Chamonix, the northern French Alps offer many sites and landscapes whose picturesque attraction first opened the way for their tourism development

and then their evolution towards geotourism practices; other sites fell into oblivion however (Hoble´a 2014). A corpus of nearly 400 geosites (museums, caves, gorges, quarries, mines and geotrails) was inventoried in 2009 across the six alpine countries (Cayla 2009), highlighting the importance of these areas. The 25 gorges and canyons included are successful destinations for tourists, such as the Breitach gorge (with 300 000 annual visitors) in Germany, the Aar gorge (with 120 000 annual visitors) in Switzerland and the Bletterbach canyon (with 50 000 annual visitors) in Italy. Tourism is a key economic sector for these countries, with over 300 million tourist nights spent in the Alps every year (CIPRA, www.cipra.org); further, about five million visitors discover a geosite during their stay (Cayla et al. 2009). This paper traces the touristic trajectories of three spectacular gorges located in the Alpine foreland and the southern Jura: the gorges of the upper Rhoˆne (Ain/Haute-Savoie), the Sierroz (Savoie) and the Fier (Haute-Savoie). All three are located within a distance of 50 km from each other (Fig. 1). These geosites and geomorphosites were selected because of their geoscientific interest and their additional values that have contributed to their tourism development (Panizza & Piacente

From: Hose, T. A. (ed.) 2016. Appreciating Physical Landscapes: Three Hundred Years of Geotourism. Geological Society, London, Special Publications, 417, 199–213. First published online February 11, 2015, http://dx.doi.org/10.1144/SP417.10 # 2016 The Author(s). Published by The Geological Society of London. All rights reserved. For permissions: http://www.geolsoc.org.uk/permissions. Publishing disclaimer: www.geolsoc.org.uk/pub_ethics

200

N. CAYLA ET AL.

Fig. 1. Map of the study area.

2003; Reynard 2009; Duval-Massaloux & Gauchon 2010). They have all experienced a ‘heritagecreation itinerary’ (‘itine´raire de patrimonialisation’) (Gauchon 2002) or a ‘heritage trajectory’ (‘trajectoire patrimoniale’) (Portal 2010) with different results (Panizza 2001). The history of the tourist practices of these three sites are examined in this paper in order to facilitate an understanding of the interactions between tourism and heritage through time and the history of the usages of these sites. The three gorges were deeply carved by rivers through middle Cretaceous (Barremian– Aptian) limestone after the last marine intrusion at the western margin of the Alps. Their excavation can be dated to around 10 Ma during the Mio-Pliocene, after the deposit of the last molasses (superimposed) and presumably controlled by the Tortonian phase of Alpine folding (antecedent), with reactivations of subglacial and/or pro-glacial character during the Pleistocene glacial cycles. This limestone constitutes an important mass with a varying thickness up to 300 m in the subalpine massifs. It is witness to an ancient carbonate platform that bordered the northern Tethyan margin around 107 –114 Ma.

This palaeo-environment led to a reef facies called the ‘Urgonian’, a pure and compact limestone. The three gorges studied are located in the pre-Alps at an altitude of less than 400 m. They incise an anticlinal ridge (as do the Fier and Sierroz rivers) or the bottom of a syncline depression (as has the Rhoˆne River) formed by Urgonian limestone folded during Late Cretaceous time.

The ‘loss of the upper Rhoˆne’: a geoheritage trajectory abruptly interrupted At the Franco-Swiss border 30 km SW of Geneva, near the town of Bellegarde-sur-Valserine in the department of the Ain, a famous geotouristic site disappeared in 1948. The loss of the upper Rhoˆne, already well known at the end of the eighteenth century, was due to the disappearance of the Rhoˆne at low water to the bottom of a deep gorge, subsequently drowned in its entirety under the floodwaters of the Ge´nissiat dam. Nowadays, only a significant iconography remains of these two centuries of tourism. There, the Rhoˆne rushed into a narrow canyon carved in the Urgonian limestone

FROM TOURISM TO GEOTOURISM: FRENCH ALPINE FORELAND

and, during the period of lower water, completely disappeared from view for a few hundred metres in length. The ‘Loss of the upper Rhoˆne’ appellation given to this picturesque site was, strictly speaking, a misnomer because the water did not penetrate the rock, but was just hidden at the bottom of the gorge by the huge scree boulders. Apart from the loss of water in the abyss of Malpertuis, the confluence of the Rhoˆne and the Valserine rivers a few hundred metres downstream completed a tourist’s visit which was characterized by many as quite spectacular. It attracted many visitors on the way to Geneva in the eighteenth century. Horace Benedict de Saussure (Saussure 1779) describes it carefully in the first volume of his Voyages dans les Alpes. He suggested the best period to make a visit was during the spring or the summer when the Rhoˆne totally disappeared; he even indicated the best point of view as on the bridge of Lucey, which no longer exists. Travellers’ tales multiplied during the nineteenth century and the site, appearing in many travel guides, gained some repute. Albanis Beaumont (1755–1812) recounted his trip in his book, written in English, published in 1800 (Albanis Beaumont 1800). One of the illustrations (see Fig. 2) shows the technique used at that time to descend

201

into the gorge. On the centre of the picture a visitor climbs a ladder maintained by a guide. A few years later, the ‘Loss of the upper Rhoˆne’ is prominently noted in Volume 2 of the monumental work edited by Isidore Taylor, Voyages pittoresques et romantiques dans l’ancienne France in the tome on the Franche-Comte´ region published in 1825 (Taylor 1825), and in Volume 2 of the Ge´ographie Universelle written by Elise´e Reclus (Reclus 1877). Reclus was the first to describe the hydraulic engineering works, undertaken in 1871, along the course of the Valserine River that facilitated the electrification of the town of Bellegarde. Throughout the nineteenth century visual representations of the locality multiplied, such as the lithography presented in the work undertaken on the personal command of Emperor Napoleon III to celebrate the annexation of Savoy to France (Charpentier 1864). The book shows the outstanding landscapes of Savoy, Haute-Savoie and Nice, French provinces since 1860. The lithography of the ‘Loss of the upper Rhoˆne’ shows (geo)tourists on the new bridge of Lucey (see Fig. 3). Numerous tourist postcards of the site also encouraged its national promotion. Geoscientists were also interested in the site for its stratigraphical, palaeontological and geomorphological interest. In 1817, Alexandre Brongniart

Fig. 2. (Geo)tourists visiting the ‘Loss of the upper Rhoˆne’ during their travel to Geneva or to the spa resort of Aix-les-Bains. Lithography (p. 41) in Volume 1 of Albanis Beaumont’s (1800) Travels from France to Italy through the Lepontine Alps.

202

N. CAYLA ET AL.

Fig. 3. The new bridge of Lucey in 1860 offered the best viewpoint to the ‘Loss of the upper Rhoˆne’. Lithography in Nice et la Savoie (p. 45). Collections Public Libraries Chambe´ry SAV D12 1864.

FROM TOURISM TO GEOTOURISM: FRENCH ALPINE FORELAND

(1770–1847) discovered in the glauconitic chalk which overlies the Urgonian limestone the same fossils he had already observed in the chalk cliffs of Folkstone and Normandy, as well as in the compact limestone at the top of the Fiz chain near Chamonix (Cuvier & Brongniart 1828). This observation allowed him to confirm in a speech at the Academy of Sciences the superiority of palaeontological characteristics for determining the comparative age of rocks, the essence of biostratigraphy. In 1878, Eugene Renevier (1831–1906) discovered that the rock layering of the area was in a normal stratigraphic succession (see Fig. 4; Renevier 1854). For over a century, several generations of geologists have contributed to the indepth knowledge of this geomorphosite (Renevier 1854; Daubre´e 1887; Martel 1912; Lugeon 1922; Jayet 1927; Gignoux & Mathian 1950). The heritage value of the ‘Loss of the upper Rhoˆne’ is important, both scientifically as well as for its ‘additional values’ (in the sense of Reynard 2005): aesthetic, educational and cultural. This site was therefore selected in the Rhoˆne-Alpes regional list of the national inventory of geological heritage,

203

even if it was difficult because the nomenclature of the inventory does not provide for the case of the missing geosite (Reynard et al. 2011). Along with the construction of the heritage path, a series of development projects have gradually threatened the site. As early as in 1796 the Sieur de Monville Boissel attempted to navigate through the gorges in the light of future development that would allow timber floating from the forests of the Swiss mountains to the port of Marseille. This attempt ended in failure (Boissel de Monville 1794). Passionate opposition against the proposal for hydro-electric development of the Rhoˆne in the early twentieth century was made by a figure of French karst research, Edouard-Alfred Martel (1859–1938). A defender of geological and landscape heritage, of which the ‘Loss of the upper Rhoˆne’ was the flagship, after several visits in 1897, 1902 and 1910 Martel published a richly illustrated book that included a plea for saving this unique heritage which he described as ‘one of the most curious sites in the world’ (Martel 1912). Unfortunately, his plea was in vain. After more than 40 years of gestation, the Ge´nissiat dam was built 10 km downstream of Bellegarde;

Fig. 4. First geological study of the ‘Loss of the upper Rhoˆne’ from Renevier’s 1854 Me´moire ge´ologique sur la Perte du Rhoˆne et ses environs (p. 82). It provided a geological interpretation of the confluence of the Rhoˆne and the Valserine and cliffs bordering the loss of the upper Rhoˆne.

204

N. CAYLA ET AL.

this resulted in the flooding and consequential covering with silt of this significant geoheritage site, and a great loss to geotourism. From 1978 however, the periodic discharge of the dam’s reservoir reveals the former bed of the canyon that ceased ‘plunging in the lapse of memory the famous loss of the Rhoˆne’ (Le Dauphine´ Libe´re´, 20/05/2010; Reynard et al. 2011; SaintPierre 2013; Fig. 5). This ephemeral event attracts the curious and nostalgic (geo)tourist. Only a few people today are able to locate the precise location of this ‘loss’, especially as the Valserine has a similar phenomenon just upstream of Bellegarde. Identified on maps as the ‘loss of Valserine’, it is still visible today and in the form of a miniature replica and imitation of the original ‘loss’. The case of the ‘Loss of the upper Rhoˆne’ is not unique; many large reservoirs, as well as many smaller construction works, potentially threaten geomorphosites worldwide. For example, a dam that would have destroyed the Portas de Almoura˜o geomonument in Portugal’s Narturtejo geopark, an epigenic gorge (Neto de Carvalho & Rodrigues 2010),

would have been built if it wasn’t for the mobilization of local activists.

The Sierroz Gorge: a famous geosite sunk into obscurity Three kilometres NE of the spa town of Aix-lesBains, another gorge of previous international renown is now something of a derelict tourist site enclosed within the peri-urban extension of Aix agglomeration. The picturesque waterfall of Gre´sysur-Aix is the result of the confluence of the Sierroz waters with those of the Deysse River. Following the waterfall, the river rushes through a narrow gorge that cuts the Urgonian limestone over a few hundred metres in length to a depth of 20 m in places. One of its first visitors was Horace Benedict de Saussure, who described the way the water was used at that time to power mills and presses (Saussure 1796). This activity was already mentioned on the Sardinian map of 1728 (Be´relle 2013). Because the cascade was so close to the spa town

Fig. 5. A temporary view of the remains of the geosite (Bridge of Lucey) during the periodic discharge of the Ge´nissiat dam’s reservoir (pictures from Roger Fillion) in 2012.

FROM TOURISM TO GEOTOURISM: FRENCH ALPINE FORELAND

of the lake of Bourget, visitors (almost 8000 guests in the city in 1862) were numerous and famous (Gauchon 1997). What triggered the romantic reputation of the gorge was an accident which took place on 10 June 1813. Hortense de Beauharnais (1783– 1837), daughter of the Empress Josephine (the first wife of the Emperor Napoleon I), regularly stayed in Aix-les-Bains from 1810. During a walk to the waterfall of Gre´sy-sur-Aix in 1813, her young lady companion and childhood friend, the baroness Adele de Broc, fell into the rapids of the river and drowned in front of her. This tragedy quickly transformed the site into a ‘must see’ of romantic tourism (Pomini 2008), and the baroness was immortalized by the monument erected at the scene of the event by Queen Hortense (see Fig. 6). The destination was popularized across Europe by the great romantic authors Alphonse de Lamartine (1790– 1869) and Alexandre Dumas (1802–1870). Similarly, many royal personages popularized the destination, notably Queen Victoria (1819–1901), Emperor Dom Pedro of Brazil (1825–1871) and Empress Elisabeth of Austria (1837– 1898). When staying in Aix in 1832, Alexandre Dumas describes his visit to the waterfall and gorge with a focus on its hydro-morphological aspects: ‘When I visited it, the water was low, and let dry the opening of three funnels, which have fifteen to eighteen feet deep, and in the inner walls of which the water has carved a communication gnawing the rock, descending in this way down to the bed of a stream leaking thirty feet deep between two so close banks that it is easy to jump from one bank to another’ (Dumas [1834] 1980, p. 243). The visit which undoubtedly had the greatest impact was that of the imperial couple. Following the annexation of Savoy to France in 1860, Napoleon III, the son of Queen Hortense, discovered the places which had so profoundly affected his mother. To develop tourism, a path was first cut into the rock. During the second part of the nineteenth century, a walkway of some 300 m was built. In 1880, the construction of a dam by the company of the Sierroz gorges downstream of the site allowed navigation through the canyon by steamboat over a distance of 1 km (see Fig. 7). In 1910, the Sierroz gorge was classified as ‘natural monuments and sites of artistic character’ (Law ‘Beauquier’: 21 April 1906) becoming the first natural site protected by the French state in Savoy (Gauchon 2002). Being largely transformed by tourism activities, the Sierroz gorge was not threatened by new developments such as those which threatened the loss of the upper Rhoˆne. This distinction seems to have been put forward as a means of official recognition of the quality of this site at a time when the shores of Lake Bourget were welcoming more than 40 000 visitors a year (Connille 1998).

205

After World War II, a populist form of tourism replaced that of the crowned heads. During the 1970s, safety measures for tourist activities were strengthened and the authorities first prohibited the navigation and then actual physical access to the gorge. Gradually, the site and its access sank into obscurity, with the old mills crumbling away and the walkway becoming only a derelict tourist site wedged between the railway and the road (see Fig. 8). Despite this, for the centennial of the site’s original protection in 2010 the association: ‘Au coeur des Gorges du Sierroz’ was created. It aims to develop knowledge of and safeguard and enhance the geomorphosite. Its ultimate goal is the re-opening of the site so that it can become a popular walk once again. The picturesque waterfall of Gre´sy/Aix and the Sierroz gorge are not so spectacular, but their location a few kilometres away from a major tourist destination made them a famous tourist attraction from the first half of the nineteenth century. Although the geoscientific interest of the site is limited, its scenic quality and the tragedy coupled with the succession of famous visitors (especially the imperial couple) have contributed to its fame. This gorge is now a memorial site for the history of the city of Aix-lesBains, enriched by its botanical interest. Since 1992, these sites have been listed with other remarkable natural areas of the Savoy department.

The Fier gorge: created by and for the tourism industry The Fier River is a tributary of the Rhoˆne, which originates in the massif of Aravis-Bornes. It bypasses the town of Annecy and its lake by the north and crosses several gorges before it flows into the Rhoˆne downstream of Seyssel. It was during the inauguration of the Annecy –Aix-les-Bains railway line by the Railway Company of ParisLyon-Me´diterrane´e (PLM) on 5 July 1866 when the idea emerged for a tourism development of the Fier gorge. Indeed, the small railway station of Lovagny was only a few hundred metres from the gorge, an ideal trip to complete a visit to the Castle of Montrottier. The castle and the gorge are intrinsically linked, with the former built on a hill isolated by the erosion of the River Fier. A deep depression, the only trace of the former bed of the river, borders the north of the hill which plunges steeply on the south into the River Fier’s gorge. Some tourists were already coming to view the gorge at that time, as shown in the lithography prepared in 1864, but it was impossible to penetrate its deeper places (see Fig. 9). The ‘Operating Company of the Galleries’ was created on 30 June 1868, the infrastructure and

206

N. CAYLA ET AL.

Fig. 6. Monument erected to the memory of Ade`le de Broc who died in the Sierroz gorge during her visit with Hortense de Beauharnais. Collections Public Libraries Chambe´ry SAV C 120.

FROM TOURISM TO GEOTOURISM: FRENCH ALPINE FORELAND

Fig. 7. Tourist visit on a steamer to the Sierroz gorge. Touring Club de France, 1900.

207

208

N. CAYLA ET AL.

Fig. 8. Sierroz gorge and the Baronne de Broc Memorial in 2013; the remains of a popular tourist (geo)site sunk into obscurity.

amenities were completed in 1869 and the inauguration took place on 21 July of the same year (Descotes 1870). Several journalists were invited to the inaugural visit including Ame´de´e Achard and Adolphe Joanne, who then published a paper in L’Illustration, journal universel (Joanne 1869) that brought national attention to the event (see Fig. 10). The inaugural tour described by Franc¸ois Descotes in the Courrier des Alpes on 27 and 29 July 1869 explains the various key elements of the visit; 250 m of walkway, suspended more than 28 m above the bottom of the canyon, allowed discovery of the otherwise inaccessible gorge. At the exit, the walk ends beside the picturesque location of the Mer des Rochers, with an amazing patterned mega lapiaz. This type of karstic erosion was produced under forest cover but now the rock is scoured and naked. It is possible to observe kamenitza, a form of chemical erosion, also referred to as footprints, near the river bed that can evolve from potholes created by mechanical erosion. Boulder chaos can also be observed; this is a consequence of the erosion caused by the river flow that undermines the base of the rocky layers, resulting in the collapse of the upper part giving rise to the blocks. All these geomorphosites illustrate various stages of the erosive action of the river. Many other, perhaps lesser, geomorphological curiosities enhance the walk. Less than three months after its opening, 6000 (geo)tourists had discovered this new attraction. This attendance has constantly grown despite several devastating floods over a century or so (on 3 October 1888, 23 December 1918, 9 and 24 November 1944, 8 December 1944 and, the most recent and most destructive, 30 September 1960). In February 1943 the gorges

were registered under the law on the protection of natural monuments and sites. Today the site annually receives more than 80 000 visitors and ranks in the top ten busiest Haute-Savoie tourist sites. In 2009, ‘la clairie`re des curieux’ was fitted out to show visitors the geoheritage interest of the Fier gorge. At the entrance, a panel explains the origin of the River Fier (see Fig. 11). Other thematic panels explain the different geoscientific particularities of the geomorphosite, an educational aspect that plays an important role in geotourism (Gordon 2012; Hose 2012; Martin 2014).

