Essays on the History of Transportation and technology [First edition.] 9781940425030, 1940425034

Emory Kemp is the founder and director of the Institute for the History of Technology and Industrial Archaeology at West

295 30 5MB

English Pages 426 Year 2014

Report DMCA / Copyright

DOWNLOAD PDF FILE

Recommend Papers

Essays on the History of Transportation and technology [First edition.]
 9781940425030, 1940425034

  • 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

E MO RY L . K E M P

Essays on the

History of Transportation

and Technology

Foreword by LANCE E. METZ

Introduction by ROBERT J. KAPSCH

E SSAYS ON T H E H IS TORY OF T R A NSPORTAT ION A N D T ECH NOLOGY

Essays on the History of

E MO RY L . K E M P Transportation and Technology Foreword by LANCE E. METZ Introduction by ROBERT J. KAPSCH

West Virginia University Press Morgantown 2014

West Virginia University Press, Morgantown 26506 © 2014 West Virginia University Press All rights reserved First edition published 2014 by West Virginia University Press Printed in the United States of America 21 20 19 18 17 16 15 14

123456789

ISBN: Cloth: 978-1-938228-81-0 EPUB: 978-1-940425-04-7 PDF: 978-1-940425-03-0 Library of Congress Cataloging-in-Publication Data: Essays on the history of transportation and technology / Emory L. Kemp [and others]. -- First edition. pages cm Includes bibliographical references. ISBN 978-1-938228-81-0 (cloth : alk. paper) -- ISBN 1-938228-81-2 (cloth : alk. paper) -- ISBN 978-1-940425-03-0 (pdf) -- ISBN 1-940425-03-4 (pdf) -- ISBN 978-1-940425-04-7 (epub) -- ISBN 1-940425-04-2 (epub) 1. Technology--History. 2. Engineering--History. 3. Transportation--History. I. Kemp, Emory L., author. T15.E78 2014 627'.1309--dc23 2013042732 Previous versions of many of these essays appeared in Canal History and Technology Proceedings, Volumes X, XXI, XXII, XXIII, XXIV, XXV, XXVI, XXVII, XXIX, and XXX. Book and jacket design by han Safel Front jacket image: Roadway designs, by Emory L. Kemp. Back Jacket image: he wood deck of the Wheeling Suspension Bridge before repairs were made, 1956. (L. L. Jemison, West Virginia State Road Commission)

To Lance Earl Metz

Foreword

ix ¶

Preface

xi

¶ Introduction

1

he 1859 Wheeling Custom House: A Harbinger of Iron and Steel Skeletal Framing

7

Charles Ellet Jr. (1810–1862): Portrait of an Engineer

23

A horoughfare hrough the Howling Wilderness: he Weston & Gauley Bridge Turnpike

45

he Pulaski Skyway—Railway Economic heory Applied to Superhighway Design

91

James Finley and the Origins of the Modern Suspension Bridge

142

CONTENTS French Movable Dams in America

168

Building the Tennessee-Tombigbee Waterway

187

Benjamin Franklin homas and the Introduction of the French Needle Dam into the United States

253

John Jervis and the Hydraulic Design of the Old Croton Aqueduct

287

he Muskingum Navigation

313

French Movable Dams on the Great Kanawha River

353

he Little Kanawha Navigation

385

Acknowledgments

410 ¶

About the Authors

411

Foreword

I

t has been my distinct honor and pleasure to serve as the editor of the National Canal Museum’s Canal History and Technology Proceedings. hese volumes contain the texts and illustrations of the annual Canal History and Technology Symposiums, which are hosted and co-sponsored by Lafayette College, Easton, Pennsylvania. Over the past thirty years, I have worked with many noted contributors, but none has displayed the depth of knowledge and broad-based experience that have characterized the papers submitted by Professor Emory Kemp of West Virginia University. Professor Kemp’s papers have displayed his vast expertise in both civil and mechanical engineering. Equally important, he has managed the diicult task of setting each of his essays within the proper historical contexts. He has exhibited a particular mastery of the evolution of waterway engineering with papers on the “Little Kanawha Navigation System” (co-authored by Larry Sypolt) and “he Evolution of French Moveable Dams on the Great Kanawha.” He has thoroughly examined the creation of New York City’s modern water system in “John Jervis and the Hydraulic Design of the Old Croton Aqueduct” (coauthored by Edward Winant). Two of his papers are thorough examinations of the inluence and adoption of innovative European waterways technology in the United States: “Benjamin Franklin homas and the Introduction of the French Needle Dam in America” and “French Movable Dams on the Great Kanawha River.” he creation of what will probably become the last man-made major navigable waterway in the United States is explained in “Building the Tennessee Tombigbee Waterway.” Professor Kemp has gained national recognition for his work and designs for the rehabilitation, and in some cases, reconstruction of historic bridges. his unique expertise is evident in his papers on “James Finley and the Origins of the ix

Foreword

Modern Suspension Bridge” and later “he Pulaski Skyway” (co-authored by Dara Callender). Finally, the history of roadways is the subject of “A horoughfare hrough the Howling Wilderness: he Weston & Gauley Bridge Turnpike” (co-authored by his wife, Professor Janet Kemp). By reprinting these papers, West Virginia University Press will make them available to a wider audience. Professor Emory Kemp’s unique blend of historical and technical skills should become widely known among historians and engineers, and I hope this book will accomplish that goal. Lance E. Metz

x

Preface

M

any years ago, in fact ive decades, while a post-graduate student in engineering at Imperial College of Science, Technology and Medicine, London University, I became committed to studying Britain’s role in the beginning and later developments in the British Industrial Revolution. During this time I irst went to Ironbridge where I was greeted with a sign proclaiming “1709 where it all began.” What began was the smelting of iron with coke instead of charcoal by ironmaster Abraham Darby. His grandson later became most notable for casting the iron components for the great iron bridge completed in 1779 over the Severn River at Ironbridge and still in existence. My interest in the history of technology was piqued by noon-time lectures by Professor Alec Skempton at Imperial College. Ater coming to West Virginia University as a professor and chair of civil engineering, my interest in the history of technology continued. With the encouragement of WVU President Harlow, I relinquished my chair in civil engineering, becoming a professor of history and starting a program in the history of science. What began as an isolated interest was greatly enhanced and led to the founding of the Institute for the History of Technology and Industrial Archeology. he new Institute was formed on the essential elements of the history of technology and in addition was intended to foster the new discipline of industrial archeology. he principal activities included the following: Documentation according to U.S. national standards of historic sites in partnership with the U.S. Park Service. Selected documents were submitted to the Library of Congress. Restoration of selected bridges and buildings, listed on the National Register of Historic Places. xi

P r e fa c e

Publications comprising papers presented at the annual symposium on canals and technology at the National Canal Museum, Easton, Pennsylvania, comprise the essays in this collection. Several of the restoration projects received professional awards. Many graduate students were involved in the activities of the Institute, two of whom received Ph.D.s. he students participated in ield documentation, measured drawings, archival photographs, and contextual site histories. For more than a decade the Institute enjoyed more than $13 million in external support. he essays in this book represent many years of research in the history of technology while I served as Director of the Institute. Emory L. Kemp

xii

Introduction Robert J. Kapsch

F

or over ity years, Dr. Emory Kemp has been researching and writing on the history of engineering, especially American civil engineering. his book represents a selection of those writings. Dr. Kemp initially developed his keen interest in the history of engineering in London, while a student at Imperial College of Science, Technology and Medicine, London University. It was there that he listened to the lectures on the history of engineering given by Professor Alec W. Skempton (the major force in the development of the British Institution of Civil Engineers’ history and heritage programs), became aware of the Newcomen Society (the oldest English-language society dedicated to the study of the history of technology), and observed the early development of industrial archeology in Britain (archeology meaning, in its broadest sense, knowledge). Upon his return to the United States, Dr. Kemp became a professor of civil engineering (later, also, of history) at West Virginia University, at Morgantown and, later, established there the Institute for the History of Technology and Industrial Archeology. He was a lifelong friend of those who developed the ield of American history of technology and engineering, including Neil Fitzsimons (the moving force in the development of American Society of Civil Engineers’ history and heritage programs), Robert Vogel (the Smithsonian curator of civil and mechanical engineering of the National Museum of American History and the driving force behind the establishment of the American Society for Industrial Archeology), Lance Metz (the intellectual leader behind the development of the National Canal Museum), Charles Peterson (the National Park Service architect referred to as “the father of historic preservation”) and numerous others. And, of course, a good friend of mine. 1

Introduction

Emory’s research interests represented in these essays can be grouped into four categories: (1) nineteenth-century improvements in building materials, particularly iron, steel, and concrete; (2) roads and bridges, especially the history and evolution of the suspension bridge; (3) canals and navigable rivers, especially those of the Ohio River and its tributaries; and (4) water supply systems. Dr. Kemp’s essays are of particular interest because they usually contain two aspects of history: of the construction of the structure and, also, of the development of the analytical science that made the construction of the structure possible. In this volume, Dr. Kemp’s interest in improvements in building materials can be especially seen in his essay “he 1859 Wheeling Custom House: A Harbinger of Iron and Steel Skeletal Framing” (chapter 1). On the eve of the Civil War, Wheeling was an important commercial center and a competitor to its neighbor to the north, Pittsburgh. It was at Wheeling that the National Road crossed the Ohio on Charles Ellet’s recently completed suspension bridge (1849). It was also here that the U.S. Government decided to build a state-of-the-art customs house, designed by Department of Treasury architect Ammi B. Young (1798–1874) and engineer Captain Alexander Hamilton Bowman (1803–1865). Dr. Kemp’s essay is a case study of how the advancements in iron and steel were gradually integrated into American building practice. At the Wheeling Custom House, these new materials were used primarily for ireprooing and wide-span loor support but not for wall support (the building’s walls were of traditional load-bearing masonry). Eventually, in Chicago, these new materials would be used in walls to build the country’s irst skyscrapers. As explained by Dr. Kemp, initial attempts to use cast-iron beams to produce larger, ire-resistant interior spans were unsatisfactory because of the brittle nature of cast iron—the underside of the beam would tend to crack under load. So rolled wrought-iron beams were used in the Wheeling Custom House. At this early date, “I”-shaped beams were just beginning to be rolled by Peter Cooper and others and nine-inch deep “I” sections were used in the Wheeling Custom House loor system. A second area of Dr. Kemp’s interests is in roads and bridges. His essay “A horoughfare hrough the Howling Wilderness: he Weston & Gauley Bridge Turnpike” (chapter 3), written with his wife, Janet K. Kemp, is a case study of early nineteenth-century highway technology built by a private organization. 2

Introduction

he Weston & Gauley Bridge Turnpike was a 110-mile-long north-south West Virginia road connecting the Staunton-Parkersburg Turnpike at Weston in the north with the James River and Kanawha Turnpike at Gauley Bridge in the south. For those unfamiliar with West Virginia geography, a north-south line drawn down the very middle of a West Virginia map would approximate the route of this turnpike. Never inancially successful, the turnpike was largely abandoned ater the Civil War. Later, parts of its route were developed as U.S. Highway 19. Roads require bridges and Dr. Kemp has maintained a long-standing interest in the history and development of bridges, especially suspension bridges. As a bridge type, the suspension bridge was very important as longer spans could be built without constructing expensive and diicult-to-build intermediate piers in the waters to be crossed. “James Finley and the Origins of the Modern Suspension Bridge” (chapter 5), details the beginnings of this bridge type in the United States. Instead of wire, Finley (1756–1828) suspended the bridge deck from chain links between towers. As explained by Dr. Kemp, Finley, in this era before the development of the science of strength of materials, had to devise his own rules of thumb and calculations to ensure the safety and stability of the bridge. His best-known work is probably Chain Bridge (1807) at Little Falls, across the Potomac River above Georgetown, Washington, D.C. Also included in this volume is “Charles Ellet Jr. (1810–1862): Portrait of an Engineer” (chapter 2). Whereas Finley was an important bridge cratsman, Ellet was an important bridge engineer. At an early age, Ellet let his position as assistant engineer at the Chesapeake and Ohio Canal at the Monocacy Aqueduct to study in Paris and was admitted to the École Ponts et Chaussées in Paris, an extraordinary achievement since Ellet had no formal education. Upon his return to the United States, Ellet became an important advocate of suspension bridges, building the Fairmount Bridge across the Schuylkill River at Philadelphia (1841–42), the 1,010-foot span suspension bridge at Wheeling (1847– 49), and others. Ellet’s life was cut short by the Civil War. Dr. Kemp’s interest in roads and bridges comes together in “he Pulaski Skyway–Railway Economic heory Applied to Superhighway Design,” written with Dara Callender (chapter 4). he Pulaski Skyway was both road and bridge. Known as the “Superhighway,” it was an elevated viaduct four miles long on 108 3

Introduction

steel spans connecting the New Jersey entrance of the Hudson Tunnel at Jersey City with Newark, completed in 1932. Dr. Kemp and Ms. Callender explain the analytical basis for this project: that the expense for this structure was justiied using an adaptation of earlier developed railroad economic analysis—today called beneit-cost analysis. he third area of Dr. Kemp’s research interests is canals and navigable rivers, especially those in the Monongahela and Ohio River valleys. He has extensively studied French-designed movable dams in the United States, as summarized in “French Movable Dams in America” (chapter 6). “Movable dams” is somewhat of a misnomer as the dams do not move, rather components within the dams are moved to control the low of water in the river to facilitate river transport and lood control. One of these movable dams is the French needle dam discussed in “Benjamin Franklin homas and the Introduction of the French Needle Dam into the United States” (chapter 8). Needle dams consisted of a frame built across the river with long and narrow pieces of wood, called needles, laid against the frame in an almost vertical position. Never intended to be watertight, the dam increased the water level of the river when the needles were in place. When the river was in lood, the needles were removed to allow lood waters to pass. Chapter 11, “French Movable Dams on the Great Kanawha River,” is a case study of how French technology was utilized on the Great Kanawha, a West Virginia tributary of the Ohio lowing northwest through Charleston to its mouth at Point Pleasant. he Great Kanawha became the irst canalized river in the United States (1898) through the use of these movable dams, particularly through the construction of ten Chanoine wicket dams. Chanoine wicket dams were developed by Jacques Maurice Chanoine (1805–1876). his type of dam controlled river low by shutters (also called wickets) that were positioned to be self-acting (i.e., open and shut through the use of water pressure). It was U.S. Army Colonel William F. Merrill (1837–1891) who irst promoted the use of this type of device in the United States. Over his career, Dr. Kemp has also maintained an active interest in the Ohio River and its navigable tributaries. he Ohio River is one of the world’s great water-navigation systems, extending 981 miles from Pittsburgh to its conluence with the Mississippi River. In addition to the Great Kanawha (chapter 11), essays on other navigable tributaries include “he Muskingum Navigation” 4

Introduction

(chapter 10) in Ohio, and, with Larry Sypolt, “he Little Kanawha Navigation” (chapter 12) in West Virginia. he Muskingum Navigation extended south by south-east from the Ohio and Erie Canal ninety-one miles through Zanesville, Ohio, to Marietta, Ohio, on the Ohio River. Unlike Ohio’s narrow mule-powered canals, the Muskingum Navigation permitted Ohio River steamboats to enter the interior of Ohio. Studied as early as 1822, it was opened for navigation in 1841 and became a federal waterway in 1887. Sypolt and Kemp detail the history of the Little Kanawha Navigation. Interest in this river dates from George Washington’s eighteenth-century trips to the backcountry. But serious interest did not begin until 1838, when the Virginia General Assembly passed an act directing a survey of the river from its mouth on the Ohio River at what is now Parkersburg, West Virginia, to the Bulltown saltworks. he Little Kanawha Navigation Company was established in 1847, but little was accomplished at that time. In 1864 the charter of the company was amended by the West Virginia legislature so that the company could improve the tributaries of the river as well as the main channel. Mill dams were removed from the river in 1866, allowing boats to navigate the lower river. By 1874 four locks and dams had been built and steamboats were able to ascend the river. Dr. Kemp also writes on contemporary water-transport projects, such as “Building the Tennessee-Tombigbee Waterway” (chapter 7). he TennesseeTombigbee Waterway is a 234-mile-long artiicial channel connecting, in the north, the Tennessee River near the intersection of the Mississippi-AlabamaTennessee state lines with, in the south, the Tombigbee River near Demopolis, Alabama. Constructed from 1972 to 1984 at a cost of almost $2 billion, this was the largest project undertaken by the Corps of Engineers and remains the largest earth-moving project in world history—310 million cubic yards of soil excavated. It may also have been the most controversial project undertaken by the Corps. Widely condemned as a colossal Congressional boondoggle, the Waterway has proved to be useful since opening in 1985, although it has not carried the tonnage initially predicted. Of particular interest in Dr. Kemp’s account of this construction efort is his description of the interaction between project-engineering decisions and the then recently passed National Environmental Protection Act (NEPA). 5

Introduction

he fourth area of Dr. Kemp’s research interests is water supply. he most important water-supply system constructed in the nineteenth century was New York’s Croton Aqueduct. New York City is surrounded by rivers of brackish waters so city oicials had to look to the north to ind a stream of adequate quality and low. In “John Jervis and the Hydraulic Design of the Old Croton Aqueduct” (chapter 9), Dr. Kemp and Edward Winant tell the story of building the forty-mile-long water-supply system from the Croton River to Manhattan. John Jervis (1795–1885) was appointed chief engineer of the project in 1836. he success of such a large project depended on an understanding of hydraulics, but at that time knowledge of open-channel hydraulics was limited and Jervis had to rely on the hydraulics research then being developed by French engineer/ scientists, particularly Antoine de Chézy (1718–1798), Pierre Louis Georges DuBuat (1734–1809), and Gaspard Clair Riche, Baron de Prony (1755–1839). Besides rudimentary analytical tools, Jervis had to contend with major construction diiculties, such as crossing the almost one-mile-long and 100-footdeep Manhattan Valley and, also, crossing the Harlem River without disrupting river traic. At the Manhattan Valley, Jervis used two cast-iron pipes three feet in diameter each and arranged as an inverted siphon. At the Harlem River he built the High Bridge, a 1,450-foot-long viaduct 114 feet above the high-water line of the river, to carry the aqueduct across in an inverted siphon. Construction of this water-supply system began in 1838 and was completed in 1842, with the exception of the Harlem River High Bridge, not completed until 1848. his water-supply system was the irst of its kind in North America and was considered equal to any water-supply system in the world at that time. he twelve essays in this book are a sample of Dr. Kemp’s writings on American civil-engineering history over the last ity years. Written by one of the few American practicing engineers who also research and write on the subject, these essays demonstrate the many contributions and engineering perspective that Dr. Kemp has brought to the subject. Robert Kapsch July 2013

6

1

he 1859 Wheeling Custom House: A Harbinger of Iron and Steel Skeletal Framing Skeletal Iron Framing

T

he Wheeling Custom House stands as an exemplar of nineteenth-century revival architectural styles, as an example of Italianate Palace revival architecture. Gothic revival architecture appeared in ecclesiastical structures such as churches, cathedrals, and educational buildings. At the same time, classical Greek and Roman styles were favored for public and private buildings such as government buildings and banks. he Italianate Wheeling Custom House its clearly into the classical-revival domain. In a narrower context the Custom House stands as an outstanding representation of a large building program that saw numerous custom houses constructed by the federal government.1 Wheeling in the nineteenth century was the second-largest urban area in Virginia, commanding a leading position in transportation sites in America and boasting a lourishing industrial base. It was notable as a center for cast- and wrought-iron products. Until the taming of the Ohio River by an extensive series of locks and dams during the latter half of the nineteenth and early twentieth centuries, the low-water head of navigation on the upper reaches of the Ohio River was a sandbar at Wheeling. As a result, even shallow drat steamboats were precluded from proceeding upstream. Regardless of this restriction, the Ohio River coupled with the Mississippi represented the most signiicant artery of commerce in the nation.2 Beginning in Cumberland, Maryland, the nation’s irst federally inanced road was inished to Wheeling in the second decade of the nineteenth century. It was later extended as far as St. Louis. On Christmas Eve of 1852, ater a 7

Chapter 1

long struggle beginning in 1828, the Baltimore and Ohio Railroad established a direct link from Baltimore to the western waters at Wheeling. By any standards it was an engineering triumph. In fact, it was the nation’s irst commercial railway stretching more than 350 miles. With the opening of the B & O, Wheeling appeared to be in a unique position to dominate a transportation network west of the Appalachian Mountains. Wheeling, in competition with Pittsburgh, was now set to become the leading industrial city on the entire upper Ohio River and to become the gateway to the west. his view of its future was coupled with a lourishing industrial base. Steamboat building, textiles, and a wide range of iron products would be enhanced in 1849 with the opening of the bridge across the Ohio River, the world’s longest suspension bridge, linking the National Road with Ohio.3 It is most signiicant that the Wheeling Suspension Bridge was erected entirely of materials produced locally, including the wire cables that support the 1,010-foot clear span deck. Figure 1.1. Exterior view of restored Wheeling Custom House, 1996.

8

T h e 1 8 5 9 Wh e e l i n g C u s t o m H o u s e

In evaluating the historical features of this bridge and its later companion, the Wheeling Custom House, iron becomes the central feature of both structures and is the essence of this narrative. he engineering aspects of this Custom House and other contemporary custom houses are most signiicant features of the era. A unifying idea is the employment of iron in pioneering ways, raising these federal buildings to one of the pinnacles of nineteenth-century civil engineering. From an aesthetic point of view the buildings clearly show the power and inluence of the Treasury Department of the Federal Government. he Department of the Treasury architect, Ammi B. Young (1798–1874), designed an elegant yet restrained series of architectural gems clothed in his Italianate Palace revival style. he simplicity of the style lent itself well to a desire to produce ireproof buildings.4 One may feel the original aim for ireproof structures was focused on the design of monumental public and private buildings. his appraisal is not correct. Suppression of ire occurred in a more humble situation, namely very large textile mills in England. he irst attempts at suppressing ire were to use cast-iron and hybrid cast-iron/timber beams. his approach proved unsatisfactory because of the brittle nature of cast iron and its tendency to exhibit laws. Early pioneers such as William Strutt (1756–1830) and Charles Bage (1752–1822), in Britain, irst used cast-iron components. William Strutt was associated with the formative years of the development of textile machinery and was involved with the famous Richard Arkwright. In 1781 Strutt witnessed the destruction by ire of a family textile mill in Northampton. At about the same time the Albion Mill in London, noted for its advanced machinery, was totally consumed by ire. Strutt introduced important innovations in ireprooing that later led to ireproof details employed in England.5 Eliminating combustible materials was necessary to ensure resistance to ire. his concern was manifest in applying iron for the principal structural and architectural elements in buildings. Fireprooing presented engineers with an open challenge to ind new uses for iron components, such as doors, windows, shutters, and stairways. Young and Captain Alexander Hamilton Bowman (1803–1865), Army engineer, were aware of these early developments and were committed to producing ireproof buildings as part of the federal building program.