Synthesis and discussion These three examples of geoheritage evolution for the same type of geomorphosite illustrate how tourism and heritage are in constant interaction (Suchet 2010). In all three cases, the tourist development, fame or tourist facilities precede, accompany or induce the heritage designation of these sites. This process resulted in the official protection of the two Savoyards geomorphosites, but failed in the case of ‘the loss of the upper Rhoˆne’. Indeed, the flooding of the latter under the waters of the Ge´nissiat dam compromised its transmission to future generations other than by historical iconographies, stories and literature monographs (Saint-Pierre 2013); perhaps this example does have the positive outcome of emphasizing the significance of such secondary geotourism materials for geohistorical studies (Hose 2008, 2012). The cases of the Fier and Sierroz gorges show that the tourism practices developed were generally

FROM TOURISM TO GEOTOURISM: FRENCH ALPINE FORELAND

209

Fig. 9. The Fier gorge and the castle of Montrottier, the first (geo)tourists in 1860 before the inauguration of the touristic visit in 1869. Lithography in Nice et la Savoie. Collections Public Libraries Chambe´ry SAV D 12 1864.

210

N. CAYLA ET AL.

Fig. 10. Inaugural visit to the Fier gorge, a new tourist attraction in the northern Alps. From L’illustration (July 1869).

FROM TOURISM TO GEOTOURISM: FRENCH ALPINE FORELAND

211

Fig. 11. Geological interpretive panels realised by Fabien Hoble´a and installed in 2009, intended to enhance the geotourism orientation of the Fier gorge (Hoble´a 2014).

chosen to be compatible with their conservation and transmission to future generations. Although early developments have generated strong landscape impacts, time has blurred their perception among the generations who did not experience the initial pre-development state of these landscapes. The age of these tourist amenities gives an additional heritage element to these sites, historical and cultural, referring to the beginning and the first ‘golden age’ of the tourist boom in the Alps and the Jura mountains. This was strongly linked to thermal and spa tourism development throughout the nineteenth century. While both Savoyard geomorphosites are in a good state of geoconservation, it is not the same for the tourist facilities that occupy them. Abandoned, and in the process of ruination in the Sierroz gorge, they are reduced to a tourist wasteland. In the Fier gorge they have been renovated and complemented by geodidactic media, bringing it into the era of modern geotourism. The current concept of geotourism (Hose 2012), which strongly mixes the geotourism with geointerpretation for educational perspectives, could help to revive the Sierroz gorges currently identified

as a ‘satellite geosite’ by the Regional Natural Park of the Bauges massif, located in the immediate vicinity. This is especially likely since the latter became a member of the European Network of Geoparks in 2011 (Zouros 2004; Martini & Zouros 2008; Desbois & Hoble´a 2013). The geopark could provide decisive support for the redevelopment project by the local association for the protection and the enhancement of the site (Pomini 2008). In all cases the sustainability of tourism practices appears as a somewhat fragile protective measure and tourism development of a geosite might conflict with other issues of territorial development that can literally lead to its loss, such as in the case of the loss of the Rhoˆne. The Rhoˆne’s touristic value was totally destroyed and the use of substitution values today confined to the economic value of hydropower production has destroyed its heritage value. The heritage interest of a site can therefore evolve following the desires and needs of a society; the Sierroz gorge example shows that the official classification of a geosite or geomorphosite does not guarantee either its protection or its longterm maintenance.

212

N. CAYLA ET AL.

The heritage process has today evolved towards a rationalization of the procedures with its own terminology, and the objectification of the selection process and site assessment (Hose 2005; Reynard 2009) based on the prominence of scientific values. Geology and geomorphology are the first target of heritage, rather than the landscape they subtend. Aesthetics are relegated to ‘additional values’ among other emerging values, including teaching (the link and reference to the past to serve the present and future). Recently, the cultural values are beginning to appear as a major lever for the assessment and interpretation of geosites (Panizza & Piacente 2003, 2005). Thus, in this way, geosites and geomorphosites in the Alps, and elsewhere in Europe, play their role in that heritage renewal. Thus, in a globalized world, territories tend to get closer to their specific heritage, which make their originality. The detailed and helpful reviews of earlier drafts of this paper by two anonymous reviewers are gratefully acknowledged. We are indebted to the volume editor, T. A. Hose, for providing additional information and for correcting and proofreading the final draft of this paper. Any errors or omissions must, of course, remain those of the authors.

References Albanis Beaumont, J. F. 1800. Travels from France to Italy through the Lepontine Alps, 1. Be´relle, A. 2013. Les gorges du Sierroz 1813–2013, un haut-lieu du tourisme alpin. La Rubrique des patrimoines de la Savoie, 31, 6 –7. Boissel de Monville, T. C. G. 1794. Voyage pittoresque et navigation exe´cute´e sur une partie du Rhoˆne re´pute´e non navigable, moyens de rendre ce trajet utile au commerce. Dupont, Paris. Cayla, N. 2009. Le patrimoine ge´ologique de l’arc alpin: de la me´diation scientifique a` la valorisation ge´otouristique. PhD thesis, Universite´ de Savoie, Laboratoire EDYTEM, France. Cayla, N., Hoble´a, F. & Gasquet, D. 2009. Place de la ge´omorphologie dans l’offre touristique de l’arc alpin: du re´el au virtuel. In: Proceedings of the Geomorphosites 2009: Raising the Profile of Geomorphological Heritage through Iconography, Inventory and Promotion, Paris. Charpentier, H. (ed.) 1864. Nice et Savoie, sites pittoresques, monuments et histoire. De´partements de la Haute-Savoie, Paris et Nantes, 2. Connille, J. F. 1998. Les gorges du Sierroz: un patrimoine oublie´. Arts et me´moire, 12, 21–38. Cuvier, G. & Brongniart, A. 1828. Description des environs de Paris, E. d’Ocagne, Paris. Daubre´e, G. A. 1887. Les eaux souterraines a` l’e´poque actuelle. Dunod, Paris, 1. Deline, P., Nussbaumer, S., Vincent, C. & Zumbu¨hl, H. J. 2012. Mer de Glace, Art & Science, ESOPE.

Descotes, F. 1870. Lovagny gorges du Fier et lac d’Annecy: itine´raire pratique, historique et pittoresque a` travers la Haute-Savoie. Chambe´ry. Desbois, J.-L. & Hoble´a, F. 2013. Le patrimoine ge´ologique, marqueur territorial du parc naturel re´gional du Massif des Bauges. Revue Espaces, 315, 122– 129. Dumas, A. [1834] 1980. Impressions de voyage: Les Alpes de la Grande Chartreuse a` Chamonix, Encre, Paris. Duval-Massaloux, M. & Gauchon, C. 2010. Tourisme, ge´osciences et enjeux de territoires. Ge´otourisme. Te´oros, 29-2, 3– 14. Gauchon, C. 1997. Des Cavernes et des Homes: Ge´ographie Souterraine des Montagnes Franc¸aises. Gauchon, Chambe´ry, Karstologia, Me´moires no. 7. Gauchon, C. 2002. Les sites naturels classe´s entre 1906 et 1930 dans les Alpes du Nord: entre tourisme et protection, bilan et actualite´. Revue de Ge´ographie Alpine, 90, 15–31. Gignoux, M. & Mathian, J. 1950. Les conditions ge´ologiques de l’ame´nagement hydro e´lectrique du Rhoˆne entre Gene`ve et Seyssel. Ge´nissiat, Special issue, La Houille Blanche, Grenoble. Gordon, J. E. 2012. Rediscovering a sense of wonder: Geoheritage, geotourism and cultural landscape experiences. Geoheritage, 4,, 65–77. Hoble´a, F. 2014. In the folds of the earth: French prealpine geomorphological landscapes. In: Fort, M. & Andre, M. F. (eds) Landscapes and Landforms of France. Springer, London & New York, 183–194. Hose, T. A. 2005. Landscapes of Meaning: Geotourism and the Sustainable Exploitation of the European Geoheritage. Abstract, Institute of Geography, University of Lausanne. Hose, T. A. 2008. Towards a history of geotourism: definition, antecedents and the future. In: Burek, C. V. & Prosser, D. (eds) The History of Geoconservation. Geological Society, London, 37–60. Hose, T. A. 2012. 3 G’s for modern geotourism. Geoheritage, 4, 7– 24. Jayet, A. 1927. Etude stratigraphique de la perte du Rhoˆne pre`s de Bellegarde: Ain, France. Eclogae Helvetiae, XX, 160–222. Joanne, A. 1869. Inauguration de la galerie des Gorges du Fier. l’Illustration Journal Universel 27e`me anne´e, LIV, 1369, 71– 73. Lugeon, M. 1922. Etude ge´ologique sur le projet de barrage du Haut-Rhoˆne franc¸ais a` Ge´nissiat. Memoire de la Socie´te´ ge´ologique de France, 4, se´r., 2, me´m. No8. Martel, E.-A. 1912. Le canyon du Rhoˆne et les projets de Barrage de Ge´nissiat, de Malpertuis et de la perte du Rhoˆne. La Ge´ographie, XXV-5, 384– 389. Martin, S. 2014. Interactive visual media for geomorphological heritage interpretation. Theoretical approach and examples. Geoheritage, 6, 149 –157. Martini, G. & Zouros, N. 2008. Geoparks. . . a vision for the future. Ge´osciences, 7/8, 182– 189. Neto de Carvalho, C. & Rodrigues, J. 2010. Managing delicate socio-environmental impacts: Naturtejo European Geopark and the building of the Alvito Reservoir at Almoura˜o geosite (Portugal). Abstracts book. 9th European Geoparks Conference, Lesvos, 84–85. Panizza, M. 2001. Geomorphosites: concepts, methods and example of geomorphological survey. Chinese Science Bulletin, 46, 4– 6.

FROM TOURISM TO GEOTOURISM: FRENCH ALPINE FORELAND Panizza, M. & Piacente, S. 2003. Geomorfologia culturale. Pitagora, Bologna. Panizza, M. & Piacente, S. 2005. Geomorphosites: a bridge between scientific research, cultural integration and artistic suggestion. In: Il quaternario Volume speciale, Geomorphological Sites and Geodiversity, 3– 10. Pomini, S. 2008. Les gorges du Sierroz. Entre reˆves et re´alite´, un lieu romantique. Gap-Editions, Challesles-Eaux. Portal, C. 2010. Reliefs et patrimoine ge´omorphologique. Applications aux parcs naturels de la facade atlantique europe´enne. PhD thesis, Universite´ de Nantes, France. Reclus, E. 1877. Nouvelle ge´ographie de la France. Hachette, Paris, 2. Reichler, C. 2002. La de´couverte des Alpes et la question du paysage. In: Lausanne, G. (ed.) Le voyage dans les Alpes, Georg, Gene`ve. Renevier, E. 1854. Me´moire ge´ologique sur la Perte du Rhoˆne et ses environs. Gene`ve, in-48. Reynard, E. 2005. Ge´omorphosites et paysage. Ge´omorphologie. Relief, processus, environnement, 3, 181–188. Reynard, E. 2009. The assessment of geomorphosites. In: Geomorphosites, Pfeil, Mu¨nchen, 63–73. Reynard, E., Hoble´a, F., Cayla, N. & Gauchon, C. 2011. Iconic Sites for Alpine Geology and Geomorphology Rediscovering Heritage? Revue de Ge´ographie Alpine: New heritage: objects, actors and

213

controversy Nouveaux patrimoines 99– 2. mis en ligne le 20 juillet 2011, http://rga.revues.org/index1435. html Rowlinson, J. S. 1998. ‘Our common room in Geneva’ and the Early exploration of the Alps of Savoy. Notes and Records of the Royal Society, 52, 221 –225. Saint-Pierre, D. 2013. Les gorges perdues du HautRhoˆne de la Suisse a` Ge´nissiat. Musnier-Gilbert Editions, Bourg-en-Bresse. de Saussure, H.-B. 1779–1796. Voyages dans les Alpes, pre´ce´de´s d’un essai sur l’histoire naturelle des environs de Gene`ve, Neuchaˆtel. Fauche-Borel, 4 volumes. Suchet, A. 2010. Le site touristique ame´nage´ des Gorges de la Fou en Pyre´ne´es franc¸aises Commercialisation, me´diation scientifique ou animation culturelle et controverse paysage`re. Te´oros, 29-2, 44– 54. Taylor, I. 1825. Voyages pittoresques et romantiques dans l’ancienne France. De L’Imprimerie de FirminDidcot, Paris. Windham, W. & Martel, P. 1744. An Account of the Glacieres or Ice Alps in Savoy, in Two Letters, One from an English Gentleman to his Friend at Geneva; the Other from Peter Martel engineer, to the said English gentleman. London, printed for Peter Martel. Zouros, N. 2004. The European Geoparks Network. Geological heritage protection and local development. Episodes, 27, 165–171.

Rediscovering geoheritage, reinventing geotourism: 200 years of experience from the Sudetes, Central Europe ´ PIOTR MIGON Institute of Geography and Regional Development, University of Wrocław, pl. Uniwersytecki 1, 50– 137 Wrocław, Poland (e-mail: [email protected]) Abstract: The Sudetes is a mountain range in Central Europe and an area of remarkable geodiversity. In recent years, the area has been promoted as a geotourist destination and various initiatives aimed at better understanding geoheritage have been implemented. An interest in scenic landscapes is not new however, and dates back to the end of the eighteenth century. Two areas within the Sudetes are cradles of local nature-based tourism. These are the granite massif of the Karkonosze in the west and the sandstone stepped plateau of Broumov Highland and Stołowe Mountains in the central part of the Sudetes. In both, physical access to the key geosites was provided as early as in the nineteenth century, while an interpretation component was added in the early twentieth century. A side-effect of political change following the end of World War II was the neglect and dilapidation of many sites, as well as the disappearance of geoheritage appreciation from the collective social memory. In the last decade many of those early achievements were rediscovered and provided the foundations for contemporary activities. An educational component based on modern science is now included in the features which were discovered as tourist attractions long ago.

Geotourism, understood as nature-based tourism with a special focus on the material legacy of Earth heritage, is often presented (sometimes perhaps unintentionally) as a new trend in global tourism. Although the term itself is relatively new and was coined by Hose in the mid-1990s (Hose 1995), gaining popularity soon afterwards, the phenomenon of travelling to see geological sites and spectacular landforms has a much longer history. The novelty of geotourism resides not so much in a new sphere of interest, but in the widespread use of modern tools to enrich experience and enhance understanding of natural phenomena. In Poland, as elsewhere, it would be inappropriate to claim that the interest in geological and geomorphological heritage of the area is a modern trend. The aim of this contribution is to show that tourism development in the mountain range of the Sudetes (Central Europe) has always had a significant ‘geotouristic’ component and the highly variable Earth heritage of the area has long been used as a foundation to build tourist products. Two regions may be considered as cradles of geotourism – the Stołowe Mountains with the Police Basin and the Karkonosze Mountains – and these will be described in more detail. It will also be demonstrated that certain tools used long ago to raise awareness in geological history proved remarkably efficient and constitute a valuable legacy of the past, still in use by modern geotourism providers.

The area Location and outline of geoheritage The Sudetes, a mountain range along the state border of the Czech Republic and Poland with a smaller NW part in Germany, belong to the circum-alpine belt of uplands and medium-altitude mountains that extends from the Rhoˆne graben in the west to the Carpathians in the east (Fig. 1a). Geologically, they are an old Variscan orogen rejuvenated by differential uplift and subsidence during Cenozoic time that occurred in response to stresses transmitted from the Alps (Zuchiewicz et al. 2007; Reicherter et al. 2008). The Sudetes are the highest range within this belt, peaking at Mount S´niez˙ka (1602 m a.s.l.). The Variscan Sudetes were composed of late Proterozoic– early Palaeozoic metamorphic and igneous rocks, mainly gneisses, mica schists, granites, amphibolites, slates and greenschists. Gravitational collapses within the orogen resulted in intramontane basins filled with thick sedimentary sequences. Numerous granite intrusions and acid to intermediate volcanism accompanied the late orogenic stage (Fig. 1b). Subsequent geological history is recorded in the discontinuous sedimentary sequence involving terrestrial and marine deposits of Permian, Early– Middle Triassic and Late Cretaceous age, among which sandstones play a significant part. Cenozoic additions to the geological complexity of the Sudetes include volcanism, mainly

From: Hose, T. A. (ed.) 2016. Appreciating Physical Landscapes: Three Hundred Years of Geotourism. Geological Society, London, Special Publications, 417, 215–228. First published online November 13, 2014, http://dx.doi.org/10.1144/SP417.2 # The Geological Society of London 2016. Publishing disclaimer: www.geolsoc.org.uk/pub_ethics

216

´ P. MIGON

Fig. 1. The Sudetes. (a) Position of the Sudetes (framed) in relation to European Variscides and Alpides. (b) General features of geology of the central and western Sudetes; the broken line indicates the position of the state border between the Czech Republic and Poland. (c) Relief of the central and western Sudetes shown through a shaded relief model. Numbers indicate the approximate altitude of summit surfaces of individual up-faulted blocks. Town names are contemporary Polish names.

basaltic, although it was largely restricted to the western part of the range (Badura et al. 2005). During Neogene time the Sudetes experienced differential uplift and subsidence, which produced the gross morphology seen today (Zuchiewicz et al. 2007; Migon´ 2011). The Sudetes are remarkably diverse geomorphologically (Migon´ 2011). The principal geomorphic

features such as dissected mountain fronts are of tectonic origin (Fig. 1c). However, they coexist with a variety of rock-controlled landforms, including ancient volcanic massifs, inselbergs, sandstonesupported plateaus and escarpments, river gorges and tors. A clear imprint on landform patterns was left by the cold periods of the Pleistocene. The evidence of past periglacial conditions is provided

REINVENTING GEOTOURISM IN THE SUDETES

by, among others, widespread block fields, frostriven cliffs, cryoplanation terraces and relict rock glaciers (Traczyk & Migon´ 2003). In the highest part of the Sudetes conditions were suitable for the development of local glaciation (Engel et al. 2010). Marginal parts of the Sudetes were twice reached by the Scandinavian ice sheet, but its erosional and depositional effects are relatively minor (Hall & Migon´ 2010).

Further Earth resources for tourism Resources for (geo)tourism are not limited to geological and geomorphological heritage. An important component of regional geodiversity is the abundance of mineral, including thermal, waters. The earliest records of the existence of bathing facilities at springs date back to the thirteenth century (Cieplice, La˛dek), while the rapid development of spa resorts began in the eighteenth century and peaked at the end of the nineteenth century (Mazurski 2012). As a response to contemporaneous trends towards outdoor recreation, spa parks were enlarged, paths to local natural attractions (rock outcrops, caves, gorges) way-marked and natural springs became tourist attractions in their own right (Potocki 2004; Mazurski 2012). The development of spas and the development of ‘geotourism’ remained in synergic relationships. On the one hand, natural rock outcrops and strangely shaped landforms in the vicinity of spas added to their attractiveness and became excursion destinations. On the other, increasing interest from visitors prompted the development of facilities for tourism (access roads, trails, lodges, observation towers), the way-marking of new trails and locally trail engineering to allow access to particularly interesting spots. A complementary resource for tourism is mining heritage. The Sudetes were rich in ores (iron, copper, tin and many others), gold and silver, exploited since medieval times, as well as coal. Most mines were still in operation during the nineteenth century however, and the only site widely recognized as a significant tourist attraction was an underground canal for coal-laden barges in the mining town of Wałbrzych.

Historical context The highs and lows of geotourism in the Sudetes through time should be considered in relation to the historical context and changing political situation. Since mid-medieval times the Sudetes were a borderland between the Kingdom of Bohemia (in the south) and various Silesian duchies (in the north), more associated with Poland. From the late fourteenth to the middle eighteenth century the

217

area was a political entity, belonging first to the Kingdom of Bohemia and later the Austrian (Habsburg) Empire. Ethnically, the Sudetes population was a mix of native Polish, Czech and immigrant German, with the latter becoming increasingly more represented, especially in the Silesian part. Austrian –Prussian wars in the middle eighteenth century again divided the area, but the Prussian (Silesian) and Austrian (Bohemian) side remained culturally close to each other and the border was not a significant barrier limiting tourism development. The southern part of the Sudetes became an independent Czechoslovakia in 1918, while the northern part remained German. However, even on the Bohemian side the Sudetes were populated mainly by the German minority. In the aftermath of World War II the former German side of the Sudetes became Poland, the entire German population was expelled and the area resettled by Polish people arriving from central and eastern Poland. Initially, there was little interest in the German history of the area and little appreciation of achievements from the nineteenth and twentieth century. Tourism was hardly a priority in post-war times and a significant part of the existing infrastructure, including provisions for geotourism, disappeared. A gap between the medieval and contemporary history existed in the minds of most people, and it is only in the last two decades or so that we have rediscovered the more recent history of the Sudetes, including the history of tourism and travel (Potocki 2004; Mazurski 2012).

Early geotourism in the Stołowe Moutains and Police Basin Geodiversity The Stołowe Mountains and the adjacent Police Basin (Fig. 2) occupy the very central part of the Sudetes; both are mainly underlain by a sedimentary series of marine origin and Late Cretaceous age. Among them, sandstones play a prominent part in terrain morphology supporting escarpments, gorges, plateaus and mesas (Fig. 3a, b). Thick sandstone beds are cut by an orthogonal fracture network, usually widely spaced (.3– 4 m), and show a variety of sedimentary structures, among which large-scale cross-bedding is the most evident (Fig. 3c). Intervening mudstones and other fine-grained clastic rocks are associated with rather subdued topography, but are important in inducing slope instability in the overlying sandstones (Pulinowa 1989). The Stołowe Mountains are an elevated multi-storey plateau bounded by sandstone-capped escarpments. Mount Szczeliniec Wielki (Great Fissured Hill) is the highest peak (919 m a.s.l.),

218

´ P. MIGON

Fig. 2. Geographical setting and geotourist highlights of the Stołowe Mountains and Police Basin. The outcrop area of Cretaceous deposits is indicated in white. Numbers indicate elevation in metres above sea level (m a.s.l.).

an isolated mesa with an extensive rock labyrinth on top. Major geomorphic features in the Police Basin are the cuesta of Broumovske´ steˇny with a network of backslope canyons, the mesas of Ostasˇ (700 m) and Hejda (628 m), and the large, heavily dissected sandstone plateau of Adrsˇpasˇsko –Teplicke´ ska´ly, with extensive ‘rock cities’ (Fig. 3d).

Beginnings Although the settlement pattern around the sandstone plateaus and mesas in the Middle Sudetes was largely established as early as in the thirteenth– fourteenth centuries (Migon´ & Latocha 2013), the canyon-riddled sandstone terrains themselves long remained inaccessible wilderness areas. The

beginning of tourism focused on rock formations dates back to the 1730s. It first developed in the Police Basin due to better accessibility. The most famous and most visited locality was the ‘rock city’ of Adrsˇpach, conveniently located next to a wide valley and a village that had existed since medieval times. One traveller’s report from 1779 (cited in Imlauf & Bohadlo 1997) mentions a gate to the rock labyrinth, a clear indicator that the area was somehow developed to serve tourists. Travelogues of early nineteenth century visitors indicate that they not only marvelled at the uniqueness of the scenery but also wondered about its origin, recalling earthquakes and powerful torrents as possible formative agents. Nevertheless, no sound explanation of the story behind the scenery could have been offered at that time.

REINVENTING GEOTOURISM IN THE SUDETES

219

Fig. 3. Geoheritage of the sandstone terrains in the central part of the Sudetes. (a) Broumovske´ steˇny, the northern escarpment of the sandstone plateau, is 200–300 m high. A wide trough excavated in weaker Permian deposits can be seen in the background. (b) The sandstone capped mesa of Mount Szczeliniec Wielki has a relative elevation of 150 m. (c) One of many tors, c. 10 m high, in the Upper Turonian sandstone on the top surface of Mount Szczeliniec Wielki. (d) One of the most famous sandstone towers in Adrsˇpach Rock City is called Milenci (Lovers), alluding to the visual resemblance to a kissing couple. All these sites were appreciated for their scenic values as early as in the beginning of the nineteenth century.

220

´ P. MIGON

Tourism development in the adjacent Stołowe Mountains lagged behind, apparently due to poorer access, scarce settlement network and higher altitude. The main plateau of the Stołowe Mountains was reached by human colonization in the early eighteenth century and the village of Karło´w (then Carlsberg) was founded in 1730. As a result of war between Austria and Prussia (1740–45), the Stołowe Mountains became a borderland soon after. During another period of tension between the two powers in 1790, plans were advanced to fortify the area. The mesa of Mount Szczeliniec Wielki was considered as a possible site for a border fortress but these plans were abandoned in favour of another location nearby at Mount Ptak. However, repeated visits of Prussian officials to the mesa helped to spread information about its curiously shaped landforms, their unusual abundance and wide vistas from mesa rims. Members of the Prussian royal family, including King Friedrich Wilhelm II, visited the area in 1790, adding fame to the locality. An increase in popularity of the site resulted in both technical and legal developments. As early as 1790 the first path to the upper mesa surface was traced across the boulder and rock slopes, using the experience of a local named Franz Pabel. In subsequent years Pabel accompanied tourists as a guide and received an official guiding license in 1813, probably the first of its kind in Central Europe (Pabel 1843). In 1825 the existing trail was extended to the highest tor on the mesa and then back to Karło´w, closing the circuit.

Johann Wolfgang Goethe: a pioneer of geotourism in the Sudetes If a definition of geotourism involving learning about geoheritage is accepted (Hose 1995; Joyce 2007), it is then debatable whether the early visitors to the sandstone area of the Central Sudetes were geotourists. Nevertheless, at least one of them certainly qualified as a geotourist: Johann Wolfgang Goethe (1749–1832), the famous German poet and writer but also a naturalist. While his travels to sites of geological interest in central Germany, western Bohemia, Switzerland and Italy are well known (Geyer et al. 2007) and have also been presented in the context of ongoing plans to develop geotourist products in Italy (Panizza & Coratza 2012), his early studies in the Sudetes are somewhat overlooked. Goethe visited the area in 1790 shortly after returning from a trip to Italy, and focused on the sandstone terrains in the central Sudetes (Chmal 2010). In late August 1790 he climbed the mesa of Mount Szczeliniec Wielki and then moved on to the ‘rock cities’ Adrsˇpach and Teplice, providing a scientific account of their morphology and possible origin. In each of these

localities plaques commemorate his visits and studies (Fig. 4).

Improvements in geotourist provisions In the early nineteenth century two ‘star’ attractions of the area were firmly established: the mesa of Mount Szczeliniec Wielki and the ‘rock city’ of Adrsˇpach. Although an educational geoscientific component of tourism was missing, it is evident from travellers’ diaries (Jaworski 1988; Imlauf & Bohadlo 1997) that the main motivation for visiting these sites was associated with the willingness to see ‘wonders of nature’ and experience close contact with bizarre rocks and landforms. Initially, (geo)tourism provision was rather simple and in each area one designated walk could have been attempted. However, the ‘discovery’ of another assemblage of sandstone walls and canyons next to the town of Teplice after a great forest fire in 1824, and its development for tourism in the late 1840s, resulted in a sort of competition between the sites and further trail engineering works to open more areas to the public. In the second half of the nineteenth century further parts of the Szczeliniec Wielki mesa were made accessible through major works in the deep open clefts along the NE escarpment. After completion of this path, tourists would have had a full overview of all the important geomorphic elements of the mesa: talus slopes on the way up and back; outer rock cliffs; boulder chaos; tors and ruiniform relief; and deep clefts opened by forward toppling of marginal parts of the mesa. The development of tourist provisions was gradually spreading out and the network of waymarked trails increased. For instance, several trails leading steeply up the northern escarpment of the Stołowe Mountains were opened to connect the foothills with the mesa of Mount Szczeliniec Wielki. The potential of non-karstic caves was also exploited. The most imposing among them, Skalnı´ chra´m (Rock Church) with its .20-m-high corridors, was a highlight of every trip to the Teplice ‘rock city’ from as early as the 1850s.

Early geotourism in the Karkonosze Mountains and surrounding areas Geodiversity The Karkonosze Mountains (Figs 5 & 6) are the highest part of the Sudetes (Mount S´niez˙ka, 1602 m a.s.l.) and the only areas that impart the flavour of a high-mountain environment (Migon´ 2008). The summit plateau is undercut by landforms

REINVENTING GEOTOURISM IN THE SUDETES

221

Fig. 4. Plaques commemorating the 1790 visit of Johann Wolfgang Goethe to the sandstone areas of the central Sudetes: (a) at Mount Szczeliniec Wielki, with a plaque commemorating the visit of John Quincy Adams in 1800 to the right; (b) in the Teplice Rock City (translation from Czech: ‘on 30 August 1790 a German poet J. W. Goethe visited the Teplice Rock City’); and (c) in the Adrsˇpach Rock City, with the bust of Goethe beneath (translation from German: ‘Johann Wolfgang Goethe – the prince of German poets – visited the Adrsˇpach Rock City to pursue the scientific study of nature in the year 1790’).

of glacial erosion, mainly cirques, whose depth ranges from 100 to 200 m (Fig. 6b). The most characteristic landforms in the Karkonosze are granite tors however, with the most striking examples up to 100 m long and 25 –30 m high (Fig. 6c). Further components of landform diversity are block fields, scree cones, glacial moraines, waterfalls, slot canyons, potholes, open clefts and non-karstic caves and scattered boulders. The area is also a mineralogical treasure, with more than 200 varieties described so far (Sachanbin´ski 2005). The beginnings of mineral prospecting date back to prehistoric times, and mineralogy has intensified since the thirteenth–fourteenth centuries. At many localities regular mining was carried out, lasting until the second half of the twentieth century and leaving a rich legacy of underground works and surface modifications. The intramontane basin of Jelenia Go´ra extends to the north of the Karkonosze. Geologically, it is a subsided part of the granite massif, downfaulted during Cenozoic time along WNW –ESE faults. The floor of the basin has retained its pre-faulting topography, typified by tens of isolated granite

hills (inselbergs), usually crowned with rock outcrops (Fig. 5). These are mainly tor clusters, but locally more extensive assemblages of ruiniform relief occur (Fig. 6d).

Early nature-based tourism The first leisure trips into the Karkonosze were organized in the sixteenth and seventeenth centuries and were targeted at two important destinations: the summit of Mount S´niez˙ka and the source of the Labe river, one of the longest rivers in Central Europe (Mazurski 2012). While both sites may be considered as geosites by modern standards, there are no indicators that the early visitors had special geoscientific interests. The vast 3608 panorama from the top of Mount S´niez˙ka was certainly a reward after a strenuous hike, but people were motivated mainly by the perspective of reaching the highest mountain of Bohemia. Notwithstanding these early explorations, the onset of actual development of tourism as a wider social phenomenon in the Karkonosze and its surroundings occurred around the turn of the nineteenth

222

´ P. MIGON

Fig. 5. Geographical setting and geotourist highlights of the Karkonosze Mountains and Jelenia Go´ra Basin. The outcrop area of Karkonosze granite is indicated in white.

century (Potocki 2004; Mazurski 2012). As well as Mount S´niez˙ka, the most valued destinations included the glacial cirques of S´niez˙ne Kotły and Mały Staw with a mountain lake in the latter, waterfalls, Mount Chojnik with ruins of a medieval castle and granite cliffs below, and bizarre rock formations, especially if these served as good viewing sites at the same time. One sign of their popularity in those distant times was the construction of inns offering shelter and refreshments. In the Jelenia Go´ra Basin vast noble estates

were converted into landscape parks following the contemporaneous spirit of Romanticism. They integrated natural elements, such as granite tors, big solitary boulders, deep open clefts and crevice caves, with a built environment (Migon´ & Latocha 2008). All these localities were appreciated for their aesthetic value and as recreation spots rather than a testimony of Earth heritage. Nevertheless, their popularity indicates where values were sought and set the stage for future developments.

REINVENTING GEOTOURISM IN THE SUDETES

223

Fig. 6. Selected geosites of the Karkonosze Mountains and the Jelenia Go´ra Basin visited since the late eighteenth century: (a) summit surface of the Karkonosze, with peat bogs and Mount S´niez˙ka in the distance; (b) glacial cirques of S´niez˙ne Kotły, with rock walls up to 200 m high; (c) Pielgrzymy, one of the highest (c. 25 m) granite tors in the area; and (d) the rock labyrinth of Mount Witosza.

Riesengebirgsverein and its role The development of tourism in the Karkonosze accelerated from the 1870s, favoured by general political stability, the increasing wealth of the Prussian and Czech societies, technical developments (especially a considerable extension of railway network) and social developments (Potocki 2004). Infrastructural developments followed in response on both sides of the state border, and in the first half of the twentieth century the Karkonosze was an established and highly popular tourist destination. Various manifestations of the growing interest in tourism was the founding of local tourist societies (Potocki 2004; Mazurski 2012). The Riesengebirgsverein (RGV, Karkonosze Society) established in 1880 was the largest and most prominent society. Its statutory aims included tracing and marking of hiking trails; as much as 300 km of trails were built in the late nineteenth and early twentieth centuries, granting access to practically each site of interest in the region (Szczepan´ski 1989) including the most attractive geosites. More advanced engineering was occasionally needed, for example in the Kamien´czyk gorge below the Kamien´czyk waterfall where a 100 m bridge attached to the

rock wall was constructed in 1892. In parallel, owners of local businesses commonly made available local geo-curiosities in the vicinity of their properties (tors, clefts, rocky channel sections), apparently to increase the popularity of the site. The aims of RGV included further activities such as supporting nature conservation, collecting movable objects of natural and cultural heritage from the area, running a regional museum and a library, regular publishing and cooperation with academia. Since the beginning of the society it has published its magazine Wanderer im Riesengebirge, which contains information about day-to-day business but increasingly also materials about local heritage, both natural and cultural. These activities, crossing the boundaries between science and society, directly bear on the issue of appreciation of regional geoheritage. Here the RGV played a pivotal role.