9

Chapter 1

Producing the First Rolled Wrought-Iron “I” Beams in America Numerous attempts were made to overcome cast iron’s predilection to fail in tension when subjected to bending. Cast-iron beams’ strength was enhanced by including wrought-iron tension rods in cast members. hese modiications, nevertheless, resulted in notable failures. Perhaps the most signiicant failure was Robert Stephenson’s (1803–1859) Dee River railway bridge.6 he wroughtiron rods terminated above the neutral axis of the cross section, resulting in large bending moments. he bridge collapsed without warning several weeks ater railway trains irst traversed it. Stephenson’s towering reputation both in mechanical and civil engineering was tarnished by this disaster. What was clearly needed was the replacement of cast iron with a ductile material having both tensile and compressive properties. Rolled wrought-iron components were the answer. Unlike the renowned Wheeling Suspension Bridge, which used only locally produced elements, the use of rolled wrought-iron beams in the Custom House

Figure 1.2. Early cast iron loor beams for mill structures in Britain.

10

T h e 1 8 5 9 Wh e e l i n g C u s t o m H o u s e

required the support of a rapidly expanding industry. Beginning with Henry Cort (1740–1800), the well- known British ironmaster, the puddling and rolling of wrought-iron shapes supplanted the ancient crat of blacksmiths, who could produce wrought iron only in very limited batches, with what was essentially a new structural material.7 he successful application of wrought-iron beams to structures was a root cause in the nineteenth century of the separation of engineering from the ancient profession of architecture. Puddling furnaces enabled the production of wrought iron in greatly increased quantities and quality, allowing structural shapes to be rolled for the irst time. Structural cross sections, especially “I” sections, were not available in suficient depths because of diiculties associated with rolling complicated crosssection shapes. Attempts were made, nevertheless, to produce wrought-iron beams using available railway rails or built-up riveted plate girders. Although not very eicient as beams, adequate bending capacity could be obtained by joining the bases of two rails at their bottom langes. he result was a curious cross section with the rail bulbs forming the top and bottom chords. What was Figure 1.3. Attempts to produce stronger and stifer components using readily available iron members.

11

Chapter 1

clearly needed was a much more eicient cross section, in a word, the “I” beam. he Trenton Iron Works, in Trenton, New Jersey, produced paired rail beams, which were used in the Cooper Union Building in New York City (1880). Peter Cooper, of locomotive Tom humb fame, was the founder of the Trenton Iron Works.8 Faced with iron ore of questionable quality, Cooper sold his business and relocated at Phillipsburg, New Jersey, and upstate New York. he most signiicant result was the use of anthracite coal as the fuel and powerful reducing agent to produce pig iron. he puddling furnaces, also using anthracite, oxidized the carbon in the molten iron to nearly zero percent, producing a bloom of wrought iron that could then be rolled into several simple shapes. Cooper’s major contribution was the rolling of “I” section members—irst a seven-inchdeep beam, and later, ater much struggle, a nine-inch-deep beam, which was employed throughout the Wheeling Custom House loor system.9 Always seeking business opportunities, the move to Phillipsburg for the production of cast iron and puddling furnaces to produce wrought iron was a success, since anthracite coal and iron ore were readily available by water and later railways. It was Cooper’s intention to set up his son, Edward, as the leading administrator of product development for this newly relocated enterprise. Edward, however, ofered the alternative of a partnership with Abram S. Hewitt. his partnership is a noteworthy example in which each ofered the necessary business and engineering expertise to establish a leading iron-industry irm. his partnership, established in 1845, managed the production of the Trenton Iron Company, enjoying a close working relationship with the burgeoning railway construction in America. he primary product, leading all others, was wrought-iron rail. he basic product for the company was not to last in the face of intense British competition. he B & O was constructed with rails imported from Wales at substantially lower prices. hus, the company sought to develop a more-promising product. One notable example is the production of wire, not only for agricultural purposes, but for the newly introduced iron-wire suspension-bridge cable. From the point of view of iron skeletal framing, the Trenton Iron Company turned its attention from cast-iron technology to wrought-iron products, which formed the technological basis for the construction of the federal building program of the 1850s. As an example, the obstacle that presented itself in producing an eicient beam was the rolling of a double-lange “I” section. By turning 12

T h e 1 8 5 9 Wh e e l i n g C u s t o m H o u s e

his attention to building systems, Peter Cooper recognized that wrought-iron rail beams lacked suicient strength to support brick jack arch loors.10 In 1847, the Trenton Iron Company undertook to produce a seven-inch-deep “I” section for use in buildings. he attempt was quite unsuccessful and cost the company 30,000 dollars. A few years later, in 1852, a second attempt was made, which this time proved successful. A recent British immigrant, William Barrow, is credited with the successful rolling of loor beams. Such new nine-inch “I” beams were to ind wide application for reasonable spans of twenty feet or more, which was usually suicient for buildings of the time.11

Structural Details of the Construction of the Wheeling Custom House of 1859 For purposes of this narrative, only the salient features of the Custom House skeletal framing are presented. he entire interior structure consists of shallowrise brick jack arches of ive-foot span resting on the recently produced nine-inch Figure 1.4. Wrought iron rail-beams used in the Cooper Union Building and a typical iron box girder. his is an early attempt to use railway rails in a system that predates the use of the 1856 9-inch beams. 13

Chapter 1

“I” beams. his loor system was supported on hollow wrought-iron box girders spanning between cast-iron columns. Since the vertical loads from the loors and the roof were loaded in compression, cast-iron columns were used throughout the structure. hese columns were designed with excessive factors of safety—as high as eight—to sustain any secondary bending stresses. An ingenious architectural feature involved a central heating system in which hot air from boilers in the basement rose up the hollow columns and was distributed laterally through the box girders and loor registers. Provisions were made at each loor by a cast-iron stool to distribute hot air into the girders as well as passing to the upper loors. his system would be described as background heating only, necessitating ireplaces in each room. he ireplaces were clothed in imported Italian marble. Framing the entire roof structure are seven-inch rail beams, which supported corrugated iron sheets to resemble Italian terracotta tile roofs. All of the lower loors utilized jack arches and nine-inch wrought-iron beams. he entire loads of the loors were simply supported on riveted wrought-iron box girders which, in turn, rested on the cast-iron columns. he central corridor has a iteen-foot clear span in contrast to the twenty-foot span in each of the side rooms. It was decided to restore the building as it was on 20 June 1863, the day that Abraham Lincoln signed a document creating the new state of West Virginia Figure 1.5. Stacked wrought iron bars, heated and then rolled into a 9-inch “I” beam.

14

T h e 1 8 5 9 Wh e e l i n g C u s t o m H o u s e

out of the western portion of the Commonwealth of Virginia. Ater the building passed into private hands, a fourth loor was added together with an additional bay on the south end of the building. hese additions were removed to meet national preservation standards. To meet modern ire-safety standards for public access, a ire stairwell was incorporated in the restoration work. When the restoration work was undertaken several short sections of the original beams became surplus. his enabled the writer to determine, by structural tests, the properties of these loor beams. his work included tension tests as well as metallurgical investigations undertaken at Loyola College, Baltimore. At the same time an investigation of the bricks and mortar determined their physical properties. he metallurgical tests of the beams revealed internal transverse cracks in both langes. At irst it appeared that additional material had Figure 1.6. Typical joints in the Custom House showing 9-inch loor beam, jack arches on 5-feet centers, iron box girders, and cast iron columns.

15

Chapter 1

been added to the beams while they were being rolled. A later and more considered opinion rests on the rolling procedure used. Because of limited capabilities it was not possible to produce a bloom with suicient material to roll such components. he alternative consisted of stacking a series of plates, which were heated in an attempt to forge all of them into one homogeneous iron bloom. It appears that at the junction of plates was an internal defect, producing a zone of weakness, which on examinations appears as the deining line between the plates. In simple terms it means that the beam needed to be rolled at a much higher heat to obtain a satisfactory forging. Nevertheless, it can be shown that the Trenton Iron Company proved that their products were quite superior to typical values for wrought iron. he only exception is the higher percentage of slag in the Trenton sample. It appears that the puddling process used did not lower the amount of slag, suggesting that they did not remove additional slag, as would normally be expected. An equally important factor rests on the amount of slag in the company’s cast-iron pigs used to produce wrought iron. With the noticeably lower carbon content, the Figure 1.7. Metallurgical tests showing internal cracks.

16

T h e 1 8 5 9 Wh e e l i n g C u s t o m H o u s e

beams’ ductility was unafected. One must conclude that these beams were of superior quality and eminently suitable for use in the Wheeling Custom House. In fact, these beams are nearly sole survivors of the irst rolling of the modern “I” beam. As a result, they need to be celebrated as a most signiicant event in nineteenth-century American engineering and architecture. he entire loor system consisting of wood looring and leveling cinders on top of the jack arches provides a level surface for the looring. he jack arches and associated nine-inch iron beams are carried on wrought-iron box girders as mentioned above. While not revolutionary compared with the nine-inch beams, the box girders are of considerable interest in the skeletal framing system. hese box girders are related closely to plate-girder bridges, which are so prevalent, even today, and have been built by the thousands in North America. Using standard plate-girder analysis it has been revealed that these girders are able, at the very least, to sustain a uniformly distributed live load of 114 pounds per square foot. his strength provides an excess compared to the 100 psf loor load speciied in building codes for public assembly. he columns carrying the loads of the girders were designed using large factors of safety. In selected cases, these factors were as high as eight, as mentioned above, providing excess strength to support the loor loads. Since the ends of the loors are supported on the exterior stone walls, all lateral wind loads are sustained by these stone masonry walls. As a result, only the interior structure can be said to employ skeletal iron framing. Having considered the iron components of the framing, it is necessary now to evaluate the jack-arch loors. Tests undertaken on the brick revealed a compressive strength of 6,050 psi, which is more than adequate. According to the original speciication, the loors were to be laid up with natural-cement mortar. Extensive examination of exposed sections of the jack-arch loor mortar conirmed the use of natural cement, which has a typical brown cast, unlike the grey Portland cement. In the absence of mortar samples for testing purposes, a value of 240 psi working stress was selected from extant concrete speciications. It can be shown that this strength is more than adequate.12 Although no written records exist, it is presumed that the loors were erected in the following manner: the nine-inch iron beams were installed on ive-foot centers, starting at the exterior stone wall, and then the jack arches were installed one panel at a time progressing throughout the length of the building. With this 17

Chapter 1

procedure there is a strong possibility that the exposed beams could turn over from the thrust of the arches. his was prevented by iron stabilizing bars connecting the top of the loor beams; these bars are still in place. his was clearly necessary since the loor beams were installed without a joint on the girders. When all of the loors were constructed these stabilizing bars were no longer necessary since the loor beams themselves prevented overturning by the adjacent stone wall. At the beginning of the restoration project Tracy Stevens, the restoration architect, engaged an engineer to perform a structural analysis of the loor system. he model used was a traditional approach by considering each structural element serially. Ater considering the shape of the jack arches, attention was focused upon the wrought-iron loor beams, which were assumed to carry the entire load of the jack-arch loors. his simple analysis showed that the loor Figure 1.8. A.B. Young’s second loor plan. Floor plan shows twenty-foot span in the business rooms together with 15’ spans in the corridor.

18

T h e 1 8 5 9 Wh e e l i n g C u s t o m H o u s e

system could support 74.4 pounds per square foot, whereas code requirements mandated 100 psf. At the time loor loads were limited in an efort not to overstress the iron beams. Another later and more sophisticated analysis considered the entire loor system, spanning longitudinally twenty feet. In this model the top of the jack arch serves as the compression zone of a composite beam in which the iron beams provide tensile strength. Using the physical properties of the components a inite element analysis (algon 20) provided stresses. In essence the hybrid loor system was divided into a series of closely spaced elements, hence the name inite element. he analysis yielded very acceptable results. For example, the center delection was 0.56 inches with a span of twenty feet. he load-carrying capacity of 124 pounds per square foot is comfortably in excess of the required 100 psf. At the interface of the mortar against the web of the beams the stress was found to be 142 psi, well below the safe value of 240 psi or less. As a result of the analytical work as well as extensive documentation of the structure, it can be concluded that the Wheeling Custom House can sustain the 100 pounds per square foot required design live load at all levels of the structure.13 he Trenton Iron Company freed engineers from the widespread cast-iron era.

Liberty and Union Many factors exemplify the historic nature of this most notable structure. As an important site in the history of nineteenth-century engineering and architecture, one can look to the following salient points in its history: 1. he birthplace of West Virginia as a new war-born state occurred in the courtroom of the Custom House. As a solid supporter of the Union cause during the Civil War, the slogan of “Liberty and Union” seems entirely appropriate. 2. he restored building can be saluted as a premier example of the federal building program of the 1850s. Its modest Italianate Renaissance Palace style represents a leading chapter in the history of architecture. Certainly a large part of its historical signiicance is the result of careful design, utilizing a new structural material, wrought iron, designed according to 19

Chapter 1

a widespread goal of ireproof structures. Ammi B. Young, architect, and Captain Bowman, engineer, should be celebrated by those interested in the birth of wrought iron, and later steel, construction in America. he nine-inch “I” beams available from the Trenton Iron Company marked the transition to iron-framed structures. 3. It would be false to ascribe to this building the claim to be the only progenitor of the skyscraper. Its form was not a true skeletal building because it was supported on external stone walls. In evaluating the contributions of American engineers in the nineteenth century, three topics emerge: the transformation of the all-timber-truss bridge into unprecedented long-span iron, and later steel, truss railway bridges; America’s leadership role in the design and construction of long-span suspension bridges, which lasted more than a century; the rise of the skyscraper in the latter third of the nineteenth century. he 1859 Wheeling Custom House is an outstanding example of an integrated approach to architecture and should be heralded for its place in civil engineering. Chapter 1 Notes 1 2

3 4

5 6

James Guthrie, Letter in House Executive Documents, 36th Congress (Washington: U.S. Government Printing Oice, 1860), 8. Emory L. Kemp and Beverly B. Fluty, he Wheeling Suspension Bridge: A Pictorial Heritage (Missoula, MT: Pictorial Histories Publishing Co., 1999), 1–6; A.W. Skempton et al. (eds.), A Biographical Dictionary of Civil Engineers in Great Britain and Ireland (London: homas Telford, 2002), 28–29, 670–672. Rebecca Harding Davis, Life in the Iron Mills (1861; New York: Feminist Press, 1972.) Sara E. Wermiel, he Fireproof Building: Technology and Public Safety in the Nineteenth-century American City (Baltimore: Johns Hopkins University Press, 2000). Tom Swailes and Joe Marsh, Structural Appraisal of Iron-ramed Textile Mills (London: Institution of Civil Engineers, 1998), 12–19. Skempton, Biographical Dictionary, 661–668.

20

T h e 1 8 5 9 Wh e e l i n g C u s t o m H o u s e

7 8

9 10 11

12 13

Charles Singer et al., A History of Technology (New York & London: Oxford University Press, 1958), 81. Robert Jewett, “Solving the Puzzle of the First American Structural Rail-Beam,” Technology and Culture 10, no. 3 July 1969, 371-391; Allan Nevins, Abram S. Hewitt: With Some Account of Peter Cooper (New York & London: Harper and Brothers, 1935). Esmond Shaw, Peter Cooper and the Wrought Iron Beam (New York: Cooper Union, 1960). Charles E. Peterson, “Inventing the I-Beam: Richard Turner, Cooper & Hewitt and Others,” Association for Preservation Technology Bulletin 12 (1980): 63–95. Charles E. Peterson, “Inventing the I-Beam, Part II: William Barrow at Trenton and John Grifen of Phoenixville,” Association for Preservation Technology Bulletin XXV (1994): 17–32. George Hool et al., Concrete Engineers Handbook (London: McGraw Hill, 1918). Patricia C. Tice, An Evaluation of the Skeletal Structure of the West Virginia Independence Hall, 1859 (Morgantown, WV: Institute for the History of Technology and Industrial Archaeology, 1995).

21

Figure 2.1. Portrait of Charles Ellet Jr. (Huntington District Archives, U.S. Army Corps of Engineers)

2

Charles Ellet Jr. (1810–1862): Portrait of an Engineer

C

harles Ellet Jr., by any measure, was a leading engineer in the nineteenth century in America. Ellet made innovative engineering solutions to a wide range of engineering enterprises including river hydraulics and navigation. His work on the laws of trade was revered by economic historians. Perhaps he is best known for his work on suspension bridges. hrough his notable 1,010-foot span of the Wheeling Suspension Bridge he received worldwide notoriety. His practical engineering work was supported by numerous publications, emphasizing the promotion of long-span wire suspension bridges. He ended his life as a colonel in the Army charged with building a ram leet on the Mississippi, which was instrumental in the Union victory at Memphis. He died of wounds received at the battle, ighting for the Union cause. Following his untimely death, his devoted wife died shortly thereater in a melodramatic Victorian way worthy of any nineteenth-century romantic novel. he cause of her death was attributed to a “broken heart.”

he Early Years Like so many famous men, Ellet was particularly inluenced by his mother.1 His grandfather, born 1746, descended from Israel Israel, diamond cutters from Amsterdam in the Netherlands. While only educated in what would be characterized as a inishing school for young ladies, Ellet’s mother, Mary Israel, was the greatest inluence in his life until he married. He was dependent on her advice for all decisions made in his early life. Born in 1780 in Philadelphia, Charles Ellet Sr., who married Mary Israel in 1807, was of Quaker background and an ironmonger. Apparently he was a direct 23

Chapter 2

descendant of Samuel Carpenter, an associate of William Penn. At the beginning of the nineteenth century in America anti-Semitism had not appeared as a national issue. In all the years of research on Ellet by the author, he has not discovered any reference to his Jewish background. his minority inheritance may explain, in part, Ellet’s stubborn reluctance to be associated with either public or private groups.2 In his renowned book Sir Samuel Smiles undertook in biographical form to present what he considered to be the ield marshals of the Industrial Revolution.3 Smiles believed this was the golden age of the essentially new profession of civil engineer. he ascendancy of the profession marked an increasing separation from the time-honored profession of architect. Smiles presented engineers in a traditional mould of master cratsmen in the new industrial era. As a rugged individualist Ellet its the epitome of such a ield marshal. As a result, he always served as an individual consultant. A case in point was the struggle to establish a society of engineers in 1852. Ellet thought such a society as the American Society of Civil Engineers would promote mediocrity in the profession. hus, there was no organized partnership to carry on his professional work ater his death. his is in stark contrast to such a irm as Boulton and Watt or, in our day, Rolls and Royce. he irm of J.A. Roebling Sons and Company promoted the almost legendary career of their founder. he well-known engineer D.B. Steinman claimed that Roebling rebuilt the Wheeling Suspension Bridge ater its collapse in a tornado. It was, in fact, Ellet and his associate McComas who rebuilt the bridge shortly ater its collapse in 1854. he confusion occurred in 1860 when Washington Roebling rebuilt the structure in much the same appearance as today. here is clearly a gap in the record of Charles Ellet Jr. regarding his early years except for a memoir by his mother written in her ninetieth year. In contrast, there is a rich source at the University of Michigan of original correspondence between Ellet and his wife. He wrote almost daily when he was away from home. his extensive archive could be the basis of a book of professional life in the antebellum period. here is also a large amount of material there focusing on the tensions between North and South. Ellet and family were staunch supporters of the Union whereas his wife, Ellie, whose family owned slaves in Virginia, favored a negotiated settlement. Few technical or political details can be found in this correspondence. 24

Charles Ellet Jr .

Although many aspiring young men “made good” without the beneit of an education, Ellet excelled as few others in self study, and his admission to the École Ponts et Chaussées in France, the world’s leading institution of engineering, must be considered exceptional.

A Young Novice Wishes to Be an Engineer he time was ripe for an ambitious youngster wishing to enter the ield of engineering. With the completion of the Duke of Bridgewater’s canal from Worsley to Manchester, England, in 1763, a veritable mania of canal building ensued. his leading technology was inally overtaken by steam-hauled railways. A later and similar action swept the United States with the opening of the Erie Canal in 1825. In America it was to be a short but very intensive period of canal building, which met a challenging entry into the transport ield with the incorporation of the Baltimore and Ohio Railroad in 1827, beginning operation in 1830. he surge for independence and the subsequent western movement were the hallmark of the celebrated historian Frederick Jackson Turner (1861–1932), who introduced this concept in a paper in the Columbian Exhibition in Chicago in 1893. Albert Gallatin, the Swiss émigré who became secretary of the treasury, delivered to Congress a widely circulated report on internal improvements in 1808.4 Many believe that this was the beginning of the internal improvement movement, but it really traces its origins to before the American Revolution by such stalwarts as George Washington and homas Jeferson. Gallatin’s report was in fact a plea for the federal government to become an active partner in this national initiative. his vision was dimmed and the federal government played only a negligible role and did not rationalize a network of transport development. Instead of being guided by the federal government as urged in the report, state and local governments, in partnership with private enterprise, took charge of the movement. Each east-coast port city from Boston to Norfolk, Virginia, attempted to support eforts to have direct communication with the former Northwest Territories and a connection to the western waters served by the Ohio River.5 A large number of transport projects resulted in not only the Erie Canal but also the Pennsylvania Mainline Canal and the establishment of the Chesapeake 25

Chapter 2

and Ohio Canal. Farther south, Washington supported the James River and Kanawha Canal.

Life on the Canal In the early stages of the Industrial Revolution, entry into the ield of engineering usually involved learning land surveying. he art of surveying rested irmly in the hands of lawyers who were responsible for real-estate transactions, deed research, and the need for land to be surveyed, mapped, and recorded.6 hus it was that Benjamin Wright, James Geddes, and Canvass White were lawyers largely responsible for building the Erie Canal. At age seventeen young Ellet joined a surveying party for laying out branch canals along the Susquehanna River valley. A surge of canals was projected with the beginning of the Pennsylvania Mainline Canal on July 4, 1826. With state support of this vast undertaking it is little wonder that political forces bolstered by local support dominated canal construction. Into this scene Ellet entered, serving as an assistant engineer. he Ellet archives do not reveal how he obtained this position even though he had no experience in canal building. Ellet entered a burgeoning profession during a golden age of individual engineers not encumbered by associating with large corporations or government agencies. here is a dearth of information on Ellet’s involvement in canal building other than his surveying work, including layout and design. On June 23, 1828, the C & O Canal Company hired Benjamin Wright as chief engineer because of his outstanding work on the Erie Canal. Ellet, always interested in professional advancement, moved to Maryland irst as a volunteer and then was hired as an assistant. A vociferous argument erupted between the C & O Company and the Baltimore and Ohio Railroad, both pushing west to join the western waters at Wheeling. he controversy focused on giving a right-of-way through the narrow gap just downstream from Harpers Ferry. he dispute was later settled with both companies securing a passage through the gap. Ellet, however, felt that the injunctions levied against the canal would severely limit its western progress. It was time for Ellet to move on. Beginning in Maryland, Ellet indicated that he needed further education. He considered either moving to Illinois or going abroad. It was during this time that he studied French and advanced 26

Charles Ellet Jr .

mathematics. He received a grant from his grandfather, Israel, to match $500 given by his mother to go to France. Despite wishing his sons to engage in agriculture, in a most interesting turn of events his father supported his plan and produced a letter of introduction to none other than the Marquis de Lafayette and an unnamed American businessman in Paris. Landing at Le Havre, Ellet arrived in Paris in 1830 and was admitted to the world’s leading program in engineering, the École Ponts et Chaussées. his is an extraordinary achievement since he had no formal education. A typical French student entered the École Polytechnique, founded in 1794, for study in science and mathematics, and ater inishing his program the student entered a one-year postgraduate course at the École Ponts et Chaussées. Counted amongst the many illustrious professors were Louis Marie Henri Navier (1785–1836); the legacy of his late uncle Gauthey (1732–1806). Coulomb (1736–1806) also deserves mention. From the point of view of Ellet’s developing career one stands out in this galaxy of stars, and this was Navier. He established a towering reputation as a lecturer, serving as both a mathematician and an engineer. He brought fresh new insights to the theoretical nature of structural analysis. In 1823, Navier published what is the most signiicant treatise on suspension bridges ever written, entitled Sur les Ponts Suspendus.7 It was this encounter with Navier’s work that irst evidenced Ellet’s commitment to suspension bridge design. Ellet spent the spring of 1831 touring public works in France, especially suspension bridges. As early as the iteenth century Chinese bridge builders employed iron-cable suspension bridges suitable for wheeled vehicles. he irst modern suspension bridge was erected across Jacob’s Creek in western Pennsylvania. he inventor was Judge James Finley and the date was 1801. Finley’s bridge employed wrought-iron chains and suspenders together with a level deck and stifening trusses lanking each side of the roadway. It had all the elements of the modern suspension bridge. In his drawing of the Wheeling Suspension Bridge Ellet included, as an insert, an early Latin-American suspension bridge at Penipe. Ellet was aware of Finley’s work but relegated his bridges to the early primitive stage of suspension bridge technology. His position was clear that he was a prophet of an exciting new technology based on sound mathematical analysis utilizing the superior properties of drawn iron wire. It should be noted that the French excelled in this type of bridge. By the mid-nineteenth century they had erected 500 wire suspension 27

Chapter 2

bridges under the aegis of the national department of Ponts et Chaussées. While we know the names and locations of the bridges Ellet visited, no comprehensive evaluations of these structures has appeared. Recently, Donald Sayenga, Tuscon, Arizona, and Michel Cotte, Tournon, France, have undertaken a study of all of these bridges visited by Ellet. He also inspected the Canal du Midi connecting the Bay of Biscay with the Mediterranean Sea, as well as a summit reservoir associated with this canal. With his interest in cultural matters, it is not surprising that during his Grand Tour he also visited sites in Switzerland, including Voltaire’s house. In a lighter vein it must be reported Voltaire said he was dying of 100,000 cups of cofee. Following his tour Ellet returned to America in early 1832 as a veritable missionary for French-style wire suspension bridges. His irst opportunity came shortly ater his return when the federal government advertised for a suspension bridge across the Potomac at Washington. Ellet responded with a design featuring two long spans of 600 and 592 feet respectively. It was a notable design worthy of the French tradition. Submitted on June 6 to the secretary of the treasury, ater the closing date for designs, the belated design was rejected; Ellet commented that the nation had lost a splendid opportunity. Rather than an elegant solution the government chose to repair the dilapidated wooden bridge.8 With his suspension bridge design rejected, Ellet accepted a position of assistant engineer on the construction of the Utica and Schenectady Railroad, with responsibility for engineering design for the western end of this line. With little prospects for building bridges he was once again a surveyor on this railway, intended to link Albany with Bufalo. His work, following an alignment that is still used today, was apparently of a high standard. In the spring of 1834, Judge Benjamin Wright appointed Ellet to survey the western extension of the New York and Erie Railroad. his led him to his irst publication, in 1835—a fortynine-page report detailing the work that led to the location of the line. At about the same time, Judge Wright became the chief engineer of the James River and Kanawha Canal Company. his large public works was intended to connect tidewater on the James with the Ohio River by way of the Kanawha River.9 A comprehensive study of the route was published detailing the engineering aspects of what would become a canal-turnpike system. 28

Charles Ellet Jr .

During the Great Depression of the 1930s, the French-style movable dams from the Falls of the Kanawha to the Ohio River were replaced by patented roller gates according to a German license. his aspect of the James River and Kanawha Canal continues today as one of the most successful river navigations in the country. Following Ellet’s assignment, a grand scheme to develop the Central Water Line system, with government support, was proposed. In essence it would develop a waterway from tidewater Virginia to the foothills of the Rocky Mountains, via the Ohio, Mississippi, and along the Missouri rivers.10 he noted railroad engineer Moncure Robinson was opposed to the James River and Kanawha Canal and it is almost amusing to note that ater a vociferous debate Ellet’s relationship with Robinson became so intense that Ellet was challenged to a duel, which never took place.11 With the economic panic of 1837 there was concern that all work would cease on the canal. During this period Ellet produced a lengthy report on the canal and its future. Ellet’s rapid rise in the profession was accelerated by his appointment as chief engineer on the canal following Wright’s retirement in 1838. As chief engineer he requested and received a salary increase to $8,000, a substantial amount for any engineer in the antebellum period. Ellet published a book, presumably written in 1838, called Laws of Trade, which is still referred to by economic historians.12 Perhaps the most important event during his sojourn in Virginia was his wedding to Elvira Daniel. She was a member of a prominent family in Lynchburg. Figure 2.2. Ellet’s drawing of proposed Potomac River suspension bridge. (Smithsonian Museum of American History) 29

Chapter 2

hey were married on October 31, 1837, at Point of Honor, the Daniel home in Lynchburg, which stands today on expansive grounds. Ellie, as he called her, was clearly the love of Ellet’s life. She devoted her life to her husband until his death. Although Ellet prepared a report on the canal in January 1839, in the same month the board of directors failed to reappoint him as chief engineer. he reason for his dismissal was a mystery to both Ellet and subsequent historians. He let the company without delay and went west to join his brothers, Alfred and Edward, on their farm in Illinois. He was engaged to design a 3,000-footlong suspension bridge across the Mississippi. Ellet’s bold design featured a 1,200-foot-long span over the main channel of the river, lanked on both ends of this span by two side spans of 900 feet each, making a total length of 3,000 feet. While Ellet was involved in the analysis of such a bridge, neither he nor anyone else really had a grasp on building the river piers for such a bridge. he pneumatic caisson later used by Colonel Eads made the construction of the river piers possible. he Eads extant iron arch bridge was completed in 1873. For Ellet’s sake and the history of suspension bridges it is a mercy that this design was not built since one of the many Mississippi freshets would in all likelihood have destroyed the towers. Probably relieved to be rid of what some thought a preposterous proposal, Ellet was paid a consulting fee of $1,000 and the matter was dropped. Born in Riedlingen, Germany, in 1769, Lewis Wernwag immigrated to America in 1786. By dint of ability and perseverance he became a renowned timber-bridge builder. He later was employed for work on the Baltimore and Ohio Railroad. In 1818, construction was completed on his celebrated 340-foot-span covered bridge entitled Colossus, over the Schuylkill River in Philadelphia. To many it was the handsomest covered bridge ever built. On his way to Illinois Ellet had stopped in Philadelphia, where he submitted a plan for a suspension bridge as a replacement for the Colossus, which had burned in 1838. If ever promotional publications paid of, the replacement of the Colossus can be linked to the publication of Ellet’s A Popular Notice of Suspension Bridges.13 A former assistant engineer to Ellet submitted a plan largely based upon Ellet’s pamphlet and, most signiicantly, employed John A. Roebling. he award of the contract to Ellet was to be a source of friction between himself and Roebling. Ater arriving in Illinois, Ellet was delighted to hear from his mother that he had won the Schuylkill River Bridge contract. 30

Charles Ellet Jr .

he stone abutments of the Colossus were undamaged by the ire which had consumed the superstructure. As a result Ellet had to design only the superstructure and not be concerned about building stone abutments. he project required only six months to complete and was opened in January 1842. It was a triumph for Ellet, which placed him in the forefront of suspension-bridge engineers. Both the bridge and its designer enjoyed public acclaim. With its main span of 340 feet and a 25-foot-wide roadway and two 4-foot walkways, he had succeeded in erecting a most impressive bridge, which in every detail represented a French wire suspension bridge on American soil. Basking in his triumph, he took his family to Cuba for the winter of 1842–1843, returning from Havana in the spring of 1843. A year later, in May 1844, Ellet let for a return visit to France and also Britain to inspect the latest advances in transport technology, especially railways. During his stay in France, he corresponded with businessmen in Wheeling about a suspension bridge across the Ohio River. Figure 2.3. Ellet’s drawing comparing the Swiss Freiburg Bridge with his Schuylkill River Bridge. (Smithsonian Museum of American History) 31

Chapter 2

his was one of many monumental sites he envisaged for long-span suspension bridges. As part of his tour Ellet agreed to secure a loan for the Schuylkill Navigation Company, but he was later informed that all of the stock had been sold and his services were not needed. he outstanding British engineer Isambard Kingdom Brunel was then involved in building an atmospheric railway on the South Devon Railway. he atmospheric system was designed to propel trains over the challenging south coast of Devon. Ellet mentions this exciting new technology in a report that has disappeared. It is a shame that these two engineers never met. In many ways they were kindred souls. Brunel is celebrated in Britain as an eminent Victorian while Ellet is largely unknown by the general public. In 1845, Ellet learned that he would not be appointed as chief engineer of the Schuylkill Navigation Company. In a somewhat surprising move he accepted an appointment in a minor role with the company to prepare a promotional pamphlet extolling the beneits of the proposed enlargement of the navigation. His appointment was for four months at a mere $250 per month. It was during this appointment that Ellet defended canal transport as opposed to steam-hauled railways. He apparently did not see a bright future for the iron horse—or was he serving a propaganda role for the company? By December 1845, Ellet secured an appointment as president of the navigation company for its enlargement. he work was successfully completed by October 1847, when Ellet resigned to pursue his promotion of suspension bridges. Although now it would be considered illegal, a business “pool” ended the conlict between the canal and railway by adjusting the tonnage carried by each and regulating transport rates.

he Two Big Bridges Ater more than a decade since Ellet proposed to bridge the Niagara Gorge, Charles Stuart, a friend of Ellet, organized two bridge companies, one in the United States and the other in Canada. he joint companies sought construction bids to provide a railway access across the gorge. Stuart envisaged connecting the Great Western Railway of Canada to the New York Central Railroad. By crossing Ontario a direct link would be established between New York City and Chicago. 32

Charles Ellet Jr .

During the summer of 1843, while stock in the bridge companies was being sold, the two most qualiied engineering consultants, Ellet and Roebling, became rivals in extolling the merits of each of their designs.14 Roebling submitted the low bid of $170,000 but the contest was not over. Ellet submitted a new proposal in which he agreed to acquire $30,000 in the bridge company’s stock. As a result Ellet gained a marked advantage over Roebling when it was announced that Ellet had received the contract for the Wheeling Suspension Bridge. Ater much negotiation the Niagara contract was awarded to Ellet. hus, in a stroke he was in charge of the two most sought-ater suspension bridge projects in the country. He was given a free hand in the details for the highway/railway bridge but the contract was not to exceed $190,000. he approved design required a clear span of 800 feet over the deep gorge of the Niagara River. he design showed a railway track in the middle with a walkway for pedestrians on each side. he designed live load was established at a generous 200 tons with a limit for train loads of 35 tons. With substantial outcroppings of rock the abutments on either side supported heavy timber frames to serve as towers. With the touch of a showman, Ellet organized a kite-lying contest to carry the irst wire of his cables across the gorge. Subsequently, more and more wires were drawn to form the multi-wire cables. Ater forming the irst cable, Ellet had a wrought-iron basket constructed, rather like a garden swing, to accommodate four people suspended from the cable and be drawn across the gorge. He collected tolls on Figure 2.4. Schuylkill River Bridge at Fairmount showing the garland system, completed by Ellet 1842. (Free Library of Philadelphia) 33

Chapter 2

what must have been a hair-raising and yet spectacular ride. Ater the erection of a single footbridge without handrails, tolls were also collected. Although it may be apocryphal, Ellet is credited with driving a horse and carriage across the bridge. As one might expect, Ellet refused to deliver the tolls to the bridge companies. At one point during a heated debate on tolls, he mounted a cannon on the American side to end the debate by force. Mercifully, armed conlict between the companies was avoided and in a larger sense between the two countries. his debate was described by Ellet as a “hornet’s nest.” he irst footbridge blew down in a windstorm and was replaced by a second version in which tolls were again charged. In the case of modern construction projects the bridge and its impedimenta belong to the contractor until the inished product is handed over to the owner. he second footbridge, opened in July 1848, cost Ellet $30,000. A month later, during the absence of Ellet, who was probably on business in Wheeling, the American company under the leadership of Lot Clark dismissed him as chief engineer. Law enforcement oicials took possession of the bridge on behalf of both companies. Having lost the original contract to Ellet, John Roebling took over the project with a magniicent structure featuring a double deck. he top deck was for rail traic and the lower one served as a highway. It should be noted that the truss was fourteen feet deep, providing suicient torsional and lexural strength to resist excessive wind pressure, which caused the demise of many early suspension bridges. Several years ago, Janet and Emory Kemp together with Donald Sayenga examined the buttress erected by Roebling, which conirmed that his abutment straddled the earlier work of Ellet. he examination disclosed wire used by Ellet. In 1877–1886, L.L. Buck completed extensive repairs of Roebling’s Niagara Suspension Bridge. Later, he headed a project that replaced Roebling’s bridge with a huge two-hinged steel arch.

he Wheeling Bridge Relieved of his responsibilities at Niagara, Ellet was free to concentrate his energies on the completion of the Wheeling bridge. Equally important for a man of delicate disposition, the elimination of the grueling 300-mile commute from Niagara to Wheeling was eliminated. hese journeys involved being a 34

Charles Ellet Jr .

passenger on primitive wagon roads. For example, it took three days by road from Erie to Wheeling. Unlike Niagara, no antagonisms developed between the directors of the Wheeling and Belmont Bridge Company and the consultant. It should be noted that the Wheeling Suspension Bridge was entirely in Virginia (later West Virginia). Perhaps the most succinct way of describing Ellet’s greatest triumph, the Wheeling Suspension Bridge, is to irst deal with the construction and engineering aspects of the structure and then to address the legal battle that ensued over navigation rights on the Ohio River. his legal battle dragged on well ater the completion of the bridge. Roebling and Ellet were the chief contenders for the bridge design. Roebling proposed a rather restrained design featuring a pier in the middle of the Ohio River. Ellet’s design, on the other hand, required a 1,010-foot span exceeding the longest span of the 870-foot bridge at Freiburg, Switzerland. Ellet’s design was both elegant and economical at the same time. It is clear that the best design won. Whereas the production of wrought iron and later steel components was not available in America till ater the Civil War, wire on the other hand could be readily produced from wrought-iron billets by a system of drawing the wire through successively smaller dies, requiring only a minimum of equipment. Figure 2.5. Ellet’s drawing of the Niagara Suspension Bridge. (Smithsonian Institution Museum Archives) 35

Chapter 2

With drawn wire produced by Bodley & Sons in Wheeling all the elements required for building the bridge were readily available in the region. Unlike large structural iron members whose length was limited by the size of the rolling mill, wire could be produced “by the mile.” By drawing the wire through successive ever-smaller dies, the wire was strengthened by work-hardening. While good-quality wrought-iron members had a breaking stress of 48–50,000 psi, the drawn wire had a breaking stress of 83,000 psi. Another signal advantage of wire was that it was proof-tested in the drawing operation. hus, any laws would be readily observed. In typical French fashion, individual cables were assembled on the town side of the bridge. hese parallel wire cables consisted of 550 wires made into 4-inch-diameter cables. In the case of Roebling’s work the cables would be spun in place by his device called a spider. Roebling is oten credited for cablespinning techniques, whereas spun cables were employed by the French engineer Joseph Vicat long before Roebling was active in the ield. Ellet’s parallel wire cables, six on each side of the bridge, were hauled into position over the top of the two towers and fastened to abutments at each end of the bridge. he individual cables remained separate in what the French called the garland system. he deck loads were carried to the cables by suspenders. In some of the earliest bridges suspenders were fabricated from wire cables. In the case of Wheeling, wrought-iron rods were attached to a spreader beam resting on top of the cables. Cast iron was utilized in rollers at the tower tops while lap-forged “eye bars” embedded in the anchorage concrete engaged the wire cables. he French favored the Arc du Triomphe coniguration for bridge towers. Ellet on the other hand used a simple and yet monumental pair of stone towers. Inside each abutment is a wedge system for adjusting the shape of each cable. he wood deck was without any stifening truss. he 24-t-wide deck was supported on sizable timber loor beams which in turn were held at each end by the iron suspenders attached to the cables. he bridge opened with great acclaim in October 1849. he work had encountered few engineering challenges in its construction. Ellet, with justiication, could take pride in his greatest accomplishment. In 1854, a violent tornado swept up the Ohio River, striking broadside the light, unstifened deck. 36

Charles Ellet Jr .

he deck writhed in agony, with the surface of the deck acting like a kite. he upstream cables were thrown into the river while the downstream cables were unseated and came to rest over the tower arch. he cables were hoisted back into place, allowing the bridge to be reopened on a one-lane basis. his temporary work, which took a matter of a month, was under the direction of Ellet and his former assistant, McComas. Later, John A. Roebling’s son Washington undertook a major repair in 1860, where the most notable aspect was the grouping of the garland strands into a pair of larger cables. he rebuilding of the bridge has been attributed to John A. Roebling and not his son. he bridge remains in service with many original features. Over the years numerous repairs have been made, including a major restoration to celebrate the 150th anniversary of its opening. From the very beginning of ideas about the crossing of the Ohio River there were storm clouds on the horizon. A commercial rivalry ensued between Pittsburgh and Wheeling regarding the ascendancy of commerce and industry on the Upper Ohio River. he issue was not essentially of an engineering nature, but a political contest between the two cities. It was a contest for control of transport links to the west before the ascendancy of railways. he case centered on whether or not the bridge, with a Figure 2.6. Ellet’s drawing of the Wheeling Suspension Bridge. (Emory Kemp, Wheeling Suspension Bridge, A Pictorial Heritage [Institute for the History of Technology and Industrial Archeology, 1999])

37

Chapter 2

90-t clearance above low water, constituted a hazard to navigation. Ater the opening of the bridge in 1849, attempts were made by river steamers to burn the bridge deck by erecting elevated smokestacks. he arson attempts, however, were unsuccessful. At the same time, for three years the courts, including the Supreme Court, became involved. In a profound sense, the central issue became a test case for the right to cross navigable streams by roadway or railway bridges. A decision was made in 1852 declaring the bridge part of a national post road. As a result, Ellet obtained a decisive victory and the bridge was saved for posterity.

Figure 2.7. A view of the completed Wheeling Bridge showing the twelve wire cables in a garland arrangement. (Eli Bowen, Rambles in the Path of the Steam Horse [Philadelphia: Bromwell and Smith, 1855])

38

Charles Ellet Jr .

Control of River Navigation While Ellet was supervising the construction of the Wheeling Suspension Bridge, he became interested in what would later be called river hydraulics. His special interest focused upon both lood control and low-water augmentation of navigation channels. As an indication of his fascination with the subject, he named his son Charles Rivers Ellet. he continuing controversy with Pittsburgh took on added intensity with the inal arrival of the Baltimore & Ohio Railroad in Wheeling. Ellet argued for a route that followed the bank of the Ohio River from Fish Creek to Wheeling. he B & O entered Wheeling along this route on Christmas Eve, 1852. he great engineering enterprise to connect Baltimore with the western waters was completed. At that time, Wheeling apparently had a strong possibility of becoming the queen city of the Upper Ohio River. With his heralded success Ellet turned his attention to his latest interest, which he believed had national signiicance and would greatly enhance his reputation as an engineer. In typical Ellet fashion he entered the ield by publishing his engineering vision in a 64-page pamphlet published by the Smithsonian Institution. It was in this pamphlet that he emphasized the eicacy of the use of reservoirs constructed on Ohio River tributaries, including the Monongahela and Allegheny rivers. On the basis of his reputation and this Smithsonian report, he petitioned Congress for support to build selected reservoirs. he Ellet memorial gained support in the Senate, which established a select committee. He enjoyed a favorable response and the committee recommended a survey of reservoir sites. A reduced bill passed the Senate, allocating the amount of $20,000 to accomplish a detailed survey. he bill did not pass the House of Representatives but Congress authorized a survey for lood prevention on the lower Mississippi. As one might expect, Ellet applied to be put in charge of the survey. Although approved for the position, Ellet would not accept it because it required him to join with three engineers. he situation was resolved when Ellet was permitted to act as an independent consultant apart from the Corps of Engineers assigned to the project. A very thorough report resulted from the Corps of Engineers team. his work, entitled Report on the Physics and Hydraulics of the Mississippi by 39

Chapter 2

Humphrey and Abbot, was published in 1861 and was heralded at home and abroad as a brilliant piece of work. It is not surprising when Humphrey later assumed the directorship of the U.S. Topographical Engineers that this study remained a standard reference in Army engineering circles for decades. With the Corps of Engineers responsible for navigation only and not lood control or low-water augmentation in navigation channels, the main practical thrust of this work can be summarized as the levees-only approach to river control. As we have seen, severe looding on the lower Mississippi and the resulting failure of levees in the New Orleans region proved that this approach was inadequate. As a policy it stubbornly remained in efect until replaced in 1937. Ellet’s cogent report stressed the concept of reservoirs on tributaries to control looding, and low-water augmentation in the summer months. He did not discard the idea of levees but thought they could supplement his reservoir scheme. With the construction of more than a dozen control dams in Pennsylvania, Maryland, and West Virginia, Ellet’s vision has been vindicated. Figure 2.8. A view of the island side of the Wheeling Suspension Bridge. (William E. Barrett, Historic American Engineering Record, Washington, D.C.)