Increasing interest in geoheritage The early twentieth century witnessed increasing efforts to popularize local geoheritage and increase laypersons’ understanding of the geological past, which became better and better understood by

224

´ P. MIGON

academics conducting their research in the Karkonosze. Various actions were undertaken and a few deserve special mention. First, the role of Wanderer im Riesengebirge needs to be acknowledged. Numerous articles about local geological and geomorphological curiosities were published, usually 1–3 pages long, providing both description and interpretation. They addressed issues such as the extent and chronology of mountain glaciation (a hot topic around 1900s), the origin of block fields and tors and the diversity of granite textures. Second, a novel means to show the geology and its relationship to landforms was a model crosssection of the West Sudetes, erected in 1902 in Jelenia Go´ra upon an initiative of local RGV activists. At a scale of 1:5000, it is 20 m long and 0.6 m high (Fig. 7). It was built using local rocks on a south-facing slope, so that the background was provided by the main ridge of the Karkonosze

and visitors had an opportunity to relate the model representation to the reality. Third, a catalogue of sites of special geological and geomorphological interest in the region was published (Gu¨rich 1914). The book contains descriptions of around 100 localities (Table 1) accompanied by photographs and, occasionally, by plans and sketches. This book, published exactly 100 years ago, is surprisingly modern in approach and style and still serves as a major reference source for geoheritage-oriented activities in the region. Last but not least, in the 1920s and 1930s the nature conservation movement began and the value of a number of geosites was highlighted by establishing legal protection. Nature reserves were created in geomorphologically significant localities such as all glacial cirques in the Karkonosze, at outcrops (beyond the Karkonosze) of Cretaceous sandstones on the top of a cuesta near Lwo´wek S´la˛ski

Fig. 7. Geological cross-section of the West Sudetes in Jelenia Go´ra: (a) central part of the section, after renovation; and (b) explanation board, which is a part of a historic nature trail (stop no. 8) in the town’s park. The text describes the history of the site and the photograph below shows G. Gu¨rich, who came up with the proposal to build the geological model. Photographs courtesy of Andrzej Paczos.

REINVENTING GEOTOURISM IN THE SUDETES

Table 1. Geosites included in Gu¨rich’s (1914) catalogue. Note that one site may contain references to various geological and geomorphological phenomena. Theme

Primarily geological interest Outcrops of specific rock types Contact phenomena at granite/ country rock boundary Rare rock textures Mineralogical sites Exposures of specific joint patterns Primarily geomorphological interest Tors Bizarre shapes of rock outcrops Weathering pits and related forms Glacial cirques Waterfalls and gorges Potholes Block fields Mass movement phenomena Semi-permanent snow patches

Number of sites described 17 6 2 1 12 59 25 53 7 6 4 3 1 1

(then Lo¨wenberg) and the basaltic plug of Ostrzyca with its spectacular block fields and scree slopes (Potocki 2004). Most isolated tors, some caves and specific rock outcrops (e.g. aplites with particularly regular, chessboard-like fracture patterns) were declared nature monuments.

The legacy of early ‘geotourism’ The legacy of pre-World War II nature-oriented tourism is twofold. The relevant aspects are (1) accessibility and (2) information and awareness. However, while some continuity and appreciation may be seen regarding the former, much of the previous understanding of the value of local geoheritage was lost and past achievements were largely neglected. Reasons for this dual treatment of the historical legacy were ideological and can be linked with an officially encouraged trend to erase the Prussian/German period (1742–1945) from the regional history. Sadly, it was precisely those two centuries, and particularly the last 50 years prior to WWII, when most developments in ‘geotourism’ occurred. Nevertheless, the material base for post-World War II tourism was largely provided by the already existing infrastructure. Many hiking routes followed old trails, while most mountain lodges continued to play the same role. The top geotourist attractions, such as the sandstone labyrinths on Mount Szczeliniec Wielki, ‘rock cities’ in Adrsˇpach

225

and Teplice, glacial cirques and imposing granite tors of the Karkonosze, remained heavily visited and appreciated for their unusual appearance. At the same time however, many local geoheritagebased attractions fell into neglect. Firstly, the newcomers to the area did not have any positive emotional relationship with these local curiosities, apparently because they generally did not consider the Sudetes as their homeland. Secondly, many sites had been associated with various elements of cultural heritage, including commemorating plaques, rock-carved inscriptions and buildings bearing names of important locals, all indicating the German past. With the prevailing hostile attitude to the German legacy, this resulted in widespread devastation as well as omissions from available guidebooks. Consequently, without conservation and management trails deteriorated, infrastructure fell into disrepair, vegetation encroached and in 20 –30 years many once-popular sites disappeared from the collective memory. The general loss of awareness about the value of geoheritage requires a more complex explanation. Clearly there was very little general interest in the written legacy of the German past. Research focused on modern (i.e. pre-1945) history was discouraged and ideologically biased, not expected to show past achievements. Also, little knowledge of German language among the Polish population did not help. One can also argue that the main efforts in post-war times had to be directed elsewhere, to rebuild the local economy and establish societal links, not necessarily to conservation of local geoheritage. The above circumstances coincided with a marginal interest in geoheritage from the tourism industry in general, however. In most available guidebooks sites of geomorphological interest were presented without interpretation whereas rock exposures, that is, sites of primarily geological interest, were hardly mentioned at all, except in very specialized publications aimed at professional geologists (Grocholski 1969). As a consequence, the general public were not only unaware of a large number of geosites existing in the area, but lacked understanding of their geological and geomorphological significance due to the lack of interpretation.

Recent developments: more than 100 years after Although sporadic examples of a more focused interest in selected elements of geoheritage (e.g. caves, volcano remnants and sandstone rock formations) can be found in the 1980s and increasingly in the 1990s (Migon´ & Latocha 2010), it is really the opening decades of the twenty-first century that have witnessed an explosion of activities. There

226

´ P. MIGON

has been a surge of publications of different kinds, aimed at different readerships. In particular, many popular science publications – in the form of both guidebooks exploring the geoheritage of the entire Sudetes (Cwojdzin´ski & Kozdro´j 2007; Stachowiak et al. 2013) or specific regions within their confines (e.g. Knapik 2008; Knapik & Migon´ 2011) – and maps (e.g. Wojewoda 2013) as well as more general publications focused on interpretation rather than itineraries (e.g. Bogdan´ski 2006; Migon´ 2012), are now available. Various on-site initiatives towards geotourism are also being implemented. For example, in the Karkonosze National Park there are currently nine educational trails which either focus specifically on geoheritage or involve it as a significant component. Special access was provided in the Go´ry Stołowe National Park, by means of wooden boardwalks, to a group of selected geosites located away from marked trails. Of particular importance is a growing interest in past mining activity and several once-closed mines of coal, gold, tin and uranium, some dating back to late medieval times, are now open to the public. Perhaps the most visible sign of geoheritage promotion was the status of National Geopark being awarded to the Karkonosze National Park in 2010 (Knapik et al. 2011). All the developments outlined above have their parallels on the Czech side of the state border signified by new educational trails, information panels, open-air and indoor rock collections and museums at former mining sites, as well as a plethora of published materials. While reviewing the current developments, the fact that the foundations of some of these activities were provided 100 years (or more) ago should not be overlooked. In the sandstone regions of Stołowe Mountains and Broumovsko we use the same engineered trails as our predecessors did in the middle of the nineteenth century. In fact, a much-needed rearrangement of tourist flows on the mesa of Mount Szczeliniec Wielki was possible by reopening trails erected in the 1870s which had been closed after World War II. In the Jelenia Go´ra Basin, revitalization of Romantic granite parklands continues. Sites once popular among visitors and then forgotten are brought back into focus and their access routes repaired. Since many of these are of direct geological and geomorphological interest, it is hoped that provision of correct interpretation will follow. The geological model of the West Sudetes described earlier, devastated and forgotten, was thoroughly renovated and accompanied by new interpretation panels. It is now considered as both a geosite as well as an object of cultural heritage. At a more theoretical level, pre-1945 literature on the subject is being rediscovered. For example, the catalogue by Gu¨rich (1914) was heavily used in

studies providing the scientific background for creation of the National Geopark in the Karkonosze.

Conclusions The conclusions offered by this study are of two kinds. First, the example of the Sudetes demonstrates that interest in natural features, mainly geomorphological as these are most scenic and spectacular, and appreciation of physical landscape has a long tradition. In this respect, geotourism history in the Sudetes parallels its development in Britain (Hose 2008). It goes back for more than 200 years and involves truly spectacular achievements such as: the provision of first access to highly difficult terrains of sandstone labyrinths and ‘rock cities’ as early as in the late eighteenth century; innovative means of showing geology to the public as exemplified by the model cross-section of the West Sudetes in Jelenia Go´ra; and important publications which have stood the test of time (e.g. Gu¨rich 1914). In terms of providing access to geoheritage sites, our predecessors of a century ago were often more efficient than we are now. Many sites, once popular and scientifically important, have almost lost their value due to uncontrolled growth of secondary vegetation. We are still to arrive at a good working solution on which geosites to conserve and how. Second, and at a more theoretical level, one might ask if it is justified to use the term ‘geotourism’ to describe eighteenth, nineteenth and early twentieth century travels to the natural curiosities of the Sudetes. Much depends on how we define geotourism and whether interpretation and learning, accompanied by increasing awareness and geoconservation initiatives, are its indispensable components, as implied by most attempts to define the phenomenon (e.g. Hose 1995, 2008; Joyce 2007). The answer may be provided in three steps, going deeper and deeper into the essence of geotourism. Was the wealth of outstanding natural features, particularly landforms supported by specific rock types, recognized and promoted as key assets of the region? The answer is firmly ‘yes’ and many recreational activities of the time were performed at localities of significant geoheritage value. Was physical access to geosites upgraded and information provided for better in-depth understanding? Again the answer is generally ‘yes’, although a lag time between providing physical access (nineteenth century) and interpretation (early twentieth century), as well as a limited availability of the latter, should be noted. Finally, a key question is did tourists appreciate, learn about and better understand the geoheritage of the region as a consequence of their visits? Honestly, we do not know. This factor of

REINVENTING GEOTOURISM IN THE SUDETES

huge uncertainty surrounds the entire contemporary geotourism, which works hard towards excellence in provisions but has little means by which to evaluate the experience gained. My participation at the ‘Appreciating Physical Landscapes: Geotourism 1670– 1970’ conference as well as part of my research in this field were supported by the statutory fund of the Department of Geography and Regional Development. I thank J. Potocki for his comments on the tourism history part of the paper. Editorial advice from T. Hose was greatly appreciated.

References Badura, J., Pe´cskay, Z., Koszowska, E., Wolska, A., Zuchiewicz, W. & Przybylski, B. 2005. New age and petrological constraints on Lower Silesian basaltoids, SW Poland. Acta Geodynamica et Geomaterialia, 2, 7– 15. Bogdan´ski, J. 2006. Geologiczny raj Krainy Wygasłych Wulkano´w [Geological Paradise of the ‘Land of Extinct Volcanoes]. Stowarzyszenie Kaczawskie, Ms´ciwojo´w (in Polish). Chmal, H. 2010. Johann Wolfgang Goethe (1749– 1832) – przyrodnik [Johann Wolfgang Goethe (1749– 1832) – a naturalist]. Czasopismo Geograficzne, 81, 173– 185 (in Polish). Cwojdzin´ski, S. & Kozdro´j, W. 2007. Sudety. Przewodnik geoturystyczny [The Sudetes. Geotourist guide]. Pan´stwowy Instytut Geologiczny, Warszawa (in Polish). Engel, Z., Ny´vlt, D., Krı´zˇek, M., Treml, V., Jankovska, V. & Lisa, L. 2010. Sedimentary evidence of landscape and climate history since the end of MIS 3 in the Krkonose Mountains, Czech Republic. Quaternary Science Reviews, 29, 913– 927. Geyer, M., Bissig, G., Maul, G., Meissner, M., Peterek, A., Pustal, I. & Ro¨hling, H.-G. 2007. Goethe und die Geologie – Ein geotouristisches Nutzungskonzept zu den geologischen Betrachtungen in den Schriften Johann Wolfgang von Goethes. Schriftenreihe der deutschen Gesellschaft fu¨r Geowissenschaften, 51, 61– 66. Grocholski, W. (ed.) 1969. Przewodnik geologiczny po Sudetach [Geological guidebook to the Sudetes]. Wydawnictwa Geologiczne, Warszawa (in Polish). Gu¨rich, G. 1914. Die geologischen Naturdenkma¨ler des Riesengebirges. Beitra¨ge zur Naturdenkmalpflege, 4, 141–324. Hall, A. M. & Migon´, P. 2010. The first stages of erosion by ice sheets: evidence from central Europe. Geomorphology, 123, 349– 363. Hose, T. A. 1995. Selling the story of Britain’s stone. Environmental Interpretation, 10, 16– 17. Hose, T. A. 2008. Towards a history of geotourism: definitions, antecedents and the future. In: Burek, C. V. & Prosser, C. D. (eds) The History of Geoconservation. Geological Society, London, Special Publications, 300, 37–60. Imlauf, L. & Bohadlo, S. 1997. Adrsˇpach – Teplice. Skalnı´ labyrint [Adrsˇpach – Teplice. Rock Labyrinth]. Gate, Na´chod (in Czech).

227

Jaworski, K. 1988. Turys´ci polscy w Teplicko-Adrsˇpachskich Skałach w XVIII i XIX wieku [Polish tourists in the Teplice-Adrsˇpach Rocks in the 18 and 19C]. In: Tomczak, M. (ed.) Poutnici. Informator Krajoznawczy. SKPS, Wrocław, 9 –62 (in Polish). Joyce, B. 2007. Geotourism, Geosites and Geoparks: working together in Australia. The Australian Geologist, September, 26– 29. Knapik, R. 2008. Przewodnik geoturystyczny po Karkonoskim Parku Narodowym [Geotourist Guide to the Karkonosze National Park]. Karkonoski Park Narodowy, Jelenia Go´ra (in Polish). Knapik, R. & Migon´, P. 2011. Atlas. Georo´z˙norodnos´c´ i geoturystyczne atrakcje Karkonoskiego Parku Narodowego i otuliny [Atlas. Geodiversity and Geotourist attractions of the Karkonosze National Park and its Buffer Zone]. Karkonoski Park Narodowy, Jelenia Go´ra (in Polish). Knapik, R., Migon´, P., Szuszkiewicz, A. & Aleksandrowski, P. 2011. Geopark Karkonosze – georo´z˙norodnos´c´ i geoturystyka [Karkonosze Geopark – geodiversity and geotourism]. Przegla˛d Geologiczny, 59, 311 –322 (in Polish). Mazurski, K. R. 2012. Historia turystyki sudeckiej [The History of Tourism in the Sudetes]. Wierchy, Krako´w (in Polish). Migon´, P. 2008. High-mountain elements in the geomorphology of the Sudetes (Bohemian Massif) and their significance. Geographia Polonica, 80, 101–116. Migon´, P. 2011. Geomorphic diversity of the Sudetes – Effects of global change and structure superimposed. Geographia Polonica, 84 (Special Issue Part 2), 93–105. Migon´, P. 2012. Granitowy krajobraz Kotliny Jeleniogo´rskiej – dopełnienie Geoparku Karkonosze [Granite landscape of the Jelenia Go´ra Basin – a complement to the Karkonosze Geopark]. Przegla˛d Geologiczny, 60, 528 –533 (in Polish). Migon´, P. & Latocha, A. 2008. Enhacement of cultural landscape by geomorphology. A study of granite parklands in the West Sudetes, SW Poland. Geografia Fisica e Dinamica Quaternaria, 31, 195 –203. Migon´, P. & Latocha, A. 2010. Zro´z˙nicowanie abiotycznych elemento´w s´rodowiska i jego wykorzystanie w rozwoju funkcji turystycznej regionu sudeckiego [Geodiversity and its role in the development of tourism in the Sudetes]. In: Ciok, S. & Migon´, P. (eds) Przekształcenia struktur regionalnych – aspekty społeczne, ekonomiczne i przyrodnicze. Instytut Geografii i Rozwoju Regionalnego, Uniwersytet Wrocławski, Wrocław, 397 –417 (in Polish). Migon´, P. & Latocha, A. 2013. Human interactions with the sandstone landscape of Central Sudetes. Applied Geography, 42, 206– 216. Pabel, F. 1843. Kurze Geschichte der Bekanntwerdung und Anlagen-Einrichtung der Heuscheuer. W.W. Klambt, Neurode. Panizza, M. & Coratza, P. (eds) 2012. Il ‘Viaggio in Italia’ di J.W. Goethe e il paesaggio della geologia. ISPRA, Roma. Potocki, J. 2004. Rozwo´j zagospodarowania turystycznego Sudeto´w od połowy XIX wieku do II wojny s´wiatowej [The Development of Tourist Infrastructure in

228

´ P. MIGON

the Sudetes Since the mid-19C till the End of World War II]. Plan, Jelenia Go´ra (in Polish). Pulinowa, M. Z. 1989. Rzez´ba Go´r Stołowych [Relief of the Stołowe Mountains]. Prace Uniwersytetu S´la˛skiego, Katowice (in Polish). Reicherter, K., Froitzheim, N. et al. 2008. Alpine Tectonics north of the Alps. In: McCann, T. (ed.) The Geology of Central Europe. Volume 2: Mesozoic and Cenozoic. Geological Society, London, 1233– 1285. Sachanbin´ski, M. 2005. Minerały Karkonoszy i ich najbliz˙szego sa˛siedztwa [Minerals of the Karkonosze and their close vicinity]. In: Mierzejewski, M. P. (ed.) Karkonosze. Przyroda nieoz˙ywiona i człowiek. Wydawnictwo Uniwersytetu Wrocławskiego, Wrocław, 161– 260. Stachowiak, A., Cwojdzin´ski, S. et al. 2013. Geostrada sudecka – przewodnik geologiczno-turystyczny

[Sudetic Geostrada – Tourist and Geological Guide]. Pan´stwowy Instytut Geologiczny – Pan´stwowy Instytut Badawczy, Warszawa (in Polish). Szczepan´ski, E. 1989. Towarzystwo Karkonoskie (1880– 1945) [Karkonosze Society (1880– 1945)]. S´la˛ski Labirynt Krajoznawczy, 1, 75– 86 (in Polish). Traczyk, A. & Migon´, P. 2003. Cold-climate landform patterns in the Sudetes. Effects of lithology, relief and glacial history. Acta Universitatis Carolinae, Geographica, 35 (Supplementum 2000), 185–210. Wojewoda, J. 2013. Mapa geoatrakcji krainy Go´r Stołowych i Broumovskich steˇn [Map of geo-attractions of the Stołowe Mountains and Broumovske´ steˇny]. Plan, Jelenia Go´ra. Zuchiewicz, W., Badura, J. & Jarosin´ski, M. 2007. Neotectonics of Poland: an overview of active faulting. Studia Quaternaria, 24, 5 –20.

Appreciating loess landscapes through history: the basis of modern loess geotourism in the Vojvodina region of North Serbia DJORDJIJE A. VASILJEVIC´*, SLOBODAN B. MARKOVIC´ & MIROSLAV D. VUJICˇIC´ Department of Geography, Tourism and Hotel Management, Faculty of Sciences, University of Novi Sad, Trg Dositeja Obradovic´a 3, 21000 Novi Sad, Serbia *Corresponding author (e-mail: [email protected]) Abstract: Loess is wind-blown sediment that covers extensive areas in the middle latitudes. Much of the loess in Eastern and Central–Eastern Europe has been redeposited by the River Danube and its tributaries. The case study area (Vojvodina region) encompasses the confluence area of the Danube, Sava and Tisa rivers. This region includes the most complete and the thickest loess– palaeosol sequences found in Europe, a valuable record of palaeo-climates over the past two million years only recognized in the closing decades of the twentieth century. Long before then however, several enthusiasts, engineers and travellers recognized and appreciated loess as a significant natural phenomenon. Among them was Luigi Ferdinando Marsigli (1658–1730) who gave the first scientific description of European loess in his outstanding multivolume work Danubius Pannonico Mysicus where he drew and explained notable loess–palaeosol exposures along the Danube River. Many other loess observations were also recorded by a number of international travellers, whose illustrated travelogues (mainly published in the nineteenth century) mentioned and illustrated loess observations along the Danube and its major tributaries. This chapter explores the interplay of scientific loess research and its geo-historical literary aspects as the foundations of modern loess geotourism.

If not the largest European river, the Danube is certainly the most important because it has always had an immense role in merging different nations, cultures, traditions, customs and, above all, regions. Moreover, navigation on the Danube provides an adventurous and exploratory voyage, which is especially the case within the territory of today’s Serbia which was previously incorporated within many great empires such as the Roman, Ottoman and Austro-Hungarian. This cultural diversity, combined with the beautiful and picturesque scenery of Danubian riverbanks and surroundings, has been an inspiration to artists, poets and musicians and also writers, travellers, naturalists and scientists. While navigating this river, early explorers transferred their travel experiences into the written word in the form of notes, diaries, logs and other literary forms, often with pictures or gravures, in which they made their comments, opinions and impressions about the surrounding area, structured according to the chronology of their journeys. This travel literature (as it was subsequently defined) became an important medium for broadening people’s horizons (Klemun 2007) and was the fundamental method for wealthier people (the bourgeoisie) to acquire world knowledge (Stagl 1995). This was especially the case from the end of the eighteenth to the end of the nineteenth centuries when people hoped to educate themselves through long-distance and long-lasting journeys,

for which the Danube provides a perfect itinerary. Brenner (1990) describes how there was an explosion of travel literature throughout this period, which occurred in many different areas of culture and knowledge (Klemun 2007). Most publications made some attempt at analysis related to people and their physical appearance, nationality, education level, business and trading, settlements, buildings and tradition. These publications and artworks also included nature and landscape shaped by their view from the river, or sometimes the land on trips ashore. In the context of nature, this chapter will involve only the abiotic (non-living) part of the natural surroundings which in modern literature has been defined as geodiversity (Gray 2004). Among many values of geodiversity, the most recognized and appreciated by the general public is the visual or aesthetical value (Hose 2010). This quality is therefore the starting point for geodiversity appreciation in the beginning as ‘it refers quite simply to the visual appeal provided by the physical environment’ (Gray 2004, p. 81). Accordingly, this chapter describes certain publications and travellers who admired the banks of the Danube within the territory of the Autonomous Province of Vojvodina (north Serbia), with particular reference to loess profiles and sequences as significant parts of geodiversity, the best developed of their kind in this area of Europe.

From: Hose, T. A. (ed.) 2016. Appreciating Physical Landscapes: Three Hundred Years of Geotourism. Geological Society, London, Special Publications, 417, 229–239. First published online November 3, 2014, http://dx.doi.org/10.1144/SP417.5 # The Geological Society of London 2016. Publishing disclaimer: www.geolsoc.org.uk/pub_ethics

230

´ ET AL. D. A. VASILJEVIC

Nothing less than loess The main ‘objects of admiration’ in this chapter are loess–palaeosol sequences or simply loess as part of Earth’s geodiversity (Vasiljevic´ et al. 2011a, b, 2014; Solarska et al. 2013; Dong et al. 2014). This geological resource provides wind-blown sediments which cover extensive areas in the middle latitudes, affecting 20% of Europe’s land surfaces (10% of the world’s land surfaces). Chronologically, the formation of loess deposits is related to the glacial cycles during Pleistocene time when dusty material was transported and deposited by wind during glacial periods. During warmer (interglacial) periods, deposition was affected by various climatic conditions associated with higher temperature, higher precipitation and an increase in chemical and biological activity (Vasiljevic´ et al. 2011b), which results in soil formation. In a scientific sense, loess has gained great importance in the reconstruction of past climates since the relatively recent pioneering works of Kukla (1977) and Heller & Liu (1982). It was proved by these authors that loess was formed during glacial periods, when climates were cold and dry, and is interlayered with soil bodies, representing the shift from cold and dry to warm and wet climatic conditions (e.g. Kukla 1977; Pecsi 1990). Loess –palaeosol sequences are therefore interesting geological formations which describe the Earth’s climate history over many glacial and interglacial cycles, or even geological ages, and represent important terrestrial palaeoclimatic and palaeoenvironmental archives (Vasiljevic´ et al. 2014). Although considered as ‘soft rock’ or even soil sediment (Derbyshire 2001), loess areas provide unique, attractive and appealing landscape: deep loess gorges; wavy plateaus; steep and high loess profiles; pseudokarst landforms such as loess caves and sinkholes; and dry valleys, gullies or pyramids (Lukic´ et al. 2009; Vasiljevic´ et al. 2014). These places attracted travellers of the past and continue to attract modern landscape tourists (today defined by Hose 2008a, 2010 as geotourists) and naturelovers and even inspire artists (Gray 2004; Hose 2008a; Vasiljevic´ et al. 2011a). The most complete and thickest loess deposits are found in plains (e.g. Pampean and Russian), plateaus (e.g. Chinese Loess) and along river basins (e.g. Danube, middle Rhine, Mississippi and middle Yellow River). The geographical setting of this study, the Vojvodina region, is situated in the southeastern part of the Carpathian (Pannonian) Basin and encompasses the confluence area of the Danube, Sava and Tisa rivers where loess and loesslike sediments cover almost two-thirds of the region, including six loess plateaus (Fig. 1). Unlike other European loess–palaeosol sequences which

generally formed during Late Pleistocene time and were characterized by smaller total thicknesses, the loess in the Vojvodina region is significantly older and thicker. This amazing accumulation of glacial dust, intercalated with interglacial palaeosols, preserves the most complete European terrestrial environmental record of the last million years (Markovic´ et al. 2011).

Danube (steam) voyages: new inspiration for travellers and writers As the loess–palaeosol sequences in the Vojvodina region are best revealed and most visible at steep cliffs of riverbanks, the best way to observe and admire them is from the river. Concurrently, the easiest and fastest method of travel in the nineteenth century was by long, deep and wide rivers, which is certainly how the Danube can be characterized. The most important events in the development of Danube cruises overlapped with the industrial revolution of the nineteenth century. In 1813 the Austrian government began using steamboats, which were a very successful and cost-effective means of river transport. Early Danube cruises did not however meet with much success (as the companies did not possess suitable levels of organization and infrastructure) until two Englishmen (Andrews and Prichard) founded the First Danube Steamship Company (Erste Donau-Dampfschiffahrts-Gesellschaft in German, or DDSG) in 1829. This company was able to meet the Austro-Hungarian government’s expectations with the steamship Franz I and started travelling the route between Vienna and Budapest as early as 1830. As of this time, the history of Danube navigation is more or less identical, coincident with the history of the DDSG. This innovation in river transport enabled people of that time to travel ‘from Black Forest to Black Sea’ (see Millet 1893) and visit vast parts of the continent, connecting western and eastern parts of Europe. Most of the Danube cruises in the nineteenth century had an initial embarkation point in Vienna; some even finished downstream at the orifice of this gigantic river or even continued to several ports of the Black Sea. Some voyages began in Passau, from where the Danube was navigable by the ships of that time. The voyages provided engaging visual experiences, passing through several countries, capital cities, fortresses, castles, gorges, forests and other natural and human-made attractions. With the comparative luxury and comfort onboard (sometimes much better than could be found in the accommodation along the riverside), these cruises launched a new trend in pan-European travel. Historians of cultural–scientific travel generally agree that, for the aspiring bourgeoisie hoping

LOESS LANDSCAPES HISTORY, NORTH SERBIA

231

Fig. 1. Map of locations and areas referred to in the chapter.

to educate themselves, travelling became a central method of acquiring world knowledge (Stagl 1995; Klemun 2007), much as the aristocracy on their Grand Tour. As a consequence, there was an explosion of travel literature to support the new class of travellers, describing many aspects of Danubian regions (people, places, nature, traditions, clothes, customs, food, drinks, anecdotes and monuments). The authors were established journalists, poets, artists, historians, geographers and sometimes scientist of various disciplines. Additionally, these writers were of British, Irish, German, Austrian and even American origin, and brought diverse aspects to their writing in terms of attitude and approach to the travelled surroundings. All this resulted in a great variability of publications and artistic works (paintings, gravures, sketches and photographs). These publications and works were used as manuals and guides by people contemplating travel in this part of Europe, whether by the river or by other means of transport. At the end of the nineteenth century (1895), DDSG advertised in the most eminent newspapers of Vienna their Danube Guide (Donau Fu¨hrer in German). The advertisement informed readers of a

new booklet entitled Danube from Passau to the Black Sea (Die Donau von Passau bis zum Schwarzen Meere in German), which contains descriptions of Danubian banks in the formerly mentioned area, the list of available ships and general travel instructions and information. This was the first printed guidebook for Danube travel created for all people who travelled or intended to travel this river. The guide’s first print run was 25 000 copies, distributed free of charge in DDSG offices in Vienna and Budapest. It was later reprinted for more than four decades (Kostic´ 2012).

Loess writing by Danube travellers in the 19th century The abovementioned writers described both their natural surroundings and the manmade environment in equal measures. Although they were mostly writers, journalists or artists, their observations and descriptions of the surrounding topography and landscape were comprehensive, detailed and surprisingly precise. There is quite a concordance among them that the upper Danube’s banks from Ulm (Germany) to Vienna (Austria) are much more

232

´ ET AL. D. A. VASILJEVIC

interesting before entering the vast Pannonian Plain. The monotonous and flat relief of this fertile plain did not exalt these travellers, although its flooded region provides immense areas of wetlands and forests with great scenery and biodiversity, today designated nature protected areas (e.g. in Serbia there are Special Nature Reserves Gornje Podunavlje and Karadordevo and Nature Parks Tikvara and Begecˇka jama).

From vast flatness to charming hills The flat and monotonous scenery of the Danubian surroundings of the Pannonian Plain changes drastically after Mohacs (Hungary), which is today the last big settlement in Hungary before the Danube enters the western border of the Vojvodina region, where it continues to flow as the border between Croatia and Serbia. This is the area where the Danube slows down as it reaches the Frusˇka Gora Mountain and changes its course towards the east, cutting into its northern foothills. Here, loess accumulation and sedimentation is of greatest thickness, which resulted in the creation of plateaus, gullies, steep cliffs and even pseudokarst landforms. The first to give a geographical description of the area where the Danube drastically starts to change in scenery was talented Irish writer and journalist Michael J. Quin (1796–1843), who wrote: ‘ . . . about two hours after we left Mohacs . . . the banks indeed were still low and sandy, which detracted from its beauty. In the distance on the right, a sugarloafed mountain, rising above the summit of a range of hills, indicated an approaching change of scenery . . . ’ (Quin 1835, p. 44). This enthusiastic description greatly resembles the introduction to a geotourist brochure of the Frusˇka Gora Mountain written in an old-fashioned style. A similar observation was made by Johann Georg Kohl (1808– 1878; Fig. 2a), a German travel writer, historian and geographer, whose perception was brief but concise: ‘ . . . below Mohacz we entered a more interesting part of Danube’ (Kohl 1844, p. 256). On the other hand, his observations throughout the whole voyage were geographically based and precise, which was the case with the flatness of the Pannonian part of Vojvodina: ‘the parallelogram lying between the Danube and the Theiss, a desert level district . . . this is the celebrated Batshka . . . ’ (Kohl 1844, p. 256). Following Quin’s enthusiasm by both taking in and writing about this picturesque scenery was Andrew Archibald Paton (1811–1874; Fig. 2b), a Scottish diplomat, writer and traveller who travelled upstream from the Black Sea. As he spent several years in the Middle East, he travelled from Beirut through the Greek islands, Turkey, Black Sea to port Varna in Bulgaria. From there the voyage was continued on the roads to Ruse, a Bulgarian port on the

Danube. He was very anxious to visit as ‘ . . . Vienna and Pesth [today’s Budapest] offered no attractions in the month of August’ and he had great expectations as he ‘ . . . felt impatient to put in execution my long cherished project of travelling through the most romantic woodlands of Servia’ (Paton 1845, p. 21). Paton revealed his desire to visit this area and enjoyed drastic changes in scenery as he continues: ‘ . . . the right shore soon becomes somewhat bolder, and agreeably wooded hills enliven the prospect. This little mountain chain is the celebrated Frusca Gora, the stronghold of the Servian language, literature, and nationality . . . ’ (Paton 1845, p. 21). This kind of travel culture certainly makes Quin and Paton potential geotourists with an interest in landscape, as they showed admiration of and curiosity towards many aspects of the visited area, giving them characteristics of the modern tourist. Several years after these authors, an analogous judgement was made by Isaac Salomon Kapper (1821–1879; Fig. 2c), a Bohemian-born Austrian writer and medical doctor (with a PhD degree from Vienna University) who used the literary pseudonym of Siegfried Kapper. Among his many excellent fairytales and poems (written both in German and Czech), he wrote a book entitled Voyages throughout the South Slavic countries in Summer 1850 (Su¨dslavische Wanderungen im Sommer 1850 in German) with the following impressions on this particular area and loess landscape: ‘ . . . steamboat Albrecht apart from Cˇerevic´ [small village on the right Danube bank, Figure 1] . . . even earlier, right bank gradually started to change its monotonus ˇ erevic´ those hills almost landscape . . . downstream C unnoticeably changed into cut-edged mountains that cover south parts of Batshka [today Bacˇka] and Srem districts . . . On the other side, left bank is still flat . . . ’ (Kapper 1851, p. 1). Although not an expert in geosciences, Kapper provided excellent depictions of the surrounding landscape, and certainly showed substantial appreciation for the varying morphological processes and geodiversity. A similar, but not as geoscientific, observation was made by Walter Copeland Jerrold (1865– 1929), an English writer, biographer and newspaper editor (deputy editor of The Observer) whose literary works earned him a portrait in the National Portrait Gallery in London (Fig. 2d). Among his many biographies and travel books, in his masterpiece Danube from 1911 he gave his impressions: ‘While on the left side the country continues flat, on the right are low cliffs. Ujlak [Today Ilok is a Croatian town on the right Danube bank with beautiful loess cliffs; Figs 1 and 3], on a rocky height, one of the most picturesquely situated places along this part of the river . . . as the river nears the foot of the hills on the right, the scenery becomes more attractive and more varied’ (Jerrold 1911, pp.

LOESS LANDSCAPES HISTORY, NORTH SERBIA

233

Fig. 2. Portraits of the authors referred to in the chapter: (a) Johann Georg Kohl; (b) Andrew Archibald Paton; (c) Siegfried Kapper; (d) Walter Copeland Jerrold; (e) John Paget; and (f) Luigi Ferdinando Marsigli.