40

Charles Ellet Jr .

War on the Mississippi A staunch supporter of the Union efort in the Civil War, Ellet had ideas on how to win this conlict. He wrote to many about his ideas and must have driven President Lincoln et al. to despair. One of his ideas was to bear fruit as a result of his badgering both the Navy and the War Department about the use of ram boats on the rivers. While the Navy stood aloof, he received a commission as a colonel in the Army to build a leet of rams to battle with Confederate forces aloat. Ellet marshaled his forces, converting a number of river boats into rams by greatly strengthening their hulls with heavy timbers and itting ram bows to create a modern-day example of ancient rams used in the classical world. he ram leet sailed down the Mississippi to join in the battle of Memphis. he rams were instrumental in a Union victory. To the chagrin of the Navy, son Charles Rivers Ellet took the surrender of the city and hoisted the Union lag over the city building. Figure 2.9. he wood deck before repairs were made, 1956. (L. L. Jemison, West Virginia State Road Commission)

41

Chapter 2

Despite his stunning contribution to victory at the battle of Memphis, the Navy tended to discount his contribution. During the battle, the impetuous Ellet stood outside the pilot house on his ram in an exposed position and was shot in the leg. Most descriptions of this incident indicated that he was later recovering, but his wound became infected, resulting in his death on June 21, 1862. here are, however, several hints in the record that he may have contracted measles, which were a severe malady in adults.

An Eminent Career If Ellet had spent his career in England, he clearly would have rated an entry in Sir Samuel Smiles’s Lives of Engineers. His career had all the elements for a romantic Victorian novel. In a true Victorian melodrama, within a fortnight his beloved wife, Ellie, is said to have died of a “broken heart.” Perhaps the best evaluation of his character was written by Hiram Chittenden, a U.S. Army Engineer oicer, who wrote to Ellet concerning his behavior Figure 2.10. Colonel Ellet’s ram boat before the battle of Memphis. (Kemp collection)

42

Charles Ellet Jr .

in connection with his appointment as chief engineer on the James River and Kanawha Canal. Again, your energy, while it has not failed to command the respect and admiration of the board, and confessedly its you eminently for encountering the peculiar diiculties with which the progress of the work will be attended, has nevertheless a spice of impetuosity, together with an uncompromising tenacity to adopted conclusions and an impatience of diferent views, which the . . . are fain to believe will not characterize you ater a time. To be perfectly frank, there have been expressions of yours in your intercourse with the board of directors . . . liable in their apprehension . . . to be construed as dictatorial, if not sarcastic . . . 15

His contributions to engineering are truly monumental, including the introduction of long-span wire suspension bridges to America, numerous publications on suspension bridges, lood control on navigable streams with reservoirs, Figure 2.11. Charles Rivers Ellet at the time of the battle of Memphis. (Michigan University Library)

43

Chapter 2

and a little-known but signiicant role in the Civil War. He was clearly one of the most talented engineers in antebellum America. Chapter 2 Notes 1

2 3 4 5 6 7 8 9 10 11 12

13 14 15

Gene D. Lewis, Charles Ellet, Jr., he Engineer as Individualist, 1810–1862 (Urbana: University of Illinois Press, 1968), 1–30. his is the only full-length biography of Ellet. Lewis, Charles Ellet, Jr., 5. Samuel Smiles, Lives of the Engineers (London, 1862). Department of the Treasury, Report of the Secretary of the Treasury [Albert Gallatin] on the Subject of Public Roads and Canals [1808] (New York, 1808). Emory L. Kemp, he Great Kanawha Navigation (Pittsburgh: University of Pittsburgh Press, 2000). Ronald E. Shaw, Erie Water West: A History of the Erie Canal, 1792–1854 (Lexington: University of Kentucky Press, 1990). Claude Navier, Sur les Ponts Suspendus (Paris, 1823). Lewis, Charles Ellet, Jr., 27. Ibid. Ibid., 9–14. Ibid., 45. Charles Ellet, Jr. “Laws of Trade” in An Essay on the Laws of Trade: in Reference to the Works of Internal Improvement in the United States (Richmond: P. D. Bernard, 1839). Charles Ellet, Jr., A Popular Notice of Wire Suspension Bridges (pamphlet, 1839). Donald Sayenga, Ellet and Roebling (Easton, PA: Canal History and Technology Press, 2001). Lewis, Charles Ellet, Jr., 44.

44

3

A horoughfare hrough the Howling Wilderness: he Weston & Gauley Bridge Turnpike Emory L. Kemp and Janet K. Kemp

I

n 1849 work was begun on a turnpike road through a wild and largely unsettled area of western Virginia. Considering the very limited resources and lack of any local engineering experience, it was a bold undertaking indeed. What the proponents of this enterprise lacked in technical and inancial resources they made up for by a display of broad-based local support and great enthusiasm. here was nothing unique in this approach to building a national transportation system since it had been used earlier for roads, canals and railways in both Britain and America. In an age of progress turnpikes in both countries were built in response to local pressures and not according to any national plan. Little wonder that many failed to provide the expected inancial returns, but they did provide arteries for industry and commercial development and in that sense they were successful. he Weston & Gauley Bridge Turnpike, stretching nearly 110 miles through a very rugged Appalachian landscape, provides an excellent example of how turnpike roads were conceived, designed, inanced, constructed, and operated. To appreciate fully the history of this turnpike and its inluence on the region through which it passed, its story must be presented in the larger context of the history of road construction and as part of a larger enterprise in Virginia.

45

Chapter 3

Early Roads and Turnpikes Much has been written about Roman engineering skills and accomplishments, particularly about the vast network of Roman roads that connected all parts of the Empire with Rome.l At irst glance there would appear to be little more than a tenuous linking between a little-known and rather insigniicant mountain turnpike in nineteenth-century Virginia and the Roman roads in Europe. Closer inspection reveals that Roman engineering had a signiicant inluence on road construction technology in the eighteenth and nineteenth centuries as relected in design, speciications, and construction practices. Roads in one form or other have existed since the dayspring of civilization, but it was the Romans who irst built permanent roads, using techniques developed on an empirical basis. hese roads served both the military and commercial needs of the Empire. hey were so well built that the appellation “permanent” is most itting. he need for surfaced all-weather roads was largely the result of the use of wheeled vehicles, the tires of which chewed up the surface of dirt and even lightly graveled roads at certain times of the year. he principles of construction were twofold, namely adequate drainage and a sound foundation. he Romans appreciated the need for good drainage since a waterlogged road sub-base of ine material such as clay or silt loses most of its bearing strength. Roman roads were generally built above the elevation of the natural ground on a low causeway and were provided with generous ditches. A solid foundation was achieved by a layer of compacted earth upon which was placed a course of small stones followed by an impervious course of lime or hydraulic-cement concrete using local aggregates. he camber was built up with a crowned course of gravel supporting a wearing course, which was typically composed of large hand-placed paving stones. he collapse of the Roman Empire saw the virtual disappearance of wheeled road vehicles until the sixteenth century in Europe. he Europe that arose from the ashes of the Empire developed along and depended heavily on rivers and later canals for the movement of goods and people. Land transportation was largely a matter of pack animals, riding horses, or, for the poorer classes, walking.

46

T h e We s t o n & G a u l e y B r i d g e T u r n p i k e

Figure 3.1. Cross-sectional diagrams of a) Roman road, b) Tresaguet’s system, c) Telford’s system, d) McAdam’s system. Drawing by Emory L. Kemp.

47

Chapter 3

he Renaissance saw a rapid expansion in trade as well as a lively interest in the classical world of Greece and Rome. his resulted in a great increase in the use of wheeled vehicles for carrying passengers as well as goods, which necessitated better roads. In Tudor times, Parliament passed an act in 1555 that endeavored to make more systematic and, at the same time, strengthen the traditional organization and responsibility for road maintenance by the parishes. he Virginia system of using “male titheables” for road work stems from this English tradition.2 his system persisted from Elizabethan times until nearly the end of the Hanoverian kings, but it never was really satisfactory. Even as late as the Napoleonic War, roads in both Britain and America were in an appalling state. Yet in the Renaissance period, interest in the Romans provided an insight into the superb road system the Romans must have had. Roman achievements in road building provided an inspiration and an example for better roads. By the end of the seventeenth century, national governments, particularly the French, were in a better position to inance and construct roads and canals on a national scale. France led in this movement and in 1716 the Corps de Ponts et Chaussées was founded to supervise public works.3 Dating from this time roads in France were laid out with considerably better alignment and better drainage. he work was under the supervision of road and bridge engineers of the Corps in Paris. Pierre Tresaguet developed a new system of road building that was adopted throughout France, c. 1775. His method was to lay two courses of large stones followed by a layer of small stones, all well beaten down (i.e., compacted) to leave no gaps between them. In this way he provided a smooth wearing surface which was also virtually waterproof if well maintained. With this system, iron wheelrims pulverized enough of the surface to provide a binder for the crushed stone, which was slightly cementitious and resulted in a waterproof surface for the road. Tresaguet also provided better drainage and, with his insistence on adequate maintenance, the result was a marked improvement in roads. he Tresaguet system is shown in Figure 3.1, where it can be seen that in a sense the French method was the reverse of the earlier Roman system, with the larger stones used to form a solid base on the bottom of the road and not a wearing surface on the top. he method inspired American practice directly and circuitously through Telford and McAdam, whose methods are described below. 48

T h e We s t o n & G a u l e y B r i d g e T u r n p i k e

In England, road and canal building was accomplished by private stock companies established by Parliament and authorized to construct a road or canal for public use between speciied geographical points. In the case of roads these became toll roads or more commonly turnpikes, so-named from the trafic control bar at each tollgate. he system of turnpike roads in England was well established in the eighteenth century. Turnpike trusts levied tolls, a part of which was supposed to be used to keep the roads in good repair. he rest of the tolls, if any, were used to pay dividends on the issued stock. In both France and England, except for turnpikes, roads were built with the use of statutory labor whereby citizens were required to work a ixed number of days a year, making or repairing roads. In the American colonies before the Revolution the same method was adopted. Each county appointed a supervisor to oversee the construction and maintenance of roads. hey were also authorized to command all males over sixteen years of age to work on the roads on appointed days. Occasionally money was available for road work through private donations or the proceeds of public lotteries. By the end of the eighteenth century there was a great increase in traic on the roads and a great need for more and better roads. he old system of keeping roads in repair by local county authorities was insuicient to meet the demand. he states and federal government did not have the capital to invest in a public system of roads, the one exception being the National Road. hus, chartered turnpike companies using mixed private and public capital seemed to ofer the best alternative. he irst turnpike in the United States was authorized in 1772 by Virginia.4 From this date, turnpike companies proliferated and ater 1800 most of the states authorized toll roads. hese turnpike roads were built by private contractors using hired labor. he American road builders looked to the developments of the French and British engineers, as leaders in road construction.

he Turnpikes of Virginia he early roads in Virginia, including what is now West Virginia, following the routes taken by the early settlers in the eastern part of the state were merely trails blazed by frontiersmen or Indians. Ater the Revolutionary War new roads and the repairing of old roads came under the government of the county courts 49

Chapter 3

who could apply to the General Assembly for help in building major works such as bridges. Road building proved too costly for the counties alone and the legislature authorized the construction of toll roads where it was hoped that the tolls would pay for the cost of the road and its later maintenance. In 1772, the Virginia legislature authorized the irst toll road between Jenny’s Gap and Warm Springs in Augusta County. One of the early and most successful toll roads, the Little River Turnpike was chartered in 1811. In the early days of the turnpikes the General Assembly provided some of the money and the rest was raised by lottery. Joint stock companies were a later development. he number of turnpikes constructed in Virginia between the end of the Revolutionary War and the General Turnpike Act of 1817 was very small. here was no comprehensive plan for building roads to complement the growth and development of the state. Attention was drawn to the need for such an overall plan by a famous report on roads and canals delivered to the United States Senate in April 1808 by Albert Gallatin.5 Gallatin, who was secretary of the Treasury, believed that it was in the best interests of the new nation to have a comprehensive transportation plan to facilitate trade and develop the country—“good roads and canals will shorten distances, facilitate commercial and personal intercourse and unite, by a still more intimate community of interests, the most remote quarters of the United states.” Gallatin proposed a plan to improve connections between the Atlantic seaports and to link these ports with the Great Lakes and the western waterways by reaching the Ohio River. He felt that so great a plan as this could be accomplished only by the federal government since there was simply not enough private capital to accomplish so ambitious an enterprise, and the population of the United States was spread so thinly. Good transportation was, he felt, necessary for the good of the nation: “No other single operation, within the power of Government, can more efectively tend to strengthen and perpetuate that Union which secures external independence, domestic peace and internal liberty.” he main emphasis in his report is upon canals, which were at the time the most advantageous means of transporting goods over any distance. Roads were secondary to the canals and were to be built as links where it was not feasible to build canals. his attitude had prevailed in Britain with regard to early railways and canals. he interest in canals was relected in Virginia’s early legislation and 50

T h e We s t o n & G a u l e y B r i d g e T u r n p i k e

continued support of canal companies at the expense of railroads, especially the James River and Kanawha Canal. By the time of Gallatin’s report the population of Virginia had increased considerably and a greater part of the state was permanently settled. he need for adequate transportation was quite apparent and in 1816 the Virginia General Assembly created the nation’s irst Board of Public Works and Fund for Internal Improvement. his provided for the establishment of joint stock companies using both public and private capital. he principal function of the Board of Public Works was to supervise the internal improvements of the state. he board assessed the merits of various turnpike proposals, examining the location, construction methods, costs, etc., and recommended to the legislature that certain private companies be chartered. he board also administered the state funding for these turnpike companies. he turnpike companies were required to report annually to the Board of Public Works on the progress of their work and the board in turn made an annual report to the General Assembly. he Board of Public Works, however, had no authority or power to plan or build roads; it was only to advise the legislature and to supervise such building as the legislature had authorized. his meant that there was no centralized plan for road building in the state and the roads that were built resulted from local initiative regardless of whether they satisied the transportation needs of the state as a whole. As a result, in the irst forty years of the nineteenth century most of the turnpikes in Virginia were built in the eastern part of the state. In the western part (now West Virginia) the population was too small to raise the necessary capital. Ater 1840 there was a marked increase in turnpike companies chartered in western Virginia. But it must be noted that at this time numerous railroad companies were chartered in eastern Virginia. he turnpike as a means of transportation was being eclipsed by the railroad and western Virginia was again neglected in this respect, a situation that was remarked upon at the time by the discontented western Virginians. he weaknesses of the Board of Public Works in providing for the comprehensive transportation needs of the state was clearly felt by its most famous chief engineer, Claudius Crozet, who was oten at odds with the General Assembly. Capt. Crozet had been a French artillery oicer under Napoleon. Ater the battle of Waterloo he had let Europe, in 1816, for the United States. He served as 51

Chapter 3

a professor of engineering at the U.S. Military Academy until 1823, when he became the principal engineer of Virginia. He served in this capacity from 1823 to 1831 and again from 1838 to 1843. he break in his service was a result of his disagreement with the legislature. He was an early advocate of railroads and proposed a railroad to link the eastern and western parts of the state. he inluential members of the legislature favored canals and ater a reorganization of the Board of Public Works, Crozet resigned. Later, as the state recognized the use of steam locomotives, Crozet was reinstated as chief engineer until 1843, when the oice was abolished. He continued to serve the state as a consultant. he job of chief engineer was not an easy one but he accomplished a great deal in improving transportation in Virginia. He was responsible for two major east-west routes to link the two parts of the state, the James River and Kanawha Canal, and the Northwestern Turnpike. Crozet and his engineers conducted surveys throughout the state to determine the feasibility of roads. he private turnpike companies chartered by the General Assembly were responsible for the actual building of the roads and Crozet was oten critical of these roads. His advice with regard to location, alignment, width, and construction was oten ignored by the companies and he was critical of the resulting roads. He was also very conscious of the need for accurate surveys and maps of the state to assist in the location of new roads. However, he had great diiculty in persuading the General Assembly to provide the means to achieve a satisfactory map. Although he did produce several maps for the state, his most notable one in 1848, he was dissatisied with the results and continued to urge the need for an adequate map. On the whole, the legislators failed to appreciate his concern.6 With the interest in turnpikes at the beginning of the nineteenth century and the establishment of the Board of Public Works, the Virginia General Assembly passed a general Turnpike Act in February 1817 to regulate the incorporation of turnpike companies. his act set forth the general regulations for turnpike companies. he irst part was concerned with raising the stock. Ater the subscription books had been opened and the public notiied, half of the capital had to be subscribed before the company could be declared incorporated. he subscribers then could elect a president and ive directors to transact the business of the company. he 52

T h e We s t o n & G a u l e y B r i d g e T u r n p i k e

president and directors were empowered to buy any land necessary for the road. If a landowner would not agree, the county courts would settle the matter and award the landowner damages. he regulations for the road itself were given as follows: the road had to be sixty feet wide, eighteen feet of which to be well graveled; a summer, or side road, eighteen feet wide, was to be kept in good repair. Every ive miles a toll gate could be erected. Maximum weights were given for wagons according to the width of their wheels; the wider the wheels, the larger the weight allowed. Scales were to be erected to check weights. Tolls were ixed as: a score of sheep or hogs, 6¼c.; score of cattle, 12½c.; every horse, mare or mule, 3c.; two-wheeled carriage, 10c.; cart or wagon with wheels less than four inches, 6¼c. for each animal drawing it; cart or wagon whose wheels were four inches, but less than seven inches, 3c. for each animal; cart or wagon with wheels more than seven inches, 1c. for each animal. Troops and public state property were exempted from tolls. he subscribers had to hold an annual meeting. he road had to be kept in good repair. If the directors failed to keep it in good repair the local magistrate could suspend tolls. Persons using the road were instructed to drive to the right hand. Road construction had to begin within two years of the date of incorporation and be completed within ten years. All subsequent acts incorporating turnpike companies were bound by the regulations of this act except for the provisions that were speciically stated. he interest in building turnpikes was high and many companies sought charters. he greatest diiculty these companies faced was raising suicient funds from private investors before actual work could be done and before the state would contribute its share. Twenty percent of the private portion had to be actually collected before the state would contribute. Out of 647 companies chartered by the state, fewer than 30 percent even became operating companies, and very few of these ever made enough money to operate successfully.

Turnpike Surveys and Construction Practices When the company had enough money it could begin to build the road. he actual construction of the road varied from company to company. he 1817 act simply speciied that the road should be cleared for sixty feet, of which 53

Chapter 3

at least eighteen feet should be covered with gravel. In the western part of the state, where road builders encountered numerous diiculties, the width of the road was oten reduced and the summer or side roads dispensed with altogether. Some turnpikes were no more than improved dirt roads while others were macadamized and constructed with elaborate drainage systems. here were several manuals of road-building practice available to road builders early in the nineteenth century.7 Most of the experts agreed that a good dirt road was adequate for all kinds of vehicles, but to keep it in good condition it had to have some sort of covering to keep it dry and to provide a smooth surface for vehicles. he gravel and stone that were put upon the road were not the road itself but simply a protective coating. he method by which the coating was applied and the materials of which it should consist were the subjects of great debate. Basically two methods were advocated, as proposed by McAdam and Telford, two leading British engineers. John Loudon McAdam published his method of road making in a book called System of Road Making in 1821. he surface of his roads was covered with small irregular stones. He recommended a hard stone, broken by hand into cubical pieces that would pass through a ring 2½ inches in diameter (see Figure 3.2). he stone was broken with a hammer six inches long and weighing about one pound, with a handle three feet long if standing, or eighteen inches if sitting. He recommended a coating of twelve inches of consolidated stone. he road bed had to be properly shaped and sloped each way from the center. On this bed three inches of clean broken stones were laid in dry weather, then traic allowed on the road to consolidate the stones, or a heavy roller could be used. hen a second coating of three inches was added during a wet period, and again compacted. A third and fourth coating were laid in the same way. A cross section of a typical McAdam road is shown in Figure 3.1. homas Telford constructed his road of broken stones upon a specially prepared bed. Upon a level bed he set by hand a course of stones to form a irm pavement. he stones in the middle of the road were to be seven inches deep, ive inches at nine feet from the center, four inches at twelve feet, three inches at iteen feet from the center. All the interstices were illed with stone chips. Upon this pavement four inches of hard stones, broken to it through a two-inch ring, were laid, compacted, and then another two inches laid, for the middle eighteen 54

T h e We s t o n & G a u l e y B r i d g e T u r n p i k e

feet. Broken stone or gravel was then laid on the sides to give a convexity of the road of six inches from the center to the sides. he whole road was then covered with a binding of l½ inches of good gravel. A cross section of a typical Telford road is shown in Figure 3.1. here were proponents of both kinds of broken stone roads, but the McAdam system seems to have prevailed in Virginia. McAdam’s method was clearly easier and cheaper to build. he actual construction practice no doubt difered widely from company to company and not all macadamized roads were built to the rigorous standards advocated by McAdam.