231 –232). Similar to Kapper, his observations and descriptions of loess and landscape verifications were impressive, while his rapture for scenery makes him a potential geotourism writer and traveller. Jerrold’s voyage continued: ‘ . . . soon is seen ahead the famous fortress of Petervarad . . . ’ (Jerrold 1911, pp. 231 –232). The Petrovaradin fortress (at that time Petervarad or Peterwardein) is a major Austro-Hungarian military complex and one of the most impressive historical buildings on the Danube outside Budapest. It is located on the northern foothills of Frusˇka Gora Mountain opposite Novi Sad, which was then just a small market town in wetlands but is now the capital of Vojvodina province. As such, Petrovaradin was a popular stopping point for most ships. This was the case for John Paget (1808–1892; Fig. 2e), English writer, agriculturist and landholder who travelled widely in south-eastern Europe, especially Hungary and Transylvania, as he had married a Hungarian baroness and settled in Transylvania where he promoted

scientific agriculture. Paget took part in the 1848– 1849 Hungarian revolution as he was a friend of Count Istva´n Sze´chenyi, the Hungarian politician, theorist and writer, one of the greatest statesmen of Hungarian history. This gave him an influential position in local society and opportunities to travel and write. One of his most popular destinations was today’s Serbia, which he described: ‘The hills on the Servian side now became exceedingly pretty. They are not generally high, but nothing can be imagined more perfectly wild and picturesque. They are covered, down to the very water’s edge, with a low natural wood’ (Paget 1855, pp. 26 –27). His publications provided thorough and very accurate descriptions of Vojvodina region and its scenery, and he also wrote: ‘ . . . a low range of hills accompanied us along the west bank for some distance; and the openings which they sometimes present, disclosing their green valleys, and silver streams, and white-washed cottages, and fantastic steeples, are most beautiful’ (Paget 1855, pp. 87 – 88). His wonderment at the hilly landscape is no

234

´ ET AL. D. A. VASILJEVIC

Fig. 3. Illustration of Danube loess landscape near Ilok by William Beattie; illustration on wood, see Beattie (1844).

surprise as he paid special attention to viticulture during his agricultural research and practice. Loess hills are considered to be favourable places for vineyards, such as those of Western Iowa (USA). This comprehensive author and researcher is therefore also profiled as having geotouristic concern and interest. Besides this work, his diary (in six volumes) remains in the Sze´clienyi library of Hungarian National Museum. The first five volumes contain only observations of natural history made in France, Italy, Switzerland, Austria and Hungary. The flatness of the plain and sudden change in landscape of the downstream Danube was not unnoticed by travellers from other continents. Theodore Child (1846–1892), American journalist and art critic with connections to the famous painter James Abbott McNeill Whistler (1834– 1903), had quite critical impressions of the vast Pannonian Plain. While he ‘ . . . was on deck betimes to inspect the landscape’, ‘the flatness of the surrounding country, the vastness of the river, whose banks rise scarcely a foot above the level of the yellow water . . . all tend to make this part of the Danube very monotonous . . . ’ (Child 1889, pp. 1– 3). Moreover, Child recognized the considerable change in

scenery on the right bank side to the west, as he wrote: ‘After passing Vucovar [today Vukovar, Fig. 1] the scenery became less monotonous, and soon we entered a smiling, hilly country, fertile and beautiful as a vast park . . . ’ (Child 1889, p. 6). It is clear that all these authors made similar observations regarding the gradual alternations of scenery and topography when entering the Vojvodina region on this gigantic river. After the neverending vastness of the Pannonian Plain, they found steep loess profiles and hills exceptionally refreshing and appealing. Their precise and enthusiastic description of the surrounding landscape shows that this area did (and still does) have geotouristic potential.

Early beginnings of scientific loess admiration Roughly one century before these travellers noticed and described loess landscapes, the Italian officer, military engineer, naturalist, diplomat and writer Count Luigi Ferdinando Marsigli (1658–1730; Fig. 2f) gave the first scientific description of loess in Vojvodina. This polymath, scientist and founder of the Academy of Science of Bologna had a

LOESS LANDSCAPES HISTORY, NORTH SERBIA

multidisciplinary education and interests that brought him designation as geographer, geologist, cartographer, astronomer and even zoologist. During the last decade of the seventeenth century, Marsigli was employed to survey the boundaries between the Ottoman and Habsburg empires. As a high-ranking officer of the Austrian army he spent a lot of time in the Petrovaradin, Titel and Stari Slankamen fortresses, which are very close to significant loess deposits (Markovic´ et al. 2009). These notable loess–palaeosol exposures along the Danube river valley were drawn and analysed by him in his outstanding six-volume work Danubius Pannonico-mysicus (Danube in Pannonian and Moesian Region, see Marsigli 1726; Fig. 4a) for which he conducted thorough research on different aspects of geosciences including cartography, astronomy (for calculating longitude), geology, mining, hydrology and natural history. This publication included descriptions of the lithology of the loess banks of the Danube river showing modern soil (Terra fructifera pinguis nigra et creatacea or black fertile carbonate soil), palaeosol (Terra nigra fructifera pinguis or black fertile soil) and between them loess layers (Terra lutosa cinerive et in fragmento creatacea priabilis or yellowcinerary/ashy layer with carbonate fragments, i.e. concretions). It is evident from modern studies that many of the characteristics and localities of the loess– palaeosol sequences recognized by Marsigli remain valid to this day (Markovic´ et al. 2009) and that these areas remain sites for potential modern geotourism developments. What is interesting is that none of the nineteenthcentury Danube travellers mentioned the Titel loess plateau in their works, although this isolated loess ‘island’ located at the confluence area of Tisa and Danube rivers represents a unique geomorphologic phenomenon. Accordingly, Marsigli conducted thorough investigations and cartographic studies of this area (Fig. 4b). The plateau dominates its lower surroundings which make it aesthetically and visually attractive, in addition to the rich diversity of the loess landforms such as loess caves, pyramids, gullies and cliffs (Vasiljevic´ et al. 2011b). It took 100 years from Marsigli’s work to the first characterization of loess deposits and introduction of the term ‘loess’ by the German mineralogist Karl Caesar von Leonhard (1779–1862) in his book Charakteristik der Felsarten (Characteristics of Rocks) in 1823/24. Von Leonhard (1824, p. 722) described loess as: ‘muddy, dirty yellowishgray, earthy mixture of clay, limestone and gravel and a small amount of mica; a friable mass; a comminuted granular system’ (Jovanovic´ et al. 2014). According to Smalley et al. (2001) the ‘locus typicus’ for loess is Haarlas. The River Neckar in Germany is where Von Leonhard probably coined

235

the word ‘loess’ by borrowing one of the local names ‘Loesch’ used by country people from the Upper Rhine region to denote the yellow lime soil ‘Lischen’ or the so-called ‘snail-shell soil’ which is easily loosened by plough or spade. This late determination of loess terminology was evidently the reason for its absence from the travel books of the nineteenth century despite many very precise descriptions. While on a steamer cruise from Budapest to Constantinople, the aforementioned Theodore Child noticed that they were passing ‘ . . . in the midst of a very broad stream of dirty yellow water, flowing with a swift rippling movement between banks of brown, clayey earth, eaten away by the wash and crumbling visibly into the river . . . ’ (Child 1889, p. 1). Certain geological components evidently caught his attention and inspired this precise description and analysis. One of the nineteenth-century Danube travellers most interested in loess and geosciences was Francis Davis Millet (1848–1912, Fig. 5a), American painter, sculptor and writer who was close friends with the famous sculptor Augustus SaintGaudens and the writer Samuel Langhorne Clemens (The latter is better known by his pseudonym Mark Twain, who was even his best man). As a notable person of artistic and adventurous spirit, Millet liked to travel a lot in the high-class company of distinguished writers, artists and scientists. During one of his visits to Europe he travelled in the company of Alfred William Parsons (1847–1920, Fig. 5a), a famous English illustrator, landscape painter and garden designer who illustrated the Danube and surroundings (Fig. 5b) in From Black Forest to the Black Sea. His interest in geosciences is demonstrated at the beginning of the book where he gives several geographical descriptions such as: ‘A mile and a half below Donaueschingen the Brigach and the Brege join, and the stream here receives the name of the Danube’ (Millet 1893, p. 2). Furthermore, along the Danube his observations of loess banks in the Vojvodina region were very skilful and precise, with some astonishing details of its geological and geomorphological features. Millet’s impressions of this region were very positive as he wrote that ‘ . . . rude steps cut along a cleft were lively with girls carrying jars of Danube water to the village above and once, under a vineyard, where the vines trail over the very edge of the bank . . . ’ (Millet 1893, p. 161). As he continues the journey, somewhere ‘ . . . about half-way between Apatin and the village of Erdod [today Erdut in Croatia, Fig. 1]’ he spotted ‘. . . fertile hills now skirt the west bank, and their sunny yellow slopes looked agreeably bright and warm’ (Millet 1893, p. 161). Further on, his remarkable observational skills for geological features of loess is proven as he noticed that

236 ´ ET AL. D. A. VASILJEVIC Fig. 4. (a) Book cover of Danubius Pannonico Mysicus and (b) the map of the confluence area of Tisa and Danube rivers drawn by Luigi Ferdinando Marsigli (1726), representing Danubian loess landscape and Titel loess plateau.

LOESS LANDSCAPES HISTORY, NORTH SERBIA

237

Fig. 5. (a) Francis Davis Millet (right) and Alfred William Parsons (left) sketching Danube banks and (b) the cover page of the same publication (Millet 1893) that (it is assumed) represents Gaelic goddess Abnoba giving birth to Danube. She was worshipped in the Black Forest, who was interpreted to be a forest and river goddess.

‘the river has washed away the hills into perpendicular bluffs, which are of earth almost as hard as sandstone . . . ’ (Millet 1893, p. 161), as dry loess often has the structure similar to rock, and even certain pseudokarst forms. Throughout the book, Millet did not hide his admiration of the Danube and its surroundings. After leaving the camping ground near today’s

Sremski Karlovci, an important Serbian cultural and historical centre, ‘on the perfect summer morning . . . ’ they ‘ . . . drifted down into the silvery light of morning which glorified the river, the hills, and the distant landscape, we were in the mood to enjoy exactly what the Danube offered for our entertainment . . . ’ (Millet 1893, p. 170). Here Millet’s geological observation continues as they reached

Fig. 6. Stari Slankamen and loess landscape drawn from the Danube by Alfred William Parsons (Millet 1893).

238

´ ET AL. D. A. VASILJEVIC

‘ . . . O Szlankamen [today Stari Slankamen, Figs 1 and 6] to the east a long range of flat hills now appeared, marking the course of the sluggish Theiss, and on the opposite bank we saw great rocks, scarcely distinguishable from the hard mud bluffs, but marking a distinct geological change in the landscape . . . ’ (Millet 1893, p. 172). Millet’s publication and his detailed and factual description of geodiversity, its range, features and values, provides readers with a true geo-travel guidebook. Unfortunately, his last voyage was one of the best known and most ill-fated in the history of cruises, as he died in the sinking of the RMS Titanic on 15 April 1912. Admired for his hard work and kindness, he was last seen helping women and children into lifeboats. Almost two centuries after Marsigli, numerous Austrian and Hungarian geologists (e.g. A. Koch, H. Wolf, Gy. Halava´ts, J. Cholnoky) initiated a systematic investigation of Quaternary sediments in Vojvodina at the end of the nineteenth and beginning of the twentieth centuries. At the beginning of the second half of the twentieth century, the research of many geologists and geomorphologists from the former Yugoslavia (e.g. V. Laskarev, B. Bukurov, M. Zeremski and J. Markovic´-Marjanovic´) sparked the interest of many international scientists (e.g. J. Butrym, H. Maruszczak and A. Bronger). During the 1960s Julius Fink of the University of Vienna founded the INQUA (International Union for Quaternary Research) Loess Commission to assemble a loess stratigraphy for all of Europe. Of the more than 100 sites indicated in the original plan, 28 of these were in the Danube region (Markovic´ et al. 2009). All of this resulted in the loess profile in Stari Slankamen gully becoming the first protected geosite in the former Yugoslavia, when proclaimed a Monument of Nature in 1975. This task of loess research, conservation and promotion was taken over by the Loess Research Group of the University of Novi Sad from the late 1990s, who will continue to follow in the footsteps of their predecessors towards better appreciation of loess at the international level.

Concluding remarks Geotourism is considered to be a contemporary special-interest form of tourism, defined and expanded in the 1990s (Hose 1995). Even though this is still a new concept of travel and leisure, it has been proven that geodiversity attracted and impressed travellers from much earlier periods. As demonstrated here, most of the nineteenth-century travellers and writers were considerably focused on natural wonders, even more than on art and architecture (in contrast to modern guidebooks;

Hose 2008b). These were mainly gentlemen of considerable social and economic standing who were in the position to travel often and ‘trod in the footsteps of earlier travellers who directed followers though assorted publications commonly focused on agricultural, industrial and socio-economic reportage’ (Hose 2008b, p. 2). Accordingly, these publications were ‘simply souvenirs of summer holidays which the author has ventured to reprint in the hope that they may find favour in the eyes of the travelling public, and also of the public that is content to travel in an arm-chair by the fireside’ (Child 1889, preface). On the other hand, these authors initiated a change in the social climate that influenced and encouraged travellers to employ their intellectual ability to venture into and explain the geology of untamed areas; this additionally involved emotional reflection on landscape and its evocation in both art and literature (Hose 2008b). As a specific geological resource and unique landscape feature, loess has attracted, and continues to attract, the attention of numerous travellers who have set their impressions down in words and on canvas. This general notion is reflected through the drastic change in scenery which considerably enhances the monotonous landscape of the Danubian surroundings: steep cliffs, picturesque hills and colourful loess sections, best visible from the water, vie for the attention of these notable authors. Although almost none of these early travel writers were geoscience professionals, their descriptions and overall interest profiled them as true ambassadors of landscape appreciation. Their enthusiasm towards landscape has surely created the basis of geotourism, not only within the Danube region but on a global scale. It is hoped that these academic and artistic depictions of outstanding topographic features over the past 300 years will bring a new dimension and motive for modern researchers and travellers. The segments from their work show that travellers recognized the aesthetic beauty and importance of this geological resource several centuries ago. This research was supported by the Ministry of Education, Science and Technological Development, Republic of Serbia (grant 176020). The authors are grateful to T. A. Hose for useful comments, D. Vujicˇic´ for providing valuable publications, and U. Stankov for providing the base for the location map.

References Beattie, W. 1844. The Danube: Its History, Scenery, and Topography. George Virtue, London. Brenner, P. J. 1990. Der Reisebericht in der deutschen Literatur. Ein Forschungsu¨berblick als Vorstudie zu einer Gattungsgeschichte. Internationales Archiv fu¨r Sozialgeschichte der Deutschen Literatur, 2, Niemeyer Verlag, Tu¨bingen.

LOESS LANDSCAPES HISTORY, NORTH SERBIA Child, T. 1889. Summer Holidays: Travelling Notes in Europe. Harper and Brothers, New York. Derbyshire, E. 2001. Geological hazards in loess terrain, with particular reference to the loess regions of China. Earth-Science Reviews, 54, 231– 260, http://dx.doi. org/10.1016/S0012-8252(01)00050-2 Dong, H., Song, Y., Chen, T. & Yu, L. 2014. Geoconservation and geotourism of Luochuan Loess National Geopark. Quaternary International, 334– 335, 40– 51, http://dx.doi.org/10.1016/j.quaint.2013. 10.023 Gray, M. 2004. Geodiversity – Valuing and Conserving Abiotic Nature. John Wiley and Sons, New York. Heller, F. & Liu, T. S. 1982. Magnetostratigraphical dating of loess deposits in China. Nature, 300, 431–433, http://dx.doi.org/10.1038/300431a0 Hose, T. A. 1995. Selling the story of Britain’s Stone. Environmental Interpretation, 10/2, 16–17. Hose, T. A. 2008a. Towards a history of geotourism: definitions, antecedents and the future. In: Burek, C. V. & Prosser, C. D. (eds) The History of Geoconservation. Geological Society, London, Special Publications, 300, 37–60. Hose, T. A. 2008b. Towards a history of landscape appreciation. In: Dowling, R. & Newsome, D. (eds) Discover the Earth Beneath our Feet. Inaugural Global Geotourism Conference 2008 Proceedings. Promaco Conventions Pty Ltd., Perth, 8–18. Hose, T. A. 2010. The significance of aesthetic landscape appreciation to modern geotourism provision. In: Newsome, D. & Dowling, R. K. (eds) Geotourism: The Tourism of Geology and Landscapes. Goodfellow, Oxford, 13–25. Jerrold, W. 1911. The Danube. Methuen and Co. Ltd., London. Jovanovic´, M., Gaudenyi, T., O’Hara-Dhand, K. & Smalley, I. 2014. Karl Caesar von Leonhard (1779– 1862), and the beginnings of loess research in the Rhine valley. Quaternary International, 334–335, 4–9, http://dx.doi.org/10.1016/j.quaint.2013.02.003 Kapper, S. 1851. Su¨dslavische Wanderungen Im Sommer 1850. F.L. Herbig, Leipzig. Klemun, M. 2007. Writing, ‘Inscription’ and Fact: Eighteenth Century Mineralogical Books Based on Travels in the Habsburg Regions, the Carpathian Mountains. Geological Society, London, Special Publications, 287, 49–61. Kohl, J. G. 1844. Austria. Vienna, Prague, Hungary, Bohemia, and Danube; Galicia, Styria, Moravia, Bukovina, and Military Frontier. Chapman and Hall, London. Kostic´, D. S. 2012. On the decks of the white fleet. Guidebooks for travellers prepared by the Danube Steamship Society (1895– 1939). In: Kostic´, D. (eds) Down the Danube from Bezdan to Belgrade. Balkanological Institute of the SASA, Belgrade, 7– 34. Kukla, G. J. 1977. Pleistocene land-sea correlations. Earth Science Reviews, 13, 307–374, http://dx.doi. org/10.1016/0012-8252(77)90125-8 von Leonhard, K. C. 1824. Charakteristik der Felsarten. Joseph Engelmann Verlag, Heidelberg, 3 vols. Loess in vol. 3. (Section 89 Loess Reprinted in Loess Letters 67, 3–5 April 2012.) Lukic´, T., Markovic´, S. B. et al. 2009. The loess ‘cave’ near the village of Surduk – an unusual pseudokarst

239

landform in the loess of Vojvodina, Serbia. Acta Carsologica, 38/2, 227 –235. Markovic´, S. B., Smalley, I. J., Hambach, U. & Antoine, P. 2009. Loess in the Danube region and surrounding loess provinces: the Marsigli memorial volume. Quaternary International, 198/1 –2, 5– 6, http://dx.doi.org/10.1016/j.quaint.2009.02.001 Markovic´, S. B., Hambach, U. et al. 2011. The last million years recorded at the Stari Slankamen loesspalaeosol sequence: revised chronostratigraphy and long-term environmental trends. Quaternary Science Reviews, 30, 1142– 1154, http://dx.doi.org/10.1016/ j.quascirev.2011.02.004 Marsigli, L. F. 1726. Danubius Pannonico-mysicus – observationibus geographicis, astronomicis, hydrographicis, historicis, physicis perlustrarus et in sex tomos digestus ab Aloysio Ferd. Com. Marsili Published 1726 by Apud P. Gosse, R. Chr Alberts, P. de Hondt, Apud Herm. Uytwerf & Franc¸. Changuion in Hagæ Comitum, Amstelodami. Written in Latin. Millet, F. D. 1893. The Danube – From Black Forest to the Black Sea. Harper and Brothers, New York. Paget, J. 1855. Hungary and Transilvania, with Remarks on Their Condition, Social, Political and Economical. John Murray, London. Paton, A. A. 1845. Servia, Youngest Member of the European Family; or, a Residence in Belgrade and Travels in the Highlands and Woodlands of the Interior, During the Years 1843 and 1844. Longman, Brown, Green, and Longmans, London. Pecsi, M. 1990. Loess is not just the accumulation of dust. Quaternary International, 7/8, 1– 21, http://dx.doi. org/10.1016/1040-6182(90)90034-2 Quin, M. J. 1835. A Steam Voyage Down the Danube. Richard Bentley, London. Smalley, I. J., Jefferson, I. F., Dijkstra, T. A. & Derbyshire, E. 2001. Some major events in the development of the scientific study of loess. Earth Science Reviews, 54, 5–18, http://dx.doi.org/10.1016/S00128252(01)00038-1 Solarska, A., Hose, T. A., Vasiljevic´, D. A., Mroczek, P., Jary, Z., Markovic´, S. B. & Widawski, K. 2013. Geodiversity of the loess regions in Poland: inventory, geoconservation issues, and geotourism potential. Quaternary International, 296, 68–81, http://dx.doi. org/10.1016/j.quaint.2012.08.2057 Stagl, J. 1995. A History of Curiosity. The Theory of Travel 1550– 1800. Studies in Anthropology & History 13. Harwood Academic Publishers, Chur. Vasiljevic´, Dj. A., Markovic´, S. B., Hose, T. A., Smalley, I., Basarin, B., Lazic´, L. & Jovic´, G. 2011a. The Introduction to geoconservation of loess-palaeosol sequences in the Vojvodina region: significant geoheritage of Serbia. Quaternary International, 240/1–2, 108–116, http://dx.doi.org/10.1016/j.quaint. 2010.07.008 Vasiljevic´, Dj. A., Markovic´, S. B. et al. 2011b. Loess towards (geo) tourism – proposed application on loess in Vojvodina region (north Serbia). Acta Geographica Slovenica, 51/3, 391–406, http://dx.doi.org/10.3986/ AGS51305 Vasiljevic´, Dj. A., Markovic´, S. B. et al. 2014. Loess– palaeosol sequences in China and Europe: Common values and geoconservation issues. Catena, 117, 108–118, http://dx.doi.org/10.1016/j.catena.2013.06.005

Index Page numbers in italics refer to Figures. Page numbers in bold refer to Tables. Adam, William, The Gem of the Peak (1838–1857) 147 Adrsˇpach Rock City 218, 219 geotourism 220, 225 visit by Goethe 220, 221 Agassiz, Louis (1807–1873), glaciology 85 Agassiz Rock, Edinburgh 32 Aix-les-Bains, Sierroz gorge 204– 205 Aladdin’s Crawl, Carlswark Cavern 164–165 Alps, The Grand Tour 9 Anthropocene 34–35 art depiction of geotourism, seventeenth century Holland 73–82 depiction of landscape Eugene von Gue´rard 59– 68 Scotland 29– 30, 33 seventeenth century Holland 71–73 environmental, Scotland 35 waterfalls 43 Australia, work of Eugene von Gue´rard 63– 68 Avebury, Lord (1834– 1913), The Scenery of England . . . (1902) 142, 144 Baedeker guides 14 Baker, Matt, Shinglehook 35 Bakewell, Robert (1767–1843), geology textbooks 142, 144 Bamforth Hole, Carlswark Cavern 160, 161, 162, 163 Bartholomew’s maps 134, 150– 152 beaches, seventeenth century Dutch landscapes 73–74, 76, 77 Beattie, William (1793– 1875), Danube loess landscape 234 Beaumont, Albanis (1755–1812), Travels from France to Italy through the Lepontine Alps (1800) 201 Belfast Field Naturalists’ Club 96 Beresteyn, Claes van (1637– 1684), dune landscape 77, 81 Bergsfjord Peninsula see Øksfjordjøkelen Bible, literal interpretation of 41, 60 Blackmoor Stannary 187 Bonnard, A.H., Carclaze tin mine 190 Boswell, James (1740– 1795), on waterfalls 45, 47, 51 Bowder Stone 16, 18 Bowen, Emanuel (c. 1694– 1767), county maps 132, 133 Bradshaw’s Descriptive Railway Hand-book (1863) 192 Bredalsfjellet 90, 91 British Speleological Association 160 Broc, Ade`le de (1788– 1813), drowning in Sierroz gorge 205, 206, 208 Brongniart, Alexandre (1770–1847), on upper-Rhoˆne 201, 203 Broumovske´ steˇny 218, 219, 226 Buch, Leopold von (1774–1853) glaciology of Norway 85 travel writing 11, 60

Bull Bay, Anglesey 101, 102 Burdett, Peter Perez (1734– 1793), Survey of Derbyshire (1767) 133, 138 –139, 140 Burke, Edmund (1729–1797), the beautiful and the sublime 13, 41, 47, 53 Byng, John, 5th Viscount Torrington (1743–1813), on High Force 51 Byron, Lord (1788–1824), search for the Sublime 53 Camden, William (1551–1623), on waterfalls 43, 51 caravanning 15 Carclaze, Old Tin Pit 187–195, 188, 194 china clay production 193 early development 189 eighteenth century 189 geology 189, 191–192 guidebooks 192 nineteenth century 190–193 post-industrial interest 193– 195 Carlswark Cavern 161 historical accounts 160, 162–163 modern damage 163–165 speleothems 158, 160, 162–165 Carreck Archive, Geologists’ Association 118– 122, 127, 128 Carreck, Marjorie, Geologists’ Association archivist 121–122 Carus, Carl Gustav (1789– 1869), landscape art and geology 60, 63 Cary, John (c. 1754–1835) geological maps 141 Improved Map of England and Wales . . . (1832) 138 New and Correct English Atlas (1787) 132– 133, 138, 139 New County Atlas (1809) 132, 133, 138 New Map of England and Wales with Part of Scotland (1794) 131– 132, 133, 138, 139, 141 Caspari, Josef, Lyngen glaciers 88– 89 cassiterite 189 caves 6 Somerset 180 Stoney Middleton Dale 157 –168 caving, Stoney Middleton Dale 159, 160 Chamonix Grand Tour 9 Mer de Glace 199 Charles II, King of England, departure from Scheveningen 73, 78 Charlestown United tin mine 187, 188 Charleswork see Carlswark Cavern Chester Society for Natural Science 96, 98– 101 Annual Reports 112, 114 excursions 101, 102, 103–106, 115 Grosvenor Museum 112, 113–114

242

INDEX

Chester Society for Natural Science (Continued) interaction with other societies 108, 109 membership 103, 104 honorary members 104, 110 role of women 101, 106, 107, 108– 109 prizes 113 transport 109, 111 Child, Theodore (1846– 1892), on Danube voyage 234, 235 China Clay History Society 193–194 china clay mining, St Austell 177, 187–188, 191, 193– 195 Claude Glass 14 Coke, Lady Mary (1727–1811), Grand Tour 9 Coleridge, Samuel Taylor (1772– 1834), on waterfalls 51, 53 Colonna, Fabio (1567–1640), on Glossopetrae 12 colour layering 134, 152 Colourists 33 Constable, John (1776– 1837) landscape painting 78 Lake District 16 contours 134 Conybeare, William (1787–1857), geology textbooks 142, 144 copper mining 178, 179 Cornwall Carclaze tin mine 187–195 eighteenth century tourism 188–190 ‘Geological’ ABC 125– 128 travels of William George Maton 171, 174, 175, 176, 177– 178, 182 Cyclists’ Touring Club 15 maps 151–152 Dale, Elizabeth, The Scenery and Geology of the Peak District (1894) 147 Daniell, William (1769– 1837), painting of Loch Coruisq 29, 30 Danube River geodiversity 229 loess-palaeosequences 230, 231 travel writing 231–238 steamer cruises 230–231 Darlington, local societies 96 De la Beche, Sir Henry (1796– 1855) 13 geology textbooks 144 Report on the geology of Cornwall . . . (1839) 191, 192 Dee river steamer 109, 111 Defoe, Daniel (c. 1660–1731) on Lake District 16, 18 on waterfalls 43, 45 Derby philosophical society 96 Derbyshire guidebooks 149– 152 maps 131, 132, 134, 135, 137, 140, 141, 145, 146–153 Stoney Middleton Dale caves 157–168 Devon, travels of William George Maton 174, 175, 176– 177, 178, 179–180, 181, 182 Dixon Hewitt, Henry, Geologists’ Association photography 119, 124– 125 Dorset, travels of William George Maton 172, 175, 176, 179, 180, 181, 182

Dudley and Midland Geological Society 96, 107, 109, 118 Dumas, Alexandre (1802–1870), Sierroz gorge 205 dunes, seventeenth century Dutch landscapes 73– 74, 75– 78, 79, 80, 81 ecotourism, definition 5, 7 Eden Project 188, 193 Edinburgh Geological Society 96 Eifel landscape, Eugene von Gue´rard 62–63 enclosure of land, eighteenth century, mapping 135 English Nature 1 Engravers’ Copyright Act (1766) 138 Enlightenment 25, 36 local societies 96 Espriella, Don Manuel Alvarez (Robert Southey), Letters from England 18 estate management, mapping 135 European Geoparks Network, Arouca Declaration (2011) 5 Eyam Dale 161, 162, 165 mining heritage 159 Eyam Dale Shaft, Carlswark Cavern 161, 162, 164, 165 fall-jøkull 85, 86 Falls of Clyde 27, 28, 31, 49 Falls of Foyers 27, 28, 31, 49 Farey, John (1766–1826), geology of Derbyshire 146, 148 Farington, Joseph (1747–1821) Carclaze tin mine (1814) 190, 191 Lodore Rocks – fall and cottage distance 48 field clubs 96 fieldwork 97 Fiennes, Celia (1662–1741) 9, 15, 16, 43, 171 Through England on a Side Saddle (1695) 189 on waterfalls 51 Fier gorge 199, 200, 205, 208, 209, 210, 211 Fingal’s Cave, Staffa 27, 28, 29, 30 Fireset Mine 159 Foley, Lady Emily (1805–1900), Woolhope Naturalists’ Field Club 98 Forbes, James D., Øksfjordjøkelen glacier 85–86, 87 Fossil Grove, Glasgow 32 Frusˇka Gora Mountain 231, 232, 233 Fugledalsbreen 91 Geikie, Sir Archibald (1835–1924) Øksfjordjøkelen glacier 85, 86 Scenery of Scotland (1901) 32 Ge´nissiat dam, ‘loss of upper-Rhoˆne’ 200, 203– 204 geo-interpretation 3, 4 geoconservation 3, 4, 97 Scotland 32, 33 geodiversity 4, 7 Danube River 229 Karkonosze Mountains 220– 221 geoheritage 7, 33 Carclaze Tin Mine 187 Karkonosze Mountains 223– 225 ‘loss of upper-Rhoˆne’ 203, 208, 211 Sierroz and Fier gorges 208, 211– 212 Sudetes 215–227 geohistory 97 geointerpretation 33, 97

INDEX Geological Society of London 13, 33 Geological Survey, maps 143, 149, 151 geological timescale 25–27, 31, 36, 41, 43 Geologists’ Association, and geotourism 13, 117–128 A ‘Geological’ ABC of North Cornwall 125– 128 Carclaze tin mine 193 Carreck Archive 118– 122, 127, 128 Circular 118 field meetings 117 –118, 119, 120, 121 transport 122–124 geological guides 118 Henry Dixon Hewitt 119, 124 –125 Proceedings 118, 120 Record 118 geology, nineteenth century county accounts 142 maps 139, 141–152 Scotland 32– 33 textbooks 142–144 geomorphosites 3, 4 French Alps 199 –212 Karkonosze Mountains 220–221, 223, 224, 225 loss of 7 Geopark Shetland 33, 34 Geoparks European Network, Arouca Declaration (2011) 5 Kanawinka 64, 67, 68 Karkonosze 226 Scotland 33 Vulkaneifel 62, 67 Geoparks Programme Feasibility Study 3 geosites 3, 4 Karkonosze Mountains 220–221, 223, 224, 225 loss of 7 primary 6 secondary 6 geotourism Chester Society for Natural Science 103– 106 Cornwall 187– 195, 190– 193 Danube loess-palaeosol sequences 230–238 European timeline 3 French Alps 199 –212 Geologists’ Association 13, 117–128, 193 Lyngen Alps, Norway, nineteenth century 87–91 Peak District 15– 16 recognition and definition 2 –6, 7, 96–97 role of local societies 95–115 Scotland 33 seventeenth century Holland 73–75, 80– 81 Stoney Middleton Dale 157– 168 Sudetes 215–227 Woolhope Naturalists’ Field Club 101–103 geotourists 1 casual 5, 6– 7, 97, 117 dedicated 5– 6, 59, 85, 97, 117 educational 6 recreational 6 Gilbert, Charles Sandoe (1760– 1831), Historical Survey of Cornwall (1817) 190 Gilpin, William (1724– 1804) Lake District landscapes 17 picturesque travel 41, 47, 49 waterfalls 47 Gimli’s Dream, Merlin Cavern 161, 166, 167

243

Gin Entrance, Carlswark Cavern 161, 162, 163, 164 glaciers, nineteenth century Norway 83– 87 recession 90– 91 glaciology, nineteenth century Norway 85 Glover, John (1767– 1849), Lake District landscapes 16 Goethe, Johann Wolfgang von (1749–1832) 59 geotourism in Sudetes 220, 221 Vesuvius 61 Goltzius, Hendrik (1558– 1617), whale stranding art 73– 74, 76 Go´ry Stołowe National Park 226 Grand Tour 7– 10, 11, 13, 19, 95 Gray, Thomas (1716– 1771) aesthetic travel 47 on Goredale Scar 51 Lake District landscapes 16 Green, William (1760–1823), The Tourist’s New Guide 17, 18 Gre´sy-sur-Aix, drowning of Ade`le de Broc 205, 206, 208 Grosvenor Museum, Chester 112, 113– 114 Gue´rard, Johann Joseph Eugene von (1811– 1901) 59– 68 Australia 63–68 Australian Landscapes 65–68 Crater of Mount Gambier, South Australia 67 The Basin Banks, near Camperdown 66 View of Koroit or Tower Hill 65 Vesuvius 60–61 Vulkaneifel 61– 63 guidebooks 14–15 beautiful, sublime and picturesque scenery 47, 49 Carclaze, Old Tin Pit 190– 192 Cornwall 190 –192 cost 152 Derbyshire 15, 149– 152 Lake District 16, 17–18 Scotland 27, 31 ‘The British Tourists’ (1798– 1800) 11 Haarlem, seventeenth century landscape tourism 75, 79 hachures 133– 134 Hakewill, James (1778– 1843), Cascade of Terni 52 Hall, Elias (1764–1853), geology of Derbyshire 147 Hamilton, Sir William (1731– 1803), An Account of a Journey to Mount Etna 8, 60 Happy Union tin mine 187 Hardraw Force, payment to view 54 hardship, pleasures of, Sublime movement 27, 41, 51, 53, 60 Harrison, William J. (1845–1909), Geology of the Counties of England and of North and South Wales (1882) 142, 144 Hastings, George Lyngen glaciers 88–89 Øksfjordjøkelen glacier 85 Hatchett, Charles (1765–1847), travels with William George Maton 171, 172, 174– 179, 181, 183, 184 Hedinger, J.M., A Short Description of Castleton.. (1839) 137, 147 Hereford Free Public Library and Museum 111, 113, 114 Herefordshire Natural History, Literary, Philosophical and Antiquarian Society 97

244

INDEX

High Force 44, 49, 51 hiking 15, 95, 223 hill shading 135 History of Geology Group (HOGG) 1 Hochstetter, Ferdinand von (1829–1884) and Eugene von Gue´rard 66– 67 volcanic forms of western Victoria 65 Hogrenning, Elias Lyngen glaciers 88–89 Øksfjordjøkelen glacier 85 Holland, seventeenth century landscape art geomorphological accuracy 71– 73 as new genre 75– 79 population 71, 72 Humboldt, Alexander von (1769–1859) 59 Cascade de Tequendama 45, 46 on waterfalls 45 hundreds 132 Hutton, James (1726–1797) Siccar Point unconformity 25 Theory of the Earth 41 Huygens, Sir Constantijn (1596–1687), sea-way 74, 79 industrial development, expansion of tourism 31 Ingleton, railways and tourism 54 Jamaica, waterfalls 54– 55 James, John Thomas (1786– 1828), Sweden 11 Jars, Gabriel (1732–1769), Carclaze tin mine 189, 190 Jelenia Go´ra basin 221, 222, 223, 226 geological cross-section 224, 226 Jerrold, Walter Copeland (1865–1929), on Danube voyage 232 –233 Johnson, Dr Samuel (1709– 1784), on waterfalls 45, 47, 51 Johnston, Mary Sophia (1875– 1955), Geologists’ Association archivist 121, 122 Kamien´czyk gorge 222, 223 kaminitza, Fier gorge 208 Kanawinka Global Geopark 64, 67, 68 kaolin see china clay Kapper, Isaac Salomon (Siegfried) (1821– 1879), on Danube voyage 232, 233 Karkonosze Mountains 216, 222 geoheritage 223–225 geosites 220 –221, 223 catalogue 224, 225 geotourism 221–225 nature reserves 224 –225 Karkonosze National Park 226 karst, Fier gorge 208 killas 182, 183, 191 Kingsley, Canon Charles (1815–1875), Chester Society of Natural Science 99, 113 Kitchen, Thomas (1718– 1784), maps 133 Knipe, James Alexander (c. 1803–1882), Geological Map of the British Isles and Part of France (1843) 142, 143, 151 Kohl, Johann Georg (1808– 78), on Danube voyage 232, 233