Figure 3.2. Drawing of ring and hammer for road construction. (W. M Gillespie, A Manual of the Principles and Practice of Road-Making [New York: A. S. Barnes & Co., 1848]198)

55

Chapter 3

he Weston & Gauley Bridge Turnpike With the interest of the nation turning toward expansion westward in the early years of the nineteeth century, there was considerable interest in improving and building roads from the east to the Ohio River. In the irst half of the century there were four main east-west roads that crossed Virginia and contributed greatly to the increasing settlement of western Virginia. he irst road was the federal National Road built from Cumberland to Wheeling between 1811 and 1820. It touched only a little of northern Virginia, but it had considerable inluence on the construction of the other roads and in many ways served as a model. In 1819 the Virginia General Assembly chartered the Kanawha Turnpike as an extension of the James River and Kanawha Canal. It was inished to Gauley Bridge in 1825, reached Charleston in 1827, and was later extended to the Ohio River. he Northwestern Turnpike was begun in 1831 to go from Winchester to Parkersburg by way of Romney, Graton, and Figure 3.3. Crozet’s map of Virginia, 1848. (Virginia State Library)

56

T h e We s t o n & G a u l e y B r i d g e T u r n p i k e

Clarksburg. he road did much to open up the Monongahela Valley to settlement. Finally, in 1838 another east-west road was planned from Staunton to Parkersburg. his road was begun at both ends and had been completed from Parkersburg to Weston and from Staunton to Beverly when the money ran out, leaving a forty-six-mile gap between Beverly and Weston. In 1845 this link was made. It had long been felt that there was a need for a north-south road linking the two main east-west roads, which would provide access to parts of Braxton, Lewis, and Nicholas counties, which were increasing in population. In 1827 the Board of Public Works authorized a survey for a road from Gauley Bridge to Nicholas Court House (Summersville), and then to Haymond’s Salt Works (Bulltown), and then to Lewis Courthouse (Weston) and on to Salem. It is not clear whether the survey was made or not, but a road was not built at that time. In 1837 the Virginia General Assembly passed an act authorizing a road to be constructed from the Nicholas Court House to Gauley Bridge. In 1838 another attempt was made to build a north-south road and an act authorized a road from Weston to Charleston. his road was built and served as a link from the Monongahela Valley to the Great Kanawha but did not follow the route of the later Weston to Gauley Bridge road, and there was still no adequate road in this area (Figure 3.3). By the 1840s the population in Lewis, Braxton, and Nicholas counties increased suiciently to make possible the construction of a north-south road. According to the 1850 U.S. Census report there was a total population in Lewis County of 10,031; in Braxton 4,212; and in Nicholas 3,963. he number of families was listed as Lewis, 1,533; Braxton, 679; Nicholas, 602. Lewis County was by far the most populated area and a sizeable part of the population lived in Weston, on the Staunton and Parkersburg Turnpike. he people were anxious to see a southern link with this road. Although they had a road to Charleston, the people of Weston thought that a road opening up the interior of Lewis and Braxton counties would increase trade in Weston and they were greatly in favor of the Weston & Gauley Bridge Turnpike. In 1845 there were about sixty houses in Weston, and several shops and businesses. It was a growing town and was by far the largest on the route of the Weston to Gauley Bridge turnpike. Although the population of Braxton County was smaller than that of Lewis, a great deal of the enthusiasm for the turnpike road came from the citizens of 57

Chapter 3

this county. he number of individual subscribers to the stock of the turnpike company was considerably larger in Braxton County than any other. Sutton, the county seat of Braxton County, was a small town but at an important point on the Elk River. here were numerous grist and saw mills located in or near Sutton that used the Elk River for power. In addition, from Sutton downstream the Elk was navigable to the Great Kanawha and thence to the Ohio River, and goods were transported by boat. A good road would make it much easier for farmers to transport their grain and timber to the mills. he area near Bulltown had been settled early because there was a ready supply of salt there, which was in demand by the early settlers. In 1768, at a salt lick about one-quarter mile below the present site of Bulltown, an Indian, Captain Bull, had come from the Susquehanna River with about ive families and settled there. he Indians collected the salt and traded peacefully with the early white settlers in the area. Many of these settlers were rough people who had no respect for the Indians and seem to have been ready to ight with them on the slightest excuse. In 1772 Capt. Bull and his Indians were all killed in the massacre at Bulltown.8 he collection of salt at Bulltown remained important. In 1809 Colonel John Haymond and his brother-in-law, Benjamin Wilson Jr., erected a furnace and set up evaporating kettles. At this time the Bulltown salt works were the main source of salt in the area. he Haymond Salt Works ceased production in 1823. Salt continued to be produced in the area by John P. Byrne and Addison McLaughlin at Bulltown and by Asa Squires at Salt Lick Bridge. hese three men were instrumental in the original formation of the Weston & Gauley Bridge Turnpike and no doubt saw the road as a means of expanding their trade. he production of salt in this area does not seem to have survived the Civil War. he main occupation of the region to be served by the Weston & Gauley Bridge Turnpike was farming. he area was still thinly populated and the amount of land under cultivation was still very small. In Lewis County, out of 174,979 acres of land only 48,152 acres were improved; in Braxton, 16,111 out of 920,443; and in Nicholas, 19,335 out of 151,684. At the same time, 1850, in many eastern Virginia counties more than half the land was improved. However, there were a number of individual farms (878 in Lewis, 408 in Braxton, 418 in Nicholas) and these people welcomed improved roads as a means of transporting their produce and livestock to markets. 58

T h e We s t o n & G a u l e y B r i d g e T u r n p i k e

Although many turnpike companies never materialized and never built a road, the proposers of the Weston & Gauley Bridge Turnpike felt not only that there was need for such a road but that there was suicient local support. he directors of the company saw the road as “the connecting link of like improvements stretching from Pennsylvania to southwestern Virginia and rendering accessible the most productive and interesting portions of the commonwealth; which but recently were a howling wilderness.”9 In spite of the improvements the area was still unsettled and isolated. he directors chose to depict on the seal of the company a scene which strongly reinforces this impression. he seal is described in the irst annual report of the company: To perpetuate in history a most tragical scene which took place on the line of the road in its early settlement, in which William Given, who is yet living, and his three infant children were attacked by a ferocious bear three times, and who inally repelled and drove away the animal and saved the lives of himself and the children, we have had the seal of the company prepared with a representation of that scene.10 Figure 3.4. Seal of the Weston & Gauley Bridge Turnpike Company. Drawing by Emory L. Kemp.

59

Chapter 3

he optimism of the directors seems to have been shared also by a great number of the people in the area, and the original capital was oversubscribed by $675.00. It is quite evident that the road had the backing of the local inhabitants since there were no claims for damage by the landowners through whose property the road went, “showing the lively interest taken by the people in the road for its prosperity and permanency.”11 he problem of damages had oten plagued Virginia turnpike companies but in the case of the Weston & Gauley Bridge Turnpike a great deal of money was saved through the cooperation of the landowners. A letter written in 1850 by one of the directors of the company, John Robinson of Summersville, sums up well the hopes and optimism the residents had for the road: It is the only improvement that can ever make the counties through which it passes and will when completed be of great advantage to the state by bringing into use the hitherto hidden resources of the country. We inhabit a country that is rich in minerals such as coal, iron oar [sic] and with water power to drive all kinds of machinery, with the inest stock country if improve that I know of anywhere. If we improve those gits nature has bestowed upon us, our section of county must prosper; our lands must enhance in value and thereby give a greater revenue to the state.12

Formation of the Weston & Gauley Bridge Turnpike Company On March 25, 1848, the General Assembly of the state of Virginia passed an act incorporating the Weston & Gauley Bridge Turnpike Company with a capital stock of $30,000, of which the Board of Public Works was authorized to subscribe $18,000, or three-iths of the capital stock. With the passage of this act the organizers and promoters of the road were authorized to advertise the stock and seek stockholders. he stock was sold in shares of $25.00. Many turnpike companies in Virginia at this time were unable to raise suicient stock ater they had been incorporated and their turnpikes never became a reality. he Weston & Gauley Bridge Turnpike, however, was very successful in raising the required $12,000 of its share of the stock. A irst meeting of the stockholders 60

T h e We s t o n & G a u l e y B r i d g e T u r n p i k e

was held on October 18, 1848, at the Braxton County courthouse in Sutton, to organize the company. Asa Squire was appointed president pro tempore and John P. Byrne clerk pro tempore. hree men were appointed, one from each county through which the road was to run, to report on the amount and legality of the stock subscriptions. his committee was composed of James G. Neil of Nicholas County, Addison McLaughlin of Lewis, and P.M. Adams of Braxton. he act of incorporation had authorized the county courts to buy stock in the company in any amount they thought proper; the county courts subscribed to the company in large amounts, the balance being taken by private subscriptions. he committee reported that Braxton County had subscribed $3,000, Lewis County $1,500, and Nicholas County $6,000. Private subscriptions from Braxton totaled $1,550, and from Lewis County $625. here were no private subscribers from Nicholas. he total amount subscribed was $12,675.00. Only $12,000 was needed to be suicient for the company to proceed with the road, therefore the excess of $675.00 was returned, in the amount of $25 each, to each subscriber who had bought more than one share. It is interesting to note that the private subscriptions were in small amounts, the largest being four shares. For the $675 there were forty subscribers of whom thirteen bought one share only. he relatively large numbers of shareholders perhaps indicates the great interest and hopes shown in this road by the citizens of these isolated counties and the small amounts may well indicate the relative lack of capital in the western counties. At this irst meeting of the stockholders, the board of directors of the company was appointed. Felix Sutton was appointed president, and he remained its president until the road was completed. Sutton was the nephew of the early settler on the river Elk for whom the town of Sutton was named. He was born in 1802 and was brought to Sutton by his uncle in 1810 ater the death of his parents. When Braxton County was formed from Lewis County in 1836, Sutton was designated the county seat. It was a very small town at the time, with very little industry. Felix Sutton was one of the leading citizens of the town and active in the promotion of the road. He helped with the formation of the new state of West Virginia and represented Braxton County in the irst and second sessions of the new state legislature. He died in 1884. 61

Chapter 3

John P. Byrne was appointed the clerk and treasurer of the company. He was also at this time the clerk of the County Court of Braxton County and held this oice with the Turnpike Company until his death in February 1860, when Daniel S. Squire was appointed clerk and treasurer in April 30, 1860. According to the General Turnpike Act of the General Assembly of Virginia passed in 1817, which controlled the formation of all turnpike companies, a turnpike company should have a president, treasurer, clerk, and ive directors. he ive directors appointed at this irst meeting were James G. Neil, John H. Robinson, and John Brown of Nicholas County, John S. Camden of Braxton County, and Jonathan M. Bennett of Lewis County. hey were appointed for a one-year term. While the president and treasurer remained in oice during the whole time the road was being built, the directors changed frequently. In the summer of 1849 the Board of Public Works received several letters from concerned stockholders suggesting that since the board owned three-iths of the shares of the company the board should appoint three directors to represent its interests, and recommending suitable people.13 he board replied that since directors had already been appointed for one-year terms they could not lawfully be replaced. John Brannon, the state proxy, suggested that when their terms expired in October 1849 then the state should appoint three directors.14 On September 20, 1849, the state did in fact appoint directors—Jonathan M. Bennett, Addison McLaughlin, and John Robinson. For the next ten years the state continued to appoint diferent directors.15 In June 1849 several stockholders and interested persons wrote to the Board of Public Works recommending that a state proxy be speedily appointed to look ater the state’s three-iths subscription in the Turnpike Company.16 hey recommended John Brannon of Weston in Lewis County as a man “of sound discretion and judgment and of some experience in matters of this kind.”17 Brannon had been the superintendent of a section of the Staunton and Parkersburg Turnpike, which passed through Weston in the 1840s. In July 1849 he was appointed the state proxy, to represent the stock held by the Board of Public Works at meetings of the stockholders. He remained the state proxy until July 1851, when he resigned because “the meetings of the stockholders of that company are held, generally, so removed from my place of residence that it renders my attendance very inconvenient and indeed my engagements on 62

T h e We s t o n & G a u l e y B r i d g e T u r n p i k e

the Huttonsville Road might preclude altogether my attendance at important times.”18 He recommended that the board appoint Johnson A. Camden in his place, and Camden was appointed in September 1851. his board of directors, consisting of the president, clerk, treasurer, and ive directors, was responsible for collecting the stock payments, for letting the contracts on the road, and for overseeing the construction of the road and paying the contractors. here appears to have been some criticism of the company oicials from time to time, ranging from incompetence to mismanagement. he directors, on the whole, had little experience of road building and no doubt made mistakes. In an early letter to the secretary of the Board of Public Works, October 26, 1848, Felix Sutton had several queries as to how to proceed, “being Unacquainted with the general manner of proceeding in such business.”19 It seems to have been inexperience and vested interests that led to bad decisions on the part of the company rather than deliberate wrongdoing. John Robinson wrote in 1850 to the Board of Public Works: “As director I wish to do everything in my power to the interest of the company. It is a new thing to manage the afairs of a company to probably most of its directors at least it is to myself.”20 he directors and stockholders of the company had their own reasons for wanting the road, some desiring it to pass through their lands, others seeing it as a means of increasing their business ventures. At times, private interests may indeed have prevailed over public good. here are many instances of this. At the irst meeting of the stockholders Addison McLaughlin, later a director of the company, proposed a resolution “that the engineer be instructed in locating the Weston & Gauley Bridge Road, to cross the Little Kanawha River at the Bulltown Salt works.”21 he resolution passed and everyone knew that McLaughlin owned the salt works. his does not necessarily imply that this particular location was not in the best interests of the whole road. It was something to consider. J. Bennett, writing to the Board of Public Works in July 1849, regarding the appointment of a state proxy, remarked: I earnestly recommend the appointment of a State proxy, for this the [sic] reason that some interested land holders are endeavoring to inluence the location of the road for their individual beneit and at a meeting shortly to be held will probably determine the matter and may inlict upon the State and company an expense, unwise and unnecessary of $10,000.22

63

Chapter 3

John Brannon, the state proxy, writing to the Board of Public Works in August 1849 shortly ater he assumed that position, reported I regret to say the location has not been made in the most satisfactory manner, the inevitable result of the labors of the men who were employed to perform them on some portions of the line, by the stockholders at its annual meeting in October last and that matters have progressed so far, in many particulars, as to put it beyond the control of those representing the interest of the State and the interest of those who would be controlled by a regard for the public good.23

his suggests that the location of the road was inluenced by factors other than engineering considerations. here are other accusations of conlict of interest upon the part of the directors from time to time. John Callaghan, writing to the Board of Public Works in December 1849 to complain because he had been refused as a contractor and felt that he had been turned down because the directors were inclined to favor their own friends and relatives, says: “In short the whole business of the Weston & Gauley Bridge turnpike road seems to have gotten into the hands of a party of relatives, friends and interested persons who seem determined to ingross the whole business.”24 Since Callaghan felt himself to be the injured party his complaints must be considered in this context. But similar complaints occurred again. In 1851 B.W. Byrne had replaced Addison McLaughlin as one of the state directors of the company. Soon aterward his appointment was rescinded and McLaughlin re-appointed, for reasons that remain unclear. he reason given to Byrne was that he had moved his residence from Braxton to Upshur County, a reason that he did not ind compelling. He found much to complain about in the dealings of the company and could not understand why McLaughlin’s friends had wanted him reinstated since “he was wholly unit for the oice, from the fact that he had been drunk most, if not all the time for the last six months.”25 He suggested that McLaughlin had been re-appointed not because “there is no one else on the line that has qualiications but I presume it is for the purpose of carrying out some sinister motives.”26 What these motives were he does not state, but he implied that the afairs of the company were not wholly above board. here does indeed seem to have 64

T h e We s t o n & G a u l e y B r i d g e T u r n p i k e

been some conniving upon the part of the directors for their own interests. W.E. Arnold, writing to the Board of Public Works in September 1853 to accept his appointment as a state director, felt it necessary to report an incident to the Board of Public Works that he clearly felt to be improper. At a meeting of the company board “consisting of Messr. Camden, Brown, Cottle and President Felix Sutton, they appointed Messr. Camden and Cottle superintendents of the road, the 1st at a salary of $350 and the other at $200 and Mr. Brown engineer of the road, who has since taken a contract to make 10 or 12 miles of said road.”27 here is also some evidence of mismanagement of the funds, although from the records it is diicult to obtain a very clear picture. Jonathan Bennett, who had been a director of the company from the beginning until his resignation in 1852, became concerned about the company in 1855, when he wrote in June to the Board of Public Works, “I have no hesitation in saying there is mismanagement in the afairs of this Company.”28 He requested a copy of the account that the company had sent to the Board of Public Works. Ater he received his copy, he again wrote to the board in July 1855 that “he very state of things I supposed existed.” He claimed that “the oicers of the Company are speculating upon the money due to contractors” and that the contractors were complaining because they were unpaid. He lays the blame upon the board of directors except for two who were working to correct the abuses. Of the president, Felix Sutton, he says he “is a very honest man but very easily imposed upon.”29 Wherever the blame lay, the directors were negligent in their duties. William Arnold, writing in September 1860 to the Board of Public Works, complained: “he Weston and Gauley Turnpike is in worse condition now than ever before. Albert Lewis is wholly unit for director. In the irst place he knows nothing about roads and secondly he gives it no personal attention. he tolls would keep it at excellent condition properly administered. he public are suffering from his neglect.”30 he evidence of these letters is somewhat diicult to evaluate since they usually present only one person’s point of view, which may well be biased. here is rarely more than one letter upon any controversy. In the case of Addison McLaughlin, W. Byrne calls him totally unit for the job; Jonathan Bennett says of him “there is none who has more at heart the prosperity of this road than he—perhaps there is none better qualiied.”31 Not 65

Chapter 3

all the directors were incompetent or self-serving, as we see in a letter from John H. Robinson, a director in the beginning, written to the Board of Public Works in November 1849. He writes, having recently become a contractor for some eleven miles of the Weston & Gauley Bridge Turnpike, I feel it a duty owed to myself to resign my oice as State director. It might be thought my being road contractor and State director would conlict with the interest of the Company and this being an improvement that I have taken and have a great interest in I would wish to do nothing to mar its progress in the least.32

He, it seems, felt a scruple about a conlict of interests that others apparently did not feel. Nor, it seems, did the Board of Public Works, and Robinson remained a state director. It must, of course, be kept in mind that the men of the mid-nineteenth century were not as conscious of conlict-of-interest situations as their counterparts today. It would also have been diicult in such sparsely populated counties to have found men any more competent than the ones who served as directors and there would have been very few with any experience of road building. While it may appear that the members of the board of directors were not experienced in organizing and administering road making, they quickly made preparations for the building of the turnpike road. he general turnpike law stipulated that a road be cleared sixty feet wide, and at least eighteen feet covered with gravel, and a summer road cleared eighteen feet wide, free of stumps, rocks, holes, etc. In many of the acts incorporating speciic turnpike companies these regulations were relaxed, particularly in the western counties where smaller roads, built to less strict standards, were more feasible. In the act of March 1848 incorporating the Weston & Gauley Bridge Turnpike, it is stated that the company was subject to the provisions of the general act, except “that the said company shall not be required to pave or cover the road with stone or gravel, nor to make a summer or side road thereto; that the said road shall be cleared at least thirty feet wide and improved for a width of iteen feet and at a grade not exceeding ive degrees.”33 At the irst meeting of the stockholders in October 1848, William P. Haymond and Minter Bailey were appointed engineers to locate the turnpike from 66

T h e We s t o n & G a u l e y B r i d g e T u r n p i k e

Weston to the Nicholas County line crossing the Little Kanawha River at or near the Bulltown salt works. John Brown and James G. Neil were appointed engineers to locate the road from the Braxton and Nicholas County line to some point at or near the Falls of the Great Kanawha; in locating the road they were instructed to follow as closely as practicable the route of the county road. It was expected that the engineers would have located the road by the following year. In December 1848 Minter Bailey wrote to the Board of Public Works to “ask the Board for a heodolite.”34 hey could apparently not acquire one locally. It would be nice to know if they did receive one or not, but they apparently managed to locate the road. James Neil also wrote to the board in August 1849 to enquire what price was paid “as an average in the western part of the State to the engineers for locating roads by the day or by the mile.”35 his may indicate that they had already inished the location and were anxious to be paid. Although we have no record of the prices paid on the Weston & Gauley Bridge Turnpike, some records for the Slavin’s Cabin and Summersville Road are extant and may give a general idea of the rates of pay. In May 1853, the superintendent, James Bennett, reported that an assistant engineer received $2 a day; one man to cut brush, $15 per month; one staf or rod bearer, $15 per month; two chain bearers, $15 per month each; one man to drive stakes, $15 per month; one guide, assistant packer and camp keeper, $15 per month; one packer and horse, $1 per day; and boarding for the whole party, $2 per day.36 According to the irst annual report of the company in October 1849, the location of the road was well under way but not yet completed. William Haymond and Minter Bailey had located in their section, the road between Weston and Sutton, a distance of forty-three miles, but the remaining part south of Sutton to the Nicholas County line was not yet located. he other two engineers, John Brown and James Neil, had located twenty-ive miles of the road in Nicholas County but some part was still to be surveyed. he engineers estimated that the whole road between Weston & Gauley Bridge would be 106 miles, 68 of which had been located, 38 still to be located. hese men who had located the line of the road were local citizens and apparently had no particular qualiications for the job other than an interest in the road and a familiarity with the country. hey did not locate the road to everyone’s satisfaction. James Bennett, a Weston resident and stockholder, who 67

Chapter 3

was later engineer and superintendent of the Slavin’s Cabin road, writing to the Board of Public Works in January 1849, says, it is a matter of regret that competent experienced engineers, appointed by yourselves, were not employed in the location of these roads [he is also talking about the Weston and Fairmont Road]. Yet this misstep may be remedied in some degree, by the appointment of an experienced and scientiic superintendent, and at a salary suicient to engage the services of one such for each road would hardly be aforded, I wish to suggest the propriety of giving the making of both roads to one superintendent.37

He suggested Dr. James McCally of Clarksburg, who had superintended parts of the Northwestern Turnpike. His suggestion appears to have been ignored and at a meeting held on July 2, 1849, the directors of the Weston & Gauley Bridge Turnpike appointed L.D. Camden and H.G. Campbell as superintendents of the road and ixed the annual salary of the former at $350, and of the latter at $150.38 Mr. Camden was assigned to the division of the road north of Nicholas County and Mr. Campbell to the part south of the Nicholas–Braxton line. With most of the road located and the superintendents appointed, the board began to let contracts for the actual work of building the road. hey advertised for contractors with handbills similar to the ones in Figure 3.5. he actual speciications for the road no longer exist but they were no doubt similar to those written for the Slavin’s Cabin and Summersville Road in 1855 (see below, “Speciications for the Salven’s cabin and Summerville Turnpike Road”). he contracts were let usually in sections of ive miles. he width for the Weston & Gauley Bridge Turnpike was to be iteen feet and cleared thirty feet. At each stage in the construction of the road the superintendent had to examine the contractors’ work before he embarked on the next stage. When the contractor had completed his section and it had been accepted by the superintendent he was obliged to keep it in good repair for one year. Before a contract was signed the contractor had to obtain a bond equal to the value of the contract. he Weston & Gauley Bridge Turnpike was not macadamized over its entire length. A typical section of such a road is shown in Figure 3.6, 68

T h e We s t o n & G a u l e y B r i d g e T u r n p i k e

where the surface is being smoothed with a mule-drawn drag. Small sections around the principal towns were macadamized and separate contracts were let for those sections. In the summer of 1849 the irst contracts were let. he entire road in Lewis County from Weston to the Braxton County line, a distance of nineteen miles, was let at the average price of $310 per mile, and twenty-four miles in Nicholas County were let at $405 per mile. From these early estimates of the cost per mile, the directors reckoned the cost of the road “including bridges and other incidental expenses, at $400 per mile or $42,000 for the completions of the line of 106 miles of road.”39 Since the company had been incorporated with a capital stock of $30,000, it was obvious that this would be insuicient to complete the road. Consequently the company obtained permission to raise the capital stock to $45,000 by an act of the Virginia General Assembly on March 4, 1850. For this additional amount, Braxton County subscribed $2,000, Lewis $500, Nicholas $2,000, and individuals $505. On February 25, 1853, yet another act authorized an additional $15,000. In the second annual report, in October 1850, the president reported that work was progressing but “not as rapidly as desired.”40 Twelve miles were now completed but no more had been put under contract. In 1851 contracts were let for another thirty-two miles so that almost the whole road was under contract. In October 1852 the company reported that sixty-seven miles of road were completed and under toll. he part in Lewis County was all completed. he uninished sections were in Braxton County south of Sutton, and in Nicholas County west of Summersville. his part of the road in Nicholas County seems to have been somewhat neglected and the people of the county appear to have complained loudly. John Brannon, the state proxy, writing to the Board of Public Works in January 1851, states that the residents of Nicholas County wished to apply their part of the subscription, together with the proportional amount from the state, to the construction of the road from the Nicholas County Court House to Gauley Bridge irst, and that they cared very little for the part of the road from the Nicholas Courthouse north to the Braxton County line. he board resisted this attempt of the Nicholas County residents, which would have been detrimental to the whole road.41 69

Chapter 3

By the end of December 1853, seventy-seven miles had been completed at an average cost of $394.49 per mile. he road was inally completed by 1858, when it was reported to the Board of Public Works that it was 109¾ miles in length and all constructed. It was all under toll except for eleven miles at the south end, which had not yet been received (i.e., accepted) by the road commissioners. he inal accounting showed that the average cost per mile was $446.59. From 1853 the annual reports list the amounts of tolls collected and the cost of repairs annually.42 Figure 3.5. Advertisement for contractors on the Slavin’s Cabin and Summerville Road. (Virginia State Library)

70

T h e We s t o n & G a u l e y B r i d g e T u r n p i k e

Bridges on the Turnpike In the earlier years the construction of the road progressed fairly rapidly, presumably because they began on the easiest sections where a county road already passed. In the later years construction slowed as they encountered greater obstacles. Small bridges had to be built like the one shown in Figure 3.7, and described below in “Speciications for the Salven’s cabin and Summerville Turnpike Road,” under Bridges. he sections of the road that involved larger bridges were the last to be completed. On February 25, 1853, an act of the Virginia General Assembly authorized $30,000 for building three bridges, one over the West Fork River at Weston, called the Bendale Bridge, one over the Little Kanawha at Bulltown, and the other over the Elk at Sutton. Part of this money was also to be used for graveling and stoning the road. Under the provisions of this act, the board of the turnpike company contracted for the three bridges and also let contracts for two miles of macadamizing at Weston to Perry Lorentz for $2,205.00, and one mile through and north of Sutton to S. and A. Anawalt for $1,360.05, and one mile through Summersville to John Bell for $1,174.00. In June 1853 a contract for a Figure 3.6. Repairing a dirt road. (Virginia Highway Research Council)

71

Chapter 3

bridge across the west Fork River was let to Henderson H. Beall for $3,000.000. his bridge was built without any complications and was completed late in 1854 (see Figure 3.8). he contracts for the other two bridges were both let in June 1853 to Ira Hart; the one across the Little Kanawha, a covered bridge, for $3,400.00; the one across the Elk River for $8,448.00. hese two bridges caused more trouble. In October 1854 the president reported that the bridge across the Little Kanawha “is not in a rapid state of progress, except the masonry, which is more than half done.”43 Eventually this covered bridge was inished and was still standing as late as 1941. he Elk River Bridge caused considerable trouble. A larger bridge was needed to cross the river at Sutton and this was the largest bridge built on the Weston & Gauley Bridge Turnpike. A contract was let on June 7, 1853, in Sutton to Ira Hart “for building a bridge over the Elk River on the Weston and Gauley Turnpike at the point selected by the turnpike” for his bid of $8,448.00.44 He was to construct a bridge on the “Tresle suspension plan, similar to that on the B & O Railroad over Cheat River.” his means that they were proposing a Fink truss bridge. For some reason, which is not now clear, it was decided to alter the plans for this bridge and on July 23, 1853, this contract, with the consent of Hart, was set aside. A new contract was made with Hart “to build a wire suspension bridge over the Elk River on said road to be at least three hundred feet long and warranted to sustain ity tons equally distributed for which the company are to pay the said Hart the sum of $9,500.”45 Hart was to enter into a bond of $10,000. At this point Hart sublet the stonework and went himself to Wheeling to purchase the necessary wire. He bought the wire from Bodley and Company, who made the wire and the ixtures for the Wheeling Suspension Bridge built in 1849, and employed Mr. Downing, who had laid the wires for the Wheeling, Nashville, Charleston, Fairmont, and other suspension bridges.46 On October 3, Hart took his bond to the board. It was refused and his contract set aside, although his bond was accepted for the Little Kanawha Bridge. he board claimed that they did not know the signers of his bond. hen on October 4, the board signed a contract with Benjamin W. Byrne “to construct a wire suspension bridge over the Elk River on said companies road agreeably to the speciications iled by J. S. Camden, a superintendent of said road for which the Company is to pay to said Byrne $12,000.”47 his bridge was to be longer, being 72

T h e We s t o n & G a u l e y B r i d g e T u r n p i k e

460 feet from center to center of the towers, thirty-three feet high, and seventeen feet wide. Ira Hart was understandably annoyed and wrote to the Board of Public Works on November 1 to complain that he had begun work on his contract in good faith and could not understand why his bond had not been accepted. he Board of Public Works acted on his part and on November 9 instructed the directors of the Weston & Gauley Bridge Turnpike Company to rescind the contract with Benjamin Byrne and restore to Ira Hart his contract of July 23. he Board of Directors of the Weston & Gauley Bridge Turnpike met on December 6, 1853, to carry out the Board of Public Works’ wishes but they felt that the July contract with Hart was insuicient. Since Byrne had already built an abutment on the north side of the river, they decided, with the agreement of Hart and Byrne, that Hart should build a bridge to the same speciications as the one agreed to by Byrne, using the abutment already built. Hart was to be paid $11,500, and Byrne was to be paid $1,600 for work already done. he bridge was to be constructed by December 25, 1854. his was not the end of problems with this bridge. Presumably construction went ahead but in December 17, 1855, Felix Sutton, writing to the Board Figure 3.7. A typical small turnpike bridge. (Virginia Highway Research Council)

73

Chapter 3

of Public Works, reported that it would probably cost $12,000 or more and asked whether it should be a toll bridge. he bridge was apparently not inished at this date but must have been inished before 1857 when tolls were irst reported. A photo of this bridge is shown in Figure 3.9. He also informed the Board of Public Works that the bridge was being constructed twelve feet wide, whereas the contract speciied seventeen feet, and he felt that there should be a corresponding reduction in the price paid to the contractor.48 he Board of Public Works replied on January 12, 1856, that a toll should be charged on the bridge and that since the width had been changed, less should be paid.49 hey also wanted to know the reasons for the change in width. What happened next is unclear but it appears that the directors of the Weston & Gauley Bridge Turnpike Company refused to pay Ira Hart the full amount for the bridge because of the reduction in the width. Hart then brought a suit against the company in the Lewis Circuit Court on April 6, 1857. In the fall of 1858 Hart obtained a verdict against the company for $1,000, but the court set aside the verdict and the case was to continue. A compromise was proposed. Felix Sutton wrote to the Board of Public Works for their opinion “whether the Company should Figure 3.8. Bendale Bridge over the West Fork River at Weston. (Used by permission from Myrtle Auvil)

74

T h e We s t o n & G a u l e y B r i d g e T u r n p i k e

compromise the case or whether the suit should take its due course in law.”50 He wrote again to the Board of Public Works, in February 1860, that the “suit of Ira Hart against this Company is compromised; it will cost the Company about $12,000.”51 Apparently Hart received some additional payment but whether he was paid in full or not is unclear, nor is it known why the width of the bridge was reduced. here may have been another bridge built upon this turnpike that does not appear in the company records. It was a covered bridge over Salt Lick Fork near Flatwoods, twelve miles north of Sutton. John D. Sutton, in his History of Braxton County, quotes a letter from Harrison Kelley, who says “I was employed by Mr. Chenoweth for fourteen years in the building of bridges on the Staunton and Parkersburg Turnpike. I built the Jane Lew bridge and Figure 3.9. Braxton County, West Virginia. (Hardesty’s Historical and Geographical Encyclopedia [Chicago and Toledo: H. H. Hardesty, 1883])

75

Chapter 3

the Salt Lick Bridge over the Salt Lick Fork of the Little Kanawha, in Braxton County, myself.”52 he road was inally inished in 1858 but does not seem to have lived up to the expectations of its promoters. In 1850 one of the directors of the company, John Robinson, had written to the Board of Public Works in glowing terms of the resources of these western counties and their value to the state, saying that “this road in a few years must be proitable to the state.”53 he road may have been valuable to the residents of these remote Virginia counties but it was never proitable in terms of tolls collected. In fact the tolls barely paid for the repairs to the road. However, it is impossible to determine whether the road might eventually have been proitable since the War between the States almost completely ruined it.

he Role of the Turnpike in the Civil War In a part of the country with few roads and diicult terrain, it is hardly surprising that both sides in the Civil War used the Weston & Gauley Bridge Turnpike. At the outbreak of the war and ater Virginia had voted for secession in April 1861, both sides sought to control the Baltimore and Ohio Railroad, which ran through northwestern Virginia. Major General George McClellan, with troops from Ohio, captured Graton from the Confederates and, ater the battle at Philippi, the federal troops controlled the Monongahela Valley and the railroad by the end of July. At this time Union troops were stationed in Weston. he federal government now decided to drive into western Virginia along the Kanawha Valley. Led by General Jacob Cox, the Union troops took Charleston in July and pushed on to Gauley Bridge. Henry Wise, the Confederate general, retreated to the Greenbrier Valley. It was also planned that troops should move from the north down the Weston & Gauley Bridge Turnpike in a lanking movement. At the end of August 1861, Colonel Tyler let Weston with the Seventh Ohio Regiment and marched down the turnpike. hey met with little resistance but were troubled by bushwhackers at Powells Mountain in Nicholas County. hen at Cross Lanes they met a Confederate force and Tyler

76

T h e We s t o n & G a u l e y B r i d g e T u r n p i k e

was forced to retreat to Summersville. General Rosecrans brought more federal troops down the turnpike and was in Sutton on September 7, 1861. hree days later they fought at Carnifax Ferry and in November at Gauley Bridge, which let the federal troops in control of the Kanawha Valley. Western Virginia remained largely in the control of the Union forces for the rest of the war but the Confederates made several attempts to break through from the east. In the spring of 1863 General William E. Jones and General John Imboden led a raid into western Virginia from the Shenandoah Valley. Jones was to march by way of Morgantown and Fairmont and meet Imboden at Clarksburg. Imboden was to arrive by way of Beverly and Philippi. General Roberts, the Union general in command of the troops in northwestern Virginia, summoned his troops from various locations, including some stations along the Weston & Gauley Bridge Turnpike, and arrived at Clarksburg before Imboden and Jones. he Confederates met at Weston. From there Jones went down the Little Kanawha Valley, burning the oil wells on his way. Imboden went south down the Weston & Gauley Bridge Turnpike, which was now reported to be in bad condition from the wagon trains retreating along it. General Roberts moved into Weston. Although the turnpike was used mainly as a way of moving troops up and down, some ighting took place along its way. here was a small engagement at Bulltown in October 1863. Early in the war a hill on the north side of the Little Kanawha overlooking Bulltown had been fortiied, probably to protect the river crossing on the turnpike there. hese fortiications can still be seen. he fort was occupied by about 400 Union soldiers under Captain William Mattingly. hey were attacked on October 13, 1863, by Confederate forces under Colonel William Jackson. Although surrounded, the Union troops refused to surrender and sent for help to Clarksburg and Weston. he Confederates camped for the night at the Salt Lick Bridge. Reinforcements arrived the next day from Clarksburg and forced the Confederates to retreat. he part of western Virginia served by the Weston & Gauley Bridge Turnpike sufered not so much from major conlicts in the Civil War but from the constant depredations of raiding parties. Braxton County in particular suffered from partisan bands. he Weston & Gauley Bridge Turnpike was used constantly by both sides moving north and south from the Monongalia Valley

77

Chapter 3

and the Kanawha Valley. he road deteriorated quickly and there were no means to keep it in good repair during the war.

Ownership of the Turnpike Transferred Ater the war the new state of West Virginia was too fully occupied in organizing its government to pay much attention to its roads. In 1863 an early act of the West Virginia legislature provided for the construction and repair “of roads and bridges” but they had very little means to implement the act. In 1866 another act transferred all turnpike roads and bridges “to the several counties in which they lie.”54 he board of supervisors in the counties was designated to take over the duties of the stockholders and to charge tolls. What became of the Weston & Gauley Bridge Turnpike from this time is unclear. A road of some kind remained in use until the formation of the State Road Commission in 1917. In the 1920s there was a movement to get the country “out of the mud” with new hard-surfaced roads. he recently formed West Virginia State Road Commission had a formidable task if the Mountain State were to achieve this goal. Morgantown was typical; in 1922 there was not a single “hard” road leading out of the town in any direction. Long sections of the old Weston & Gauley Bridge Turnpike were incorporated into what is now U.S. Route 19, which was paved in the 1920s. his new road, in general, was located in the bottom of valleys and did not run along the ridges of the hills as did the Weston & Gauley Bridge Turnpike in many places. hus, there are several lengthy sections of the old turnpike that remain in essentially original condition. Especially notable is the section from Bulltown north towards Weston (see Figure 3.11). In many ways the high hopes of the directors that the Weston & Gauley Bridge Turnpike would prove to be the artery of regional development were not fulilled. hrough the ravages of war the road was in an appalling state. With the formation of West Virginia the vision that this turnpike would be a vital link in a state-wide system remained only a dream. Even today the country through which the Weston & Gauley Bridge Turnpike passed is wild and unsettled in a great part of its course. Nevertheless, this road opened up to settlement the area in Lewis County south of Weston and 78

T h e We s t o n & G a u l e y B r i d g e T u r n p i k e

stimulated growth in population and commerce in communities such as Sutton and Summersville. In the l850s the importance and potential of the Weston & Gauley Bridge Turnpike can be seen in the establishment of turnpike roads that were intended to link with this turnpike. On March 29, 1853, the Gilmer and Braxton Turnpike Act was passed to construct a road from Glenville to the Weston & Gauley Bridge Turnpike at or near the Bulltown salt works. As early as 1849 there was interest in building a turnpike from Buckhannon to the Little Kanawha River. An act was passed establishing this road on March 15, 1849; however, the project languished and was not revived until four years later. In March of 1853 the act was revised “to construct a turnpike from Buckhannon to some convenient point on the Weston & Gauley Bridge Turnpike in Lewis County.” he Weston & Gauley Bridge Turnpike, together with its tributary roads, did establish the road pattern for the area, which changed very little until the construction of Interstate Route 79. Chapter 3 Notes 1

2

3

A brief introduction to the history of Roman roads can be found in Charles Singer, A History of Technology (London: Oxford University Press, 1957), 500–508. he History Research Advisory Committee, Virginia Highway Research Council, has prepared a number of reports dealing with early roads in Virginia. A list of publications is available, gratis, from Howard Newlon, Virginia Highway Research Council, Box 3817 University Station, Charlottesville, VA 22903. With a centralized national administration and the establishment of the Polytechnique, the irst school of engineering, the French were leaders in nearly all phases of engineering, both civil and military. his resulted in notable works such as the Languedoc Canal and the development of the analysis, design, and construction of masonry structures on impressive levels of elegance and sophistication. he new nation from the Revolutionary War until the Civil War was strongly inluenced by the French. From the point of view of engineers at West Point with French methods, using French text books is most signiicant, since engineering oicers were involved in a variety of engineering work. Most important

79

Chapter 3

4

5

6

7

of these was the construction, or more accurately re-construction, of the National Road. Before the Virginia Act of 1817 turnpikes were inanced by public funds and from the proceeds of public lotteries. Ater the Turnpike Act a new system was inaugurated that difered from either the French or British precedents. It was the formation of joint stock companies, like the British turnpike trusts, but inanced by both private and public capital. he public funds were received from city, county, and state sources. his system of mixed capital was also used extensively for railroad construction in many states. Albert Gallatin, “Report to the U.S. Senate,” American State Papers, Class X, Misc. Vol. I, 724–741. Although Gallatin’s report was inluential in its plea for internal improvements, the American transportation system of roads, canals, and railways was never planned, inanced, or built on a national basis during the nineteenth century. Maps of portions of Virginia were prepared in the pre-industrial era from the beginning of the seventeenth century until the second decade of the century. For our purposes, however, the irst signiicant map showing roads was the prestigious, but not always accurate, map of Wood, Boye, and Tanner. Between 1838 and 1850 the Board of Public Works was responsible for mapping. Of the maps produced during this period, Crozet’s map of 1848 is the most useful since it shows the Virginia turnpike system at the beginning of construction of the Weston & Gauley Bridge Turnpike. In addition to the road-building speciications that formed part of the Virginia Act of 1817, a number of text books were available to engineers, which gave design information on all aspects of road layout and construction. Construction details are given on earth, gravel, and more-permanent road systems based on the work of Tresaguet, Telford, McAdam. he most important of these books are W. M. Gillespie, A Manual of the Principles and Practice of Road Making (New York: A. S. Barnes & Co., 1848); Dennis H. Mahan, An Elementary Course of Civil Engineering, for the Use of Cadets of the United States Military Academy (New York: John Wiley, 1860), 6th ed.; Sir Henry Parnell, A Treatise on Roads (London: Longman, Rees, Orme, Brown, Green & Longman, 1833), 2nd ed.; Charles Penfold, A Practical Treatise on the Best Mode of Repairing Roads (London, 1840); John Loudon McAdam, System of Road Making (London,

80

T h e We s t o n & G a u l e y B r i d g e T u r n p i k e

8

9 10 11 12 13

14 15

16

1821); John Millington, Elements of Civil Engineering (Philadelphia: J. Dobson, Smith & Palmer, 1839); homas Hughes, he Practice of Making and Repairing Roads (London: J. Weale, 1838). Roy Bird Cooke, Lewis County Sketch Book II, 4. Cooke gives a long account of this massacre and calls it “one of the worst deeds attributed to the white settlers.” Virginia Board of Public Works Annual Report 1848–49, 99. First annual report of the Weston & Gauley Bridge Turnpike Company. Ibid. Ibid. Board of Public Works Inventory, 425, John Robinson to B.P.W., 9 April 1850. Ibid., John Wilson, J.C. Spalding, John McGee, John Brannon, Geo. W. Jackson, John Morrow, Jas. McGee to B.P.W., 27 July 1849. Ibid., J.M. Bennett to B.P.W., 7 July 1849. Ibid., John Brannon to B.P.W., 30 August 1849. It has not been possible to ascertain all the names of all the directors. he company did not report the names of the directors annually until 1859. he following is an incomplete list of the directors and the years they served: 1848–49 – James G. Neil, John H. Robinson, John Brown, John S. Camden, Jonathan M. Bennett. 1850 – Jonathan M. Bennett, Addison McLaughlin, John H. Robinson. 1851 – No record. 1852 – Wm. E. Arnold, John S. Camden, John Brown, Wm. Cottle, Addison McLaughlin. 1853 – Wm. E. Arnold, Philip Dufy, Asa Squires, John S. Camden, John Brown, Addison McLaughlin. 1854 – Morgan Dyer, Wm. Arnold, Asa Squires, John Brown, John S. Camden, Addison McLaughlin. 1855 – Asa Squires. 1856 – No record. 1857 – No record. Board of Public Works Inventory, 425, J.M. Bennett, John McGee, Lewis Maxwell, Cabell Tavenner, Ro. Ervin, Jas. Bennett, John Morrow, W. E. Arnold to B.P.W., 13 June 1849.