Labe river 221, 222 Lake District guidebooks 17–18 landscape painting 16 landscape tourism 16–18 landscape tourism 10– 11, 15–18, 95 Scotland 27– 31 seventeenth century Holland 73–82 Landseer, Edwin (1769–1862) engraving of High Force or Fall of Tees 44 Scotland 31 Lassels, Richard (1603– 1668), The Grand Tour 8, 11 Le Blond, Mrs Aubrey see Main, Elizabeth Leland, John (1503– 1552) 43, 135, 171 Leonhard, Karl Caesar von (1779–1862), loess 235 Lessing, Carl Friedrich (1808–1880), Eifel expedition 63 Lhuyd, Edward (1660– 1709) 12 Lightspout Valley, Church Stretton 101 Ligorio, Pirro (c. 1510–1583) 11 Little Ice Age, Norway 83–85, 86– 87 Liverpool Geological Society 96, 108, 109 local societies 96–115 Loch Coruisq 28, 29 Lodore Falls 47, 48, 49 payment to view 54 loess 230 palaeosol sequences, Vojvodina, north Serbia 230– 238, 231 Lowe, Joseph (1865–1934), photography of Lake District 16 Lower Resurgence Entrance, Carlswark Cavern 160, 161, 163 Lower Sump, Carlswark Cavern 160, 162, 163 Lubbock, John, 1st Baron Avebury (1834– 1913), The Scenery of England . . . (1902) 142, 144 Lucey, ‘loss of the upper-Rhoˆne’ 201, 202, 204 Lyell, Charles (1797– 1875) Elements of Geology (1838) 144 Principles of Geology (1830–1833) 32, 142, 144 Lyngen Peninsula, Norway 83, 84 nineteenth century geotourists 87– 91 MacCulloch, John (1773– 1835), Fingal’s Cave 28, 29, 30 Mackay, Charles (1814–1889), The Scenery and Poetry of the English Lakes (1846) 18 McKenny Hughes, Professor Thomas (1832– 1917), Chester Society for Natural Science 99, 100 Mackintosh, D., Scenery of England and Wales (1869) 142, 144 Macpherson, James (1736–1796), Ossianic poetry 27 Main, Elizabeth (1860–1934), Lyngen Peninsula 89– 91 maps 131– 153 eighteenth century county maps 132–133 developments 135, 138– 139 plagiarism 138 strip maps 132, 135, 136 nineteenth century cost 151, 152 geological maps 139, 141 –152

INDEX Peak District 131, 132, 135, 137, 146– 153 topographic depiction and accuracy 132, 133– 135, 138– 139 colour layering 134, 152 contours 134 hachures 133–134 hill shading 135 RSA prize 138 Marliano, Bartolomeo (1488– 1566) 11 Marsigli, Luigi Ferdinando (1658–1730) 233, 234– 235 Danubius Pannonico-Mysicus (1726) 12, 235, 236 Martel, Edouard-Alfred (1859– 1938), ‘loss of upperRhoˆne’ 203 Martin, William (1767–1810), Petrificata Derbiensia (1809) 147 Maton, William George (1774–1835) 172– 174 travels in SW England 174, 175, 176 journals (1797) 171, 172, 173, 174–184 Mawe, John (1766– 1829), The Mineralogy of Derbyshire (1802) 146 Mello, Rev. J.M., Hand-Book to the Geology of Derbyshire (1876) 147 Merlin Cavern 161 Gimli’s Dream 161, 166, 167 historical accounts 165–166 modern damage 166, 167, 168 speleothems 158, 165–166 Streamway 161, 166 Michell, Philip, Carclaze tin mine 192, 193 Miller, Hugh (1802–1856), popular geology 32 Millet, Francis Davis (1848–1912), on Danube loess 235, 237–238, 237 mineralogy, travels of William George Maton 181, 182, 183 mining, Sudetes mountains 217, 221 mining heritage Carclaze Tin Mine 187–195 Stoney Middleton Dale caves 159, 166 Montagu, Elizabeth (1718–1800), on Scottish landscape 27 Montrottier Castle 205, 209 Moore, Henry, Picturesque Excursions in the High Peak . . . (1819) 147 Morden, Robert (c. 1650– 1703), maps 133 Moritz, Karl Philip (1756– 1793), Peak District 15 Mosenberg crater lake, work of Eugene von Gue´rard 62, 63 Moule, Thomas (1784–1851), map of Derbyshire (1837) 135, 137 Mount Gambier, work of Eugene von Gue´rard 67 Mount S´niez˙ka 221–222, 223 Mount Szczeliniec Wielki 217, 218, 219 geotourism 220, 225, 226 visit by Goethe 220, 221 Mount Witosza 222, 223 Murchison, Sir Roderick Impey (1792–1871) 13 Nasmyth, Alexander (1758– 1840), Lower Falls of Fyers 28 National Geographic, approach to geotourism 4–5 National Parks Sudetes Mountains 226 UK 15

245

National Parks and Access to the Countryside Act (1949) 15, 33 Natural England see English Nature naturalists’ societies 95–115 navigation, early maps 132 Neo-Romanticism 3, 71 Neptunism v. Vulcanism 60 Netherlands see Holland Newcastle, local society 96 Norway, nineteenth century tourism 83– 92 Norwich school of landscape painters 78 Novi Sad 231, 233 Nugent, Thomas (c. 1700– 1772), The Grand Tour 8 Nut Scrin, Carlswark Cavern 160, 161 Odin’s Mine, Castleton 137 Ogilby, John (1600– 1676) Brittania (1675) 132, 135 strip maps 132, 135, 136 Øksfjordjøkelen 83, 84 glaciers 85– 87 Ordnance Survey 1 inch maps 132, 139, 147, 149, 150, 151 cost 151, 152 establishment of 135, 138 Oyster Chamber, Carlswark Cavern 161, 162, 163, 164 Paget, John (1808–1892), on Danube voyage 233–234 Pannonian Plain 232, 234 Parsons, Alfred William (1847– 1920), on Danube loess 235, 237 Paton, Andrew Archibald (1811– 1874), on Danube voyage 232, 233 Peak District landscape tourism 15–16, 132 maps 131, 132, 137, 140, 141, 146–153 Stoney Middleton Dale caves 157 –168 Pennant, Thomas (1726– 1798), on waterfalls 49 Petrovaradin fortress 231, 233, 235 Phillips, William (1773– 1828), geology textbooks 142, 144 philosophical societies 96 photography Carreck Archive 118–122, 124–125 Lake District landscapes 16 Photographic Supplement to Stanford’s Geological Atlas (1913) 143 Scotland landscape 31 Picturesque movement 14, 17, 41, 132 Pielgrzymy tor 223 plagiarism, map making 138 Playfair, John (1748–1819), abyss of time 25 Plot, Robert (1640–1696) 12 Pococke, Dr Richard (1704–1765), travel writing 171 Polgooth tin mine 177, 178– 179, 187, 188 Police Basin 218 geology 217– 218 tourism 218 Quin, Michael J. (1796– 1843), on Danube voyage 232

246

INDEX

Rackett, Thomas (1755–1840), travels with William George Maton 171 –172, 174–184 railways 14 and tourism 54, 95 Cornwall 189–190, 192– 193 Fier gorge 205 Geologists’ Association 123, 193 naturalists’ society excursions 105, 109, 111 Scotland 31 Raleigh Club of Travellers 13 Ramsay, Sir Andrew Crombie (1814– 1891), Physical Geology and Geography of Great Britain (1862) 32, 142, 144 Rashleigh, Philip (1729– 1811), Carclaze tin mine 190 Reader, Thomas William (1860– 1923) albums 120, 121, 123, 124, 126 Geologists’ Association 121 recreation, geological 5 Renevier, Eugene (1831–1906), on upper-Rhoˆne gorge 203 Reynolds, James (1817– 1876), Geological Atlas of Great Britain (1860) 142– 144, 144, 145 Rhoˆne , upper, loss of 200–204 Riesengebirgsverein (RGV), Karkonosze Mountains 223 Rift Sump, Carlswark Cavern 165 roads, mapping 135 Rock Garden, Merlin Cavern 161, 165 ‘romantic gaze’ 43 Romantic movement 8, 13, 19, 95 Peak District 132 Scottish landscape 27–30, 32, 33 and waterfalls 41–55 Romanticism 41 see also neo-Romanticism Rome, The Grand Tour 9 Rowlandson, Thomas (1756–1827) A high waterfall 42 Roy, Major-General William (1726–1790), military maps 135 Royal Geographical Society 13 Royal Geological Society of Cornwall 96 Carclaze tin mine 190 Royal Society 9 Royal Society of Arts, accurate surveying prize 138 Royal Society of Edinburgh, Boulder Committee 32 Ruisdael, Jacob van (1628–1682), geomorphological accuracy 73, 75– 76, 80 ‘ruralism’ 71 Ruskin, John (1819–1900) 31, 54 Saal, Georg (1817–1870), Vulkaneifel 62– 63 St Austell Carclaze tin mine 187–195 china clay mining 187 –188, 193, 195 Salisbury Crags, Edinburgh 32 Saussure, Horace-Be´ne´dict de (1740– 1799), Voyages dans les Alpes (1779) 201, 204 Saxton, Christopher (c. 1542– c. 1610), Atlas of the Counties of England and Wales (1579) 132, 133 Scandinavia, travellers 11 Scheveningen sea-way 74, 79 seventeenth century Dutch landscapes 74– 75, 78

Scotland geology and physical landscape 25–36 tourism of awe 27–31 Scott, Sir Walter (1771–1832), Romantic movement and landscape 30 Scottish Parliament, geological wall 33 sea-way, The Hague to Scheveningen, Sir Constantijn Huygens 74, 79 Sedgwick, Adam (1785– 1875), Carclaze tin mine 190 Serbia, loess-palaeosol sequences 231, 232–238 ‘shambling’ 189 Short, Dr Thomas, Natural, Experimental and Medicinal History of Mineral Waters (1734) 160, 162 Siccar Point unconformity 25 Sierroz gorge 199, 200, 204–205, 206, 207, 208 Simpson, Samuel, map of Derbyshire (1746) 135, 137 Sites of Special Scientific Interest 33 Stoney Middleton Dale caves 157– 168 Skalnı´ chra´m 220 Skeat, Ethel (1869–1939), Chester Society for Natural Science 99 Slingsby, William Cecil, Lyngen glaciers 88–89 Smith, William (1769–1839) A Delineation of the Strata of England and Wales with Part of Scotland (1815) 132, 139, 141, 151 County Geological Atlas (1819) 132 S´niez˙ne Kotły cirque 222, 223 Society of Antiquaries 12 Somerset, travels of William George Maton 174, 175, 180– 181, 182 sough drainage, Stoney Middleton Dale 159– 160 Southey, Robert (1774– 1843) see Espriella, Don Manuel Alvarez spas, Sudetes mountains 217 Speed, John (1552– 1629) Darbieshire described (1610) 134, 135 The Theatre of the Empire of Great Britaine (1610) 132, 133 on waterfalls 51 speleothems, Stoney Middleton Dale SSSI Carlswark Cavern 158, 160, 162–165 Merlin Cavern 158, 165– 166 Staffa 27, 28, 29, 30 stalactites 162, 163 stalagmites 162, 163, 167 Stanford, Edward (1827–1904) Geological Atlas of Great Britain (1904) 143, 144, 145 Photographic Supplement (1913) 143 Stari Slankamen 231, 237, 238 steamships, and tourism River Danube 230–231 River Dee 109, 111 Scotland 31 Sierroz gorge 207 Stebbing, William Pinckard Delane (1873– 1961) Geologists’ Association archivist 120 –121 Steno, Nicholas (1638–1686) 12, 19 Stevin, Simon (1548–1620), sailing car 73, 74, 77 Stołowe Mountains 216, 218 geology 217–218 tourism 220, 226

INDEX Stoney Middleton Dale caves, Derbyshire 157– 168, 158, 161 Carlswark Cavern 160, 162–165 educational interest 159 flora and fauna 159 geology 159 industrial heritage 159– 160, 166 Merlin Cavern 165– 168 recreational caving 159, 160 Strupbreen glacier 88, 89 Strupvatnet lake 88– 89 Stub Scrin, Merlin Cavern 166 Sublime movement 13, 27, 41, 47, 51 pleasures of hardship 27, 41, 51, 53 volcanoes 60 Sudetes mountains 216–217, 216 geotourism and geoheritage 215– 227 pre-World War II 225 twenty-first century 225–226 mining 217 political history 217, 220 thermal spas 217 surveying eighteenth century mapping 133–135, 138– 139 triangulation 138, 139 Swallow Falls, payment to view 54 Switzerland, The Grand Tour 9 Taylor, John (1580– 1653) 15 Teplice Rock City geotourism 220, 225 visit by Goethe 220, 221 Terni, Cascade of 52, 53 textbooks, geological, nineteenth century 142– 144 Thomas Cook’s tours 31 Thomson, Thomas (1773–1852), on Sweden 11 Thornton Force 54 tin mining Carclaze 187–195 travels of William George Maton 177, 178– 179 Titel loess plateau 231, 235, 236 topography, depiction in early maps 132, 133– 135 colour layering 134 contours 134 hachures 133–134 hill shading 135 Torrington, Viscount see Byng, John tors, Sudetes mountains 219, 223, 225 tourism 10–11 eighteenth century 19 expansion due to industrial development 31 French Alps 199 –212 improved access 53–55 nineteenth century 13, 14–15, 19 Scotland 27– 31 Sudetes mountains 217, 218, 220– 227 twentieth century 15 tourism of awe 27– 30 ‘tourist gaze’ 10, 11 Tower Hill maar volcano 63, 64, 65, 68 trains see railways

247

triangulation 138, 139 tub boats, Carclaze tin mine 189, 190, 194 Turner, Joseph William Mallard (1775–1851) Cascade of Terni 52 High Force or Fall of Tees 44 Lake District landscapes 16 Loch Coruisq, Skye 30 Upper Falls of the Reichenbach: Rainbow 50 Tyndall, John, glaciology of Norway 85 uniformitarianism 19, 25, 54 Velde, Esaias van der (1587– 1630), dune landscape 75, 79 Vesuvius Eugene von Gue´rard 60–61 The Grand Tour travel writing 8 Victoria Falls 45 Vojvodina , north Serbia, loess-palaeosol sequences 230–238, 231 volcanism, Australia 63–65, 66, 67 Vulcanism v. Neptunism 60 Vulkaneifel European Geopark 62, 67 Wanderer im Riesengebirge 223, 224 Ward Lock Guides 150, 152 Ward, Rev. R., geology of Derbyshire 150 waterfalls 6, 27, 28 as impediment to navigation 43 improved access to 53–55 Jamaica 54– 55 pleasures of hardship 51, 53 and romantic love 55 Romantic movement 41–55 as source of power 43 twentieth century 54–55 Watergrove Mine 159 Watson, White (1760– 1835), on geology of Derbyshire 146 West, Thomas (1720– 1779) A Guide to the Lakes (1778) 17, 49 Peak District appendix 146 Picturesque movement 14 waterfalls 49, 51, 53 whale strandings, seventeenth century Dutch art 73– 74, 76 Wheal Eliza tin mine 187, 188 Wheal Martyn 188, 193 Whitaker, H.B. (ed.), Geological Atlas of Great Britain (1904) 143, 144, 145 Whitehurst, John (1713– 1788), An Inquiry into the Original State and Formation of the Earth (1778) 146 Wilkinson, Rev. Joseph (1764–1831), Select Views in Cumberland, Westmoreland and Lancashire 17 Windham, William (1717– 1761), Mer de Glace, Chamonix 199 women, role in early geological societies 101, 102, 106–109 Wonder Cavern see Carlswark Cavern Wonder Scrin, Carlswark Cavern 160, 161, 162 Woodstock Field Naturalists’ Club 96 Woodward, Horace B. (1848– 1914), The Geology of England and Wales (1876) 142, 144

248 Woolhope Naturalists’ Field Club 96, 97– 98 excursions 98, 101–102, 103, 115 transport 109, 111 geotourism 101–103 interaction with other societies 107, 109 membership 102–103 honorary members 110, 111 role of women 102, 106–108 museum 113, 114 prizes 112– 113 Transactions 111, 112, 114 Wordsworth, William (1770–1850) crossing the Alps 9

INDEX on Lake District 13– 14, 17–18, 49 on waterfalls 49 Worm, Ole (1588–1655) 11–12 Wyndham, Henry Penruddocke (1736–1819), A Gentleman’s Tour . . . (1775) 49 Yorkshire Geological Society 96 Young, Arthur (1741– 1820) Lake District landscapes 17 on waterfalls 49, 51, 53 Zandvoort, seventeenth century Dutch landscapes 73