81

Chapter 3

17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42

Ibid. Ibid., John Brannon to B.P.W., 30 July 1851. Ibid., Felix Sutton to B.P.W., 26 October 1848. Ibid., John Robinson to B.P.W., 9 April 1850. Ibid., Proceedings of irst meeting of stockholders, 18 October 1848, 4. Ibid., J. M. Bennett to B.P.W., 7 July 1849. Ibid., John Brannon to B.P.W., 30 August 1849. Ibid., John Callaghan to B.P.W. 19 December 1849. Ibid., B.W. Byrne to B.P.W. 3 September 1851. Ibid. Ibid., W.E. Arnold to B.P.W., 26 September 1853. Ibid., J.M. Bennett to B.P.W., 26 June 1855. Ibid., J.M. Bennett to B.P.W., 30 July 1855. Ibid., Wm. E. Arnold to B.P.W., 16 September 1860. Ibid., J. M. Bennett to B.P.W., 3 September 1851. Ibid., John H. Robinson to B.P.W., 22 November 1849. Acts of the Virginia General Assembly, Vol. 1847–48, 217f. Board of Public Works Inventory, 425, Minter Bailey to B.P.W., 20 December 1848. Ibid., James G. Neil to B.P.W., 18 August 1849. Board of Public Works Inventory, 392, James Bennett to B.P.W., 9 May 1853. Board of Public Works Inventory, 425, James Bennett to B.P.W., 29 January 1849. Virginia Board of Public Works, Annual Report, 1848–49, 99. Ibid. Ibid., 1850, 339. Board of Public Works Inventory, 425, John Brannon to B.P.W., 20 January 1851. List of tolls and repair costs: Year

Tolls

Repairs

1853

$764.70

$847.12

1854

$454.45

$331.05

1855

$444.93

$159.02

1857

$871.19

$867.93

82

T h e We s t o n & G a u l e y B r i d g e T u r n p i k e

43 44 45 46

47 48 49 50 51 52 53 54

Virginia Board of Public Works Annual Report, 1854, 194. Board of Public Works Inventory, 425, report of president and directors to B.P.W., 7 June 1853. Ibid., report of president and directors to B.P.W., 23 July 1853. he Wheeling Suspension Bridge was completed by Charles Ellet Jr. in 1849 amid great public acclaim. With a 1000-foot main span it was the world’s largest bridge at the time of its construction. As a result it exerted a powerful inluence on the use of suspension bridges. In Virginia (now West Virginia) suspension bridges were built in the 1850s at Huntington, Charleston, Fairmont, and Morgantown as well as across the Elk at Sutton. See A.A. Jakkula, A History of Suspension Bridges in Bibliographic Form (Texas A&M University, 1941). Board of Public Works Inventory, 425, report of president and directors to B.P.W., 4 October 1853. Ibid., Felix Sutton to B.P.W., 17 December 1855. Tolls for the bridge were irst reported in 1857 and amounted to $87.71. Board of Public Works Inventory, 425, Felix Sutton to B.P.W., 11 February 1859. Ibid., Felix Sutton to B.P.W., 22 February 1860. John Davison Sutton, History of Braxton County and Central West Virginia (Parsons, WV: McClain Printing Company, 1919), 94. Board of Public Works Inventory, 425, John Robinson to B.P.W., 9 April 1850. Acts of the West Virginia Legislature, 1866, chapter 117, page 115.

83

Chapter 3

APPENDIX Speciications for the Salven’s cabin and Summerville Turnpike Road

Location and Grades 1st

he location is indicated by stakes numbered to correspond with the stations in the ield notes. Generally, the stakes occupy the axis of the road, and indicate the grade of its surface; but, sometimes along hill-sides and in ravines, and, also, at other places where short turns occur, the stakes shall be taken as the indication of the grade according to the ield notes; and it will be the duty of the Contractor to cut the road from this indication, in a regular shape, free from unsightly and inconvenient curves, and so that the surface of the road may conform exactly to the grade set down in the notes, which nowhere exceeds ive degrees.

Width 2nd he general width of the road is to be seventeen feet, exclusive of side ditches; but embankments, and where the stakes or the notes, indicate a curve with a radius less than 100 feet, the width is to be increased, if necessary, to the maximum of 22 feet, according to the nature of the curve, and the height of the embankment.

Ditches 3rd

here must be on one or both sides of the road, as the case may require, a ditch, not less than one foot wide at the bottom, and one foot deep, with sides not steeper than 45 degrees, and suiciently inclined to convey water freely into a proper drain. hus along a hill-side the least allowance for ditches will be two feet, and on lat ground three feet. In lat lying ground, where the ditches cannot receive an adequate forward slope, their size must be increased, and the road duly raised above the adjacent ground, according to the directions of the Superintendent.

84

T h e We s t o n & G a u l e y B r i d g e T u r n p i k e

Shape 4th

hrough nearly level ground, not exceeding a slope of eight degrees, and in deep cuts, the road must be raised in the middle, above the edge of either the ditch or the side slope, one twenty-fourth part of its width. Also, embankments, where nearly level, must be rounded as just mentioned; but if they slope more than two degrees forward, they need not be raised more than half that height in the middle. An additional elevation of one-tenth of their intended height must be superadded to embankments at each point, in order to allow for settling. Along hillsides, the slope of which exceeds eight degrees, and the radius of the curve of the hill is short, the surface of the road must be made lat and inclining to the hill, so that the outer edge may be raised higher than the edge of the ditch, one-twentieth part of the width of the road. In other places along such sills, where the road is nearly straight, or where the radius of the curve extends from the hill, only two-thirds of the road shall slope to the hill, and the other third must be made level, so that when it settles it will incline to the lower side of the road.

Slopes 5th

he upper slopes, cut out of a hill, shall not be less than one-half of the angle made by the declivity of the hill with the vertical. he exterior slopes shall be those naturally taken by loose earth.

Construction 6th

Every tree within the width occupied by the road and ditches must be grubbed. Every rock or stump, showing at the surface, within the said width, is to be cut down to a depth of at least two feet below the surface of the road.

7th

Beyond the ditches, every tree within 30 feet of the centre of the road must be cut down; and where the outside of the road is suiciently level for carriages, the stumps must be shaped in a conical forms, the timber being removed from the said space.

8th

he irst operation ater the grubbing must be the removal of the vegetable mould from the surface to the foot of the embankment.

85

Chapter 3

9th

No perishable material is to be sufered to remain or be placed in the body of the road. No stone weighing more than six ounces is to be let on its surface.

10th When the exterior slope of the road is exposed to abrasion from a current at its base, it must be protected by stones of a suicient size to a suitable height.

Walling 11th Sustaining walls, when found indispensible, must rest upon a irm and level foundation. heir base must be at least two-iths, and their top one-fourth of their height. hese walls, when not otherwise stated, shall be made of dry masonry. he stones must be carefully laid, with due proportion of headers and stretchers, to the satisfaction of the Superintendent. In general, however, it will be preferable to let the earth take its natural slope, and resort to walls only when unavoidable. Timber supports should never be permitted, unless absolutely unavoidable for want of rock within practical distance.

Drains 12th he draining of the road must be carefully attended to. It is to be efected by adequate ditches, by raising the surface at least two feet above any body of water that may, at times, accumulate near it and by frequent gutters or culverts. Gutters are to be preferred where the water can be conducted over them. hey shall consist of a bed of broken stones, at least nine inches thick, on the whole width of the road, and held up at the lower side, either by a stone wall or by the natural slope of the bank, protected by a layer or large stones laid on its surface. On either plan, the outer edge of the gutter must be formed by large lat stones inclining inwardly. he size of each gutter will be regulated by the length of the road, the height of the adjacent hills, the extent of the valleys to be drained, and, in one word, by all the circumstances that may inluence the quantity and rapidity of the water discharged.

86

T h e We s t o n & G a u l e y B r i d g e T u r n p i k e

No gutter, however, shall be less than 12 feet wide, unless for reasons appearing in the notes. he depth of each is to be one twenty-fourth part the width;— this dimension is generally stated in the notes, but may be varied by the Superintendent. he fall of the gutters must not be less than three-fourths of an inch nor more than two inches per yard. here must be a gutter in every depression, whether or not designated in the notes. And at least one drain for every two hundred yards of declivity, to be located, if omitted in the notes, opposite hollows, in convex turns where the water naturally tends to cut across the road, or in artiicial turns at the point where cutting changes in embankment. Where along a graded hill, a gutter is to be made, the Contractor will be required to preserve distinctly, the level distance appropriated for the same. He will not be allowed, in any case, to cut it out of a graded surface. Along hill-sides, on approaching a gutter, it will generally be necessary to lessen, gradually, the inclination of the road to the hill, so as to make the transition from the inward slope of the road to the outward descent of the gutter, regular and easy. When large gutters are made across water courses, the stones composing the walls by which they are held up, must be lat, and laid sloping, with their outer edge raised, and the illing must consist of stones and gravel of as small a size as the current will permit. 13th Culverts, unless diferently stated in the notes, and only then when unavoidable, must no where be less than eighteen inches square, and this only when the water falling on or immediately near the road, has to be passed under it. here small culverts to consist of two small parallel walls, at least one foot thick, or curb stones covered with large lat stones, not less than six inches thick, and at least two feet of earth. Where stones are not to be had, substantial pieces of sound white oak, may, with the approbation of the Superintendent be substituted, the whole trough being well pinned together. Such culverts should have no earth over them. For large culverts requiring arches to be built, if any occur, special speciications shall be made. Small bridges will, generally, be found cheaper, and otherwise preferable.

87

Chapter 3

Upper Ditch 14th Where a long descent occurs, and the ground will allow it, a ditch may be cut out of the hill above the upper slope of the road, and parallel to it, to convey the water to the next valley or ravine, whereby some drains may be dispensed with, and the road be better preserved.

Swamps 15th Where swampy places are crossed an embankment must be raised at least two feet above the surface, and covered with clean gravel or broken stones of the usual size, or else in wet foundations an even bed of at least two layers of fascines, no where less than one foot thick, must be prepared, to support an embankment two feet thick of earth with gravel or broken stone, as just mentioned, ditches being besides cut on one or both sides, as usual, and culverts placed under the embankment, at proper intervals, unless it be practical to carry of the water at the sides. 16th If the soil any where should require a capping of stone, care must be taken so to shape the road that the capping superadded may give it its proper height and dimensions. his capping is, in general, to consist of a stratum of about ten inches thick, and composed of broken stones of not more than six ounces of weight, laid at three diferent times.

Bridges 17th For large bridges plans and speciication will be furnished; they are not to make part of the road sections. But the common bridges not exceeding 40 feet will be included in the contract for the sections they belong to. hey must be made 18 feet wide in the clear for a double track, and only 12 feet if only one track is required. his last dimension will be understood to be intended when not otherwise speciied. he abutments, when needed, to consist of dry masonry, laid carefully on a irm foundation the irst course of stones to be large and lat, and the other courses to consist of a due proportion of headers and stretchers, there being at least one header in every ive feet of the face of each course, and each header corresponding to about the middle of the interval of the headers of the preceding course, and projecting at least one foot and a half back of the stretchers, and none of the latter to be less than six inches high and one foot thick, nor longer than four times its height.

88

T h e We s t o n & G a u l e y B r i d g e T u r n p i k e

he average thickness of the abutments and wings shall not be less than onefourth of their height, and no where under 18 inches. he out-side batter shall not exceed two inches to one foot. When the length of the bridge is subdivided into small spans and trestles are used, they must consist of three or four uprights (according to width) [which] are to be laid at equal distances from each other, the sleepers being 8 inches by 12. heir ends to rest either on sills at least 12 inches square, inserted into the top of the bearers balanced over the trestles, to which bearers the sleepers must then be secured by iron bolts or clamps. he ends of the sleepers meeting on the same trestle to be pinned together. On the sleepers a looring two and a half inches thick, at least, is to be laid, with a substantial railing on each side, leaving a clear space of 12 or 18 feet of bridgeway. he planks to project nine inches beyond the sleepers. In the loor, joists ive inches square, are to intervene among the planks every 8 feet, their tops being made even with the planks by notches over the sleepers. he joists will project on each side three feet beyond the sleepers, and serve both to stifen the bridge and support the railing as follows: A string piece 8 inches square being now laid along the loor, above each one of the exterior sleepers, and connected with it by screw bolts three quarters of an inch in diameter and eight feet apart; upright posts ive inches square and eight feet asunder are to be spiked on the string pieces, each post on one of the joists, to the projecting ends of which they must be braced by pieces four inches square at about 45 degrees. he posts being then capped by a rail four inches by ive, and the spaces between them occupied by St. Andrew’s crosses three inches square, will complete the railing, the whole height of which must be at least four feet. he railing to be painted, and the sleepers being either painted or pitched, and a board about two inches wider nailed on their upper surfaces before laying the loor. he timber to be of a durable kind, such as white oak, heart pine, black walnut, and the pins black locust. 18th he elevation of each bridge must be such that there may be a short ascent to it at each end, with a paved gutter in the depression, which should be at least as low as the top of the abutment. 19th Near each bridge a descent not exceeding ive degrees must be made to a ford, if such exists; this secondary way is to be only 12 feet wide.

89

Chapter 3

20th Mile posts of locusts to be irmly planted all along, and on the same side of the road, showing the distance both to the Staunton and Parkersburg road and Summersville. 21st he Contractors will be required to keep each mile of the road in repair for one year from the time it shall be accepted. JAS. BENNETT, Eng. And Supt. Signed May 8, 1852 as part of the contract to which it was amended. Jas. Bennett, Supt. H. S. Moore

90

4

he Pulaski Skyway—Railway Economic heory Applied to Superhighway Design Dara Callender and Emory L. Kemp

E

arly in the twentieth century, highway planning and design were widely discussed. he diferent types of roads and their requirements were matters of signiicant disagreement. Many planners and politicians believed that two distinct types of highways could never coexist in an interconnected network of roads. Scenic and commercial routes were viewed as separate entities, and each road type had its detractors as well as its supporters. Many engineers could not envision the need for high-speed commercial routes until motor trafic had begun to visibly overtake all other forms of transport. For this reason, the high-speed commercial-route design type was slow to gain acceptance by planners and highway engineers. he use of scenic roads was not widespread either, and most highway designers chose to ignore the possible utility of the parkway design concept in their planning. However, shortly ater 1900 the principles of the modern scenic highway, or parkway, were deined. Developed in the late nineteenth century, the American parkway road type was based on the concepts inherent in England’s Garden Cities Movement. Although parkway development never became widespread in the United States, the New York City area became the focus of extensive, limited-access, parkway planning and construction. In the 1920s New York State began the country’s earliest experimentation in controlled access with the construction of parkways that were the pioneer examples of scenic highways. he irst such route, called the Bronx River Parkway, incorporated the natural beauty of its surroundings in the design and appearance of both the route and the structures thereon. his parkway and the others that followed, primarily in Long Island and Westchester County, were 91

Chapter 4

milestones in national highway planning; however, this type of route, intended for slower-speed passenger automobile traic, was not seen as the means through which New Jersey could accommodate the enormous volume of industrial and passenger traic found in the densely populated northern region of the state. During New York’s intensive parkway planning and construction period, New Jersey’s Highway Commission chose, instead, to experiment with the relatively new high-speed, limited-access, commercial route called the “superhighway.” Although a number of early 1920s authors used the word “superhighway” in their writing, no absolute deinition of the term existed. New Jersey state engineers, before and into the 1920s, performed numerous studies toward the goal of addressing the most critical route problems within the highway system. he most important of their concerns was the improvement of the state’s oldest route, United States Route 1. he joint planning, by the states of New York and New Jersey, of the design and construction of the Holland Vehicular Tunnel under the Hudson River forced the latter state to make its most diicult planning decision to date. It was clear that the completion of the tunnel would place numerous vehicles, which had previously entered Manhattan by ferry, onto the already busy local streets of Jersey City and the surrounding municipalities. Both the Regional Plan Association and New Jersey engineers and oicials were well aware that these roads could not accommodate the additional through traic that would be generated by the proposed Hudson River tubes. Although the Holland Tunnel proposal did not initiate New Jersey’s highway improvement plans, it did serve to focus and solidify them. he irst goal of state engineers prior to the planning of the tunnel had already been the easing of traic in the northern portion of the state, and the accommodation of vehicles transporting goods and persons across the region. By the 1920s, the extension of area routes to accommodate anticipated tunnel traic became an even more pressing need. herefore, New Jersey state engineers proposed the extension of Route 1 with a new highway across the region. By the time of the 1927 opening of the Holland Tunnel, permitting the irst motorized access between New York and New Jersey, the irst portions of the new Route 1 Extension were completed. his construction was quickly followed by the further expansion and improvement of the area highways, providing faster and more eicient access to New York City. 92

T h e P u l a s k i S k y way

Once the New Jersey State Highway Department had clariied the need for a project to accommodate future tunnel traic and to ease highway congestion between Jersey City and Elizabeth, and the Route 1 Extension project had been initiated, a design group was chosen by then-State Highway Engineer William Sloan, a railway engineering expert. Critical to the future work of his team was the fact that the two senior design engineers, Frederick Lavis and Sigvald Johannesson, also came from the railroad-engineering community. Traic studies of the 1910s and 1920s had illustrated that the increase in the volume of motorized traic was beginning to overwhelm the industrial area of northern New Jersey. he state’s Engineering Advisory Board clearly indicated the need for signiicant highway improvement for the region to accommodate both municipal concerns and inancial goals. he board developed a proposal for the Route 1 Extension, and the highway planners began to perform analyses for the establishment of a workable plan for the river crossings that would satisfy both economic concerns and the requirements of vehicular and navigable traic. he nation’s highway-and-viaduct project of the 1920s, which epitomized the characteristics of the superhighway, was the Route 1 Extension. Wyatt Brummitt was only one of many authors on highway engineering of the time who referred speciically to this design as the standard maker for the form: “A superhighway was not a speciic highway type, but rather a new transportation concept. Superhighways were routes that successfully expedited the existing transportation low, thereby creating an express route. . . . [and] New Jersey took a leading role in the development of superhighway systems nationally.”1 his was the highway type that New Jersey worked to deine, beginning with the design and construction of the Routes 1&9 Corridor and the Pulaski Skyway. he Skyway was completed in 1932 to connect the industrial heart of northern New Jersey with the Holland Tunnel into New York City. (See Figures 4.1, 4.2 and 4.3.) Its construction coincided with improvements to the New York metropolitan area highway system recommended by the country’s irst regional planning organization, the Regional Plan Association. he planning of the route (overleaf ) Figure 4.1. Photograph of Pulaski Skyway over Hackensack River spans, 1983. (Library of Congress, Prints and Photographs Division, Historic American Engineering Record, HAER NJ, 9-JERCI-10-6)

93

followed decades of improvements in national and state roadway construction and design. he vast size and impressive silhouette of the Pulaski give the viewer only a sense of its structural signiicance in relation to the development of the limited access “superhighway” form. he bridge was, when completed, the 94

world’s longest high-level vehicular viaduct. However, it is not with the design of the structure itself that the following narrative is concerned. Perhaps the greatest historic importance of the Skyway is the role that it played in the transition from railway to highway transportation systems in this 95

Chapter 4

country. he New Jersey state design team that planned the Pulaski and the approaching sections of the Routes 1&9 Corridor utilized and revised longused railroad economic planning standards. heir project appears to have been the irst American highway design to incorporate extensive economic analyses and predictive formulae for future usage. It was certainly the largest such design and served as a means of educating those in the highway engineering profession of the value of economic design.

he Structural Features of the Pulaski Skyway he Pulaski Skyway is a four-mile-long, 108-span steel viaduct connecting the city of Newark in Essex County with the Holland Tunnel approaches in Jersey City, Hudson County, New Jersey. he structure currently carries Routes U.S. 1&9 over the Hackensack and Passaic rivers, the Hackensack Meadows, the New Jersey Turnpike, Port Authority Trans-Hudson (PATH) Rapid Transit, former Erie Lackawanna and Penn Central railroad tracks, and local streets. Originally called the New Jersey High-Level Viaduct and the Newark–Jersey City Viaduct, the bridge is accessed via end ramps in Jersey City and Newark and two intermediate ramps in Kearny and Jersey City. he portions of the highway approaching the Skyway were originally designated as the Route 1 Extension project, carrying U.S. Route 1 and later N.J. State Route 25, and were named the Diagonal Highway in period literature. Currently known by the overall designation the Routes 1&9 Corridor, the entire route extends from Jersey City through Newark and into Elizabeth, and consists of cut (excavated) sections and steel-and-concrete viaduct spans. he original approach sections of the Corridor have been changed since they were opened to traic because of additional construction. However, the portion of the highway that encompasses the Pulaski Skyway retains its original structure and characteristics. he overall width of the Skyway is 56 feet 6 inches, with a clear roadway width of 48 feet 6 inches between curbs east of and through the Jersey City ramp, and 50 feet west of the ramp. here are two traic lanes in each direction, which were originally undivided but are presently separated by a steel median barrier that was installed in sections in 1955 and 1969. At the Broadway and 96

T h e P u l a s k i S k y way

Kearny ramps the mainline roadway width in each direction is 30 feet, separated by 24-foot-wide ramp roadways that enter the structure at the median location. Sections of the bridge were originally surfaced with asphalt or granite paving blocks, which were later replaced with reinforced concrete and paved with a bituminous material. he roadway surface is now composed of latex-modiied concrete (LMC), which was placed in 1982. Safety walks with integral steel-armored concrete curbs of varying height exist at both sides. he reinforced concrete fascia balustrades in the eastern river approach span are 3-foot-6 inches high, and the 4-foot-high metal railings in the remaining spans consist of steel channel rails and I-beam posts. he deck expansion joints are of several types, including steel angle and channel armoring systems with asphaltic iller material and open steel inger joints. Light standards, staggered from side to side of the bridge, are mounted to both fasciae. Drainage is provided by catch basins and cast-iron pipes in the eastend sections, and by open sections of curb throughout the remainder of the structure. he Skyway structure was designed to support H20 truck loading, weighing a total of 20 tons, as required by the existing American Association of State Highway Oicials Speciications of the late 1920s. Trucks were banned from the Skyway by Jersey City in the mid-1930s and by Hudson County in 1954. he structure currently carries only automobiles. Recent literature indicates that some methodology for the design of the viaducts was taken from American Railway Engineering Association speciications of the time, due to the lack of highway viaduct guidelines. Construction on the corridor approaching the river spans was begun in 1926, and the Skyway was erected between 1930 and 1932. Numerous contractors were used in the erection of the Pulaski Skyway itself. he general contractor for Section 1 was Charles T. Kavanaugh of Bayonne, New Jersey, and the roadway general contractor for Sections 2 through 8 was New York City’s Fredburn Construction Company. A number of small irms were hired as structural subcontractors and to perform the general roadway work. he substructure elements were built by the Tunnel Construction Corporation, Arthur McMullen Company, and Guarantee Construction Company of New York City, and the Foundation Company of Newark, New Jersey. 97

Chapter 4

Four nationally prominent twentieth-century bridge manufacturers were involved in the construction of the Skyway superstructure. Taylor-Fichter Company of Jersey City and New York City erected many of the eastern and western spans of the bridge, including all spans west of the Passaic River. he American Bridge Company of New York City was responsible for the remaining spans east of the Hackensack River. Pennsylvania’s McClintic-Marshall Company, working out of oices in Hackensack, New Jersey, built the Hackensack and Passaic River cantilever spans as well as the deck truss spans east of the Passaic. he deck truss spans west of the Hackensack River were built by the Phoenix Bridge Company of Pennsylvania. Figure 4.3. Cross-section drawing of Pulaski Skyway at cantilever span (Hackensack and Passaic Rivers) at panel points U1 and U2. (Route 25, Connecting Link, Sections 3 and 6 [Contract Drawings], 1929, New Jersey Highway Department)

98

T h e P u l a s k i S k y way

Several features of the Skyway and the adjacent Routes 1&9 Corridor were unique to highway construction of the 1920s and 1930s. he project was the irst in the United States in which concrete approach slabs extending beyond the abutments were utilized in an attempt to reduce diferential settlement. Although the system was not completely successful, it was important as an example of engineering theory being put into practice. he building of ramp systems was in the experimental stages in the early 1930s. he use of ramps entering the highway at the center of the deck was unique, and it required extensive analysis and detailing to permit the accommodation of the ramps between the trusses. he viaducts designed for the Corridor were longer than any others being built at the time. hey also utilized the relatively new practice of encasing steel members in concrete to provide protection, which was irst widely used for bridge construction in New Jersey in 1931. Experimentation was also used in the design of the deck expansion joints. As construction on the corridor progressed it became possible to determine what details were not working and to design and use diferent ones. he joints on the Skyway are located approximately 600 feet apart and have to accommodate at least 6 inches of movement. To provide this expansion, cast inger joints, which were later to become a familiar detail on long-span bridges around the country, were designed. he Skyway structural elements are connected using more than two million rivets. A total of 88,461 tons of structural steel were used in the construction of the Pulaski. his amount exceeds that used to erect the George Washington Bridge by approximately 20,000 tons.

he Birth of Modern Highway Economic heory During the second half of the nineteenth century, railroads were the primary means of transport for both goods and people within the United States. Railway design engineers learned through years of experience and experimentation the ways in which to lay out the most cost-efective routes. By the late 1880s scientiic methods of railway design and management were the norm. In the early 1900s, however, when much of the transportation focus shited to 99

Chapter 4

roads, no such standards of practice existed for highway design. Many railroad engineers, inding available work to be limited, went to work in the design of highways, taking with them their railway planning expertise. It is from early railroad practice that modern highway economic theory was formed. What follows describes the methods of layout utilized by railroads, and the gradual processes through which highway engineers began to discuss and establish similar analyses for roadway design. It provides the reader with a detailed discussion of the principles taken and revised from these early railroad methods, which were deined and used by Frederick Lavis and Sigvald Johannesson in the design of the Route 1 Extension in New Jersey. his new economic planning theory proved to be the beginning of modern highwaydesign practice.

he Deinition of Wellington’s heory of Railway Location Prominent nineteenth-century railway design engineer Arthur Mellen Wellington stated that “railways are not undertaken unless they are to be proitable, not to the general public, nor to other parties in the near or distant future, nor to those who lend money on them, but to those who at irst control the enterprise.”2 From the early days of railroading, attempts were made to determine the controlling factors that would enable railway owners to proit through the establishment of these transportation corridors. By the 1870s, railroad engineers such as Albert Fink had developed systems for the systematic construction and management of railways. Fink’s cost-accounting system, used by many American railroads of the time, allowed for new statistical control. he railroad companies of the 1880s were divided administratively into the functional departments of inance, transportation, and traic through an organizational form established at the Pennsylvania Railroad, which was combined with Fink’s cost system. his railway-management system was practiced by nearly all railroad companies for decades to follow. Although many railroad planners worked toward the standardization of cost-eicient design systems, it was Arthur Wellington who ultimately stood out as the leader in his ield. Wellington worked as Principal Assistant Engineer for 100

T h e P u l a s k i S k y way

Location and Surveys for the Mexican National Railway, and later as Assistant General Manager for Location for the Mexican Central Railway; he published, in 1887, the text on location that became the norm for railway engineering use. Critical to his theory was the statement that a railroad was a form of industry, which he deined as the manufacture of transportation. Transportation, indeed, existed before its invention, . . . but it was . . . mainly provided on a small scale by each consumer for his own use and his immediate neighbors’. With the invention of the railway irst began the manufacture of transportation for sale on a large scale and by modern processes.3

Wellington stated that even the strongest of railroad corporations existed within very narrow inancial margins for error. Even relatively small variations in initial costs, operating expenses, or revenues could mean the diference between success and failure for many proposed or operating lines. Every recommended increase in expense at any point of construction or operation had to be thoroughly veriied and justiied in terms of its future impact. he Wellington heory of Railroad Location was a means of comparing two or more proposed locations by estimating costs and then estimating the operating costs per train car considering the efect of variables such as alignment, curvature, gradient and distance. Using the train mile as a base, one could calculate the relative cost of each proposed location.4

he cost per train-mile was deined as the expense incurred in operating a given train over a length of track of a single mile. he following elements, and their comparative costs, backed by extensive historic cost documentation, were used in the presentation of the theory: fuel, water, oil and waste, engine repairs, switching engines, train wages, train supplies, car repairs, car mileage, rail renewals, adjusting track, tie (aka sleeper) renewals, earthwork and ballast, yards and structures, and station, terminal, general, and tax costs. (See Table 4.1.) he principles of Wellington’s economic theory were based on the determination that three minor and two major details of alignment were of signiicance in inancial considerations. he minor details, deined as such due to their lesser inluence on 101

Chapter 4

Table 4.1. Approximate Estimate of the Details of Operating Expenses for an Average American Road. (Source: “he Economic heory of the Location of Railways” by Arthur Mellen Wellington, 1902, New York: John Wiley & Sons)

Engines 18.0 %

Road Engines 14.4 %

Fuel Water Oil and Waste Repairs-Engines

Switching Engines Train Expenses 47.0 %

Switching Engine Wages Train Wages and Supplies 17.0 %

Maintenance Of Way 23.0 %

7.6 % 0.4 % 0.8 % 5.6 % 3.6 %

Train Wages and Supplies 15.4 %

Engine Wages Car Wages Car Supplies

1.6 % 6.4 % 8.5 % 0.5 %

Cars 12.0 %

Repairs and Renewals Mileage (a practical equivalent for repairs)

10.0 % 2.0 %

Track Between Stations 8.0 %

Renewals of Rails Adjusting Track

2.0 % 6.0 %

Road Bed 7.0 %

Renewing Ties Earthwork, Ballasting, etc.

3.0 % 4.0 %

Yards and Structures 8.0 %

Switches, Frogs, Sidings Bridges, Masonry Stations, Other Buildings

2.5 % 3.5 % 2.0 %

Total “Line” or Transportation Expenses Station, Terminal, and General Expenses and Taxes Total Operating Expenses

70.0 % 30.0 % 100.0 %

the future of railroad property, were the elements of distance, curvature, and rise and fall. he more critical elements for consideration were the amount of trafic on a given line and what Wellington termed the “ruling gradients,” so called because these grades were the worst encountered on the entire line. he cost typically utilized for purposes of comparison was one cent per “train-mile.” Despite 102

T h e P u l a s k i S k y way Table 4.2 Estimated approximate efect of great and small diferences of distance. (Adapted from Arthur Mellen Wellington, he Economic heory of the Location of Railways [New York: John Wiley & Sons, 1902]).

Item

Total Cost in Per Cent

Increase for Greater Diferences of Distance of Per Cent Varying with Distance

Total Amount Per Train-Mile

Fuel

7.6

from 67 percent to 85 percent

6.5

Water

0.4

from 0 percent to 50 percent

0.2

Oil and Waste

0.8

from 50 percent to 50 percent

0.4

Engine Repairs

5.6

from 40 percent to 57 percent

3.2

Switching Engines

5.2

from 0 percent to 0 percent

0.0

Train Wages

14.9

from 0 percent to 100 percent

14.9

Train Supplies

0.5

from 0 percent to 40 percent

0.2

Car Repairs

10.0

from 35 percent to 50 percent

5.0

Car Mileage

2.0

from 100 percent to 100 percent

2.0

Rail Renewals

2.0

from 80 percent to 100 percent

2.0

Adjusting Track

6.0

from 50 percent to 100 percent

6.0

Tie Renewals

3.0

from 100 percent to 100 percent

3.0

Earthwork and Ballast

4.0

from 100 percent to 100 percent

4.0

Yards and Structures

8.0

from 0 percent to 50 percent

4.0

Station and General

30.0

from 0 percent to 0 percent

0.0

Total

100.0

from 24.8 percent to 51.4 percent

51.4

this deceptively low cost, Wellington observed that revenue and expenses could be signiicantly inluenced by even relatively slight variations in design. he efect of distance, or length of line, according to Wellington, varied in a fashion dependent upon the factor under comparison. Some elements were ixed and varied only slightly, if at all, with changes in length, while others were considerably impacted by each increase in mileage. he general statement could be made that for short increases in length little change was observed in expense; however, as distance increases became greater the expenses connected with the increases varied more directly with the distances covered. (See Table 4.2.) Curvature impacted railroad operating costs by causing possible losses of power, limiting train lengths, and increasing the need for periodic conirmation of safe track conditions. In addition, sharper curves increased the danger 103

Chapter 4 Table 4.3 Estimated Cost Per Train-Mile and Per Daily Train of 26.4 Feet of Rise and Fall. (Source: Arthur Mellen Wellington, he Economic heory of the Location of Railways [New York: John Wiley & Sons, 1902]).

Item

Total Cost of Item

Percentage of Same Increasing with 26.4 Feet of Rise and Fall for Class C

Cost Per Train Mile of 26.4 Feet of Rise and Fall

Fuel

7.6

100.0 percent

7.6 percent

Water, Oil and Waste

1.2

50.0 percent

0.6 percent

Repairs of Engines

5.6

4.0 percent

0.22 percent

Switching-Engine Service

5.2

Unafected

0.0 percent

Train Wages & Supplies

15.4

Unafected

0.0 percent

Repairs of Cars

10.0

4.0 percent

0.4 percent

Car Mileage

2.0

Unafected

0.0 percent

Renewals of Rails

2.0

10.0 percent

0.2 percent

Adjusting Track

6.0

5.0 percent

0.3 percent

Renewing Ties

3.0

5.0 percent

0.15 percent

Earthwork, Ballast, etc.

4.0

5.0 percent

0.2 percent

Yards and Structures

8.0

Unafected

0.0 percent

Station and General

30.0

Unafected

0.0 percent

Total

100.0

9.7 percent

9.67 percent

Per Foot of Rise and Fall

0.366

Per Foot of Rise and Fall per Daily Train – Class C (round trip)

$2.67

of derailment or collision, slowed travel speeds, and reduced the feasibility of longer or heavier cars. Wellington noted that the most important and expensive location detail relative to construction costs was sharp curvature. herefore, he saw reason for establishing, in design, maximum curvatures, and determining the costs inherent in exceeding such curves. Two causes of increased expenses were related to rise and fall, or the “elevations overcome by [an] engine on gradients not exceeding in resistance the maximum of the road, and hence not limiting the length of the train.”5 he irst was direct costs in terms of the fuel used in overcoming changes in grade, and in terms of the wear and tear on equipment ascending and descending a given grade. “he second objection to gradients [was] . . . the efect which the maximum or rather ruling grade . . . [had] to increase the cost of operating the entire line, . . . not by increasing the direct expense per train-mile, but by limiting the 104

T h e P u l a s k i S k y way

number of cars to a train.”6 Wellington observed that the general cost of rise and fall was in direct proportion to the elevation climbed, and that the ruling grade cost was directly afected by the rate of the grade. (See Table 4.3.) he ruling grade was that which most adversely afected a railway’s performance and future development. Although the efect of grades could be overcome via economies of power expenditure or choice of train length or weight, careful consideration of grades was preferable to dealing with the limitations and potential dangers inherent in using steep grades. Wellington’s general rule regarding gradients was stated as follows: Follow that route which afords the easiest possible grades for the longest possible distances, using to that end such amounts of distance, curvature, and rise and fall as may be necessary, and then pass over the intervening distances on such grades as are then found necessary.7

he amount of traic a line could be anticipated to see was the result of the combination of the same factors previously discussed. Important to increasing traic was the use of the least number of trains per mile, and the careful choice of the line location and the number of “traic points” on that line. Each traic point was a tributary source of additional traic that would directly impact the general railway earnings: he equation giving the general law of increase in earnings due to an increase of tributary traic points on the same length of line . . . [showed that] the productive traic varie[d] as the square of the number of tributary sources of traic.8

his law of increment in traic was Wellington’s means of indicating a basis for the assumption of increased inancial productivity and acceptability of expenditure. (See Table 4.4.) Wellington’s theory, as published in editions during the closing decades of the nineteenth and opening decades of the twentieth centuries, was utilized throughout the railroad industry. By the time of the release of the 1936 edition of the manual of the American Railway Engineering Association’s Committee on the Economics of Railway Location, a formula had been established for 105

Chapter 4

determining the economic value of a given location. his formula was based upon the ratio between the original invested capital and the annual operating revenues less operating expenses. Train-mile costs remained in use by the railroad engineering community throughout the decades preceding and during World War II and were presented for general use not only by the American Railway Engineering Association (AREA), but also within standard engineering texts. Some highway engineers of the early twentieth century were beginning to study the possible efects of construction and vehicle operation costs in their designs, as railroad engineers had long done. However, no published documentation similar to that for railway design existed in which the theory of highway layout was deined. During the 1920s engineers were still attempting to deine a comprehensive location theory for road use. Due to Frederick Lavis’s railroaddesign experience he was intimately familiar with Wellington’s documented location theory. He was perhaps more familiar with this type of analysis than any other highway engineer of his time, and it was Wellington’s theory that provided the key to Lavis’s planning of New Jersey’s new superhighway, and the deinition of highway-location analysis that he and Johannesson formulated for their work.

he Development of the heory of Highway Location By the early 1920s engineers had begun seriously considering the need for highway design utilizing economic theory. Most American roads had originally been based on pioneer trails, the locations of which had changed little with time. Roadway routes were irmly set, as were the social concepts of how road design and funding should take place. Some of the primary factors originally inluencing roadway location had been cost, existing laws, the placement of existing centers of population, and the transportation needs of the public. here had been a relative lack of improvement in highway location practice over time. Roads [had] grown up from the pioneer trails with but small regard to the possibilities of location to accommodate more than the traic of the day. . . . Small use of the science of highway location ha[d] been made; . . . it ha[d] been found well nigh impossible to abandon the old road[s] and the projection of new roads . . . ha[d] been retarded by laws protecting the established order.9

106

T h e P u l a s k i S k y way

Number of Traic Points

Names of Traic Points

Increment Due to Each Addition

Total Traic

Table 4.4. Illustrating the Law of Increment in Traic Resulting from the Interpolation of the Additional Traic Points. (Source: Adapted from Arthur Mellen Wellington, he Economic heory of the Location of Railways [New York: John Wiley & Sons, 1902]).

1

A

0

0

2

B

1

1

3

C

2

3

4

D

3

6

5

E

4

10

6

F

5

15

7

G

6

21

8

H

7

28

A

A

B

A

C

B

C

A

D

B

D

C

D

A

E

B

E

C

E

D

E

A

F

B

F

C

F

D

F

E

F

A

G

B

G

C

G

D

G

E

G

F

G

H

B

H

C

H

D

H

E

H

F

H

G

H

(Note: he comparative aggregate traic for any number of traic points n is given by the sum of the natural numbers to n-1 inclusive.)

United States road building throughout history had been based primarily on cost, and work had generally been funded through some form of taxation. It had long been fundamental to roadway development that any new or improved construction being proposed not cost more than the beneits which could be realized upon completion of a given project. By the 1920s some writers believed that the time had come for a diferent way in which to measure the beneits of 107

Chapter 4

road improvements. hey felt that the economic gains to be considered should include both tangible (savings in construction and operation costs, and gains in time and convenience), and intangible (“appreciation on real estate value, . . . reduced cost of merchandise and produce due to lowered transportation costs, [and extended] range and ease of travel”10) beneits. However, many planners were uncertain as to how to efect changes in engineering practice, and engineering literature of the time gives no indication of any attempts being made to weigh these factors. Prior to the start of the twentieth century, engineers coming from the railroad community had provided the only source of economic analysis. One aspect of engineering design that was directly applicable to both railways and highways was the study of bridges. One of the premier bridge engineers of the late nineteenth and early twentieth centuries, John Alexander Low Waddell, wrote and published numerous papers during his long career on various aspects of the economic design of bridges. However, even Waddell focused primarily on bridges’ structural elements, and presented little information on the costs of overall layout or vehicle expenses. His Economics of Bridgework,11 published in 1921, compared the cost-efectiveness, in very general terms, of structure types, and provided documentation on the most feasible uses of materials. Despite the signiicance of his valuable “List of Factors and Conditions Afecting the Layouts of Bridges,” his publication provided the highway engineer with only a basic understanding of the importance of comparative cost studies. Waddell deined economics as the means of obtaining one’s goals while absolutely limiting one’s expenses. He wanted young engineers to realize that “the determination of the best possible layout for any proposed structure [was] . . . truly an economic problem . . . he general idea that the best possible layout [was] the one which makes the irst cost of structure a minimum [was] a fallacy.”12 However, he did not provide the means of performing the economic analysis necessary to the design and construction of the emerging modern highway bridge within this text or its more famous predecessor, Bridge Engineering.13 he deinition of an economic theory for roads took time to develop. Despite the fact that 1920s designers were beginning to interpret highway location in terms of economic feasibility, most continued to speak and write 108

T h e P u l a s k i S k y way

in very general terms. Consideration was given to the comparison of possible variations in alignment and their feasibility over various types of terrain and for various types of vehicles. No publication yet provided inancially based methods indicating how designers should perform their analyses. he inadequacy of available information was keenly felt by engineers working in densely populated districts, due to the intensiied problems of traic volume and roadway mileage. Many writers of the period indicated that highway personnel should look to the experience of railroad engineers as the most valuable existing source of planning information. “However, that experience [was] . . . becoming harder to get as the railroads [were] . . . not building many new lines.”14 Even as motor-vehicle volume rapidly increased many American engineers continued to believe that vehicular traic was not and never would be the primary means of transport in America. For that reason, some writers of the time did not believe that exhaustive economic design was always justiiable. Despite this hesitancy on the part of many, commentary by engineers from England, such as H.T. Tudsbery, indicated that by the early 1920s American engineers were already better equipped than the British to deal with the problems of scientiic roadway engineering. Traic in urban centers around the world was increasing so steadily that engineers like Tudsbery had become concerned with accommodating it using economic means. In 1925, Tudsbery published an article in the British periodical Engineering,15 wherein he noted that the costs of operating motor vehicles varied not only by vehicle type, but also for any given vehicle based upon delays, grades, weather conditions, and fuel and maintenance costs. He provided an analysis of the cost of operating various vehicles under various conditions, including the combined costs of gas (gasoline) and oil, wages, tires, repairs, depreciation, and overhead for commercial vehicles. He computed his costs per truckmile, per ton-mile, and per foot of rise and fall. His writing discussed concerns regarding at-grade crossings, and the value of regulation and separation of traic of difering speeds to maintain road capacity on heavily traveled routes. Traic delays were of great signiicance to Tudsbery’s analysis. He observed that if the cause of delays could be eliminated, a signiicant savings in both eiciency and actual operating costs might be achieved. Tudsbery observed that economic highway location was achieved by the reduction of signiicant gradients and even minimal heights of rise and fall, as well as the elimination, 109

Chapter 4

wherever possible, of movable bridges. Tudsbery’s analysis was an early example of a published work that included estimated monetary igures in the documentation of the subject. His description of the essentials of roadway location provided valuable information for engineers attempting to redeine the science of highway planning. he basic theory of economic location was simply but succinctly stated in Tudsbery’s words: he basic requirements of a good track are: (a) to provide as direct a route as possible between the selected terminals, (b) to provide a route on which traic may proceed safely, and at its maximum economic speed, (c) to provide for the safety of all traic using the track, [and] (d) to keep down the cost of the track compatibly with eiciency.16

Despite the hesitance of some American planners to commit themselves to the widespread use of the theory, others wrote strongly in its defense. In 1923, Construction Engineer O.L. Kipp of the Minnesota State Highway Department presented a paper, published in the Engineering News Record, entitled “Economic Study of Highway Design and Location,”17 in which he recommended the careful accounting of operational costs of the vehicles expected to use a road through thorough and cautious analysis of the efects of curvature, grades, and overall distance. Although the study did not consider time savings, it did incorporate operational cost savings, in addition to Iowa Experiment Station conclusions regarding maximum economic grades over varying lengths for commercial and automobile traic. Kipp focused on the importance of reducing gradients, referencing William G. Harger’s principle of expended energy described in the Engineering News Record at the start of the decade. Harger develop[ed] the principle that there is less potential energy wasted on a light grade than on a heavy grade and that the actual saving in operating cost by eliminating a foot of rise and fall over a hill is less important on light grades than on heavy grades.18

Kipp continued by stating that right-of-way and bridge structure costs were of signiicance in highway location considerations. However, his analysis of 110

T h e P u l a s k i S k y way

roadway curvature focused only on the costs involved in the volume of paving material. Much of Kipp’s analysis was devoted to the cost efects of various pavement types, and this portion of the study incorporated only fuel costs for the vehicles using the highway, concluding that the annual savings using a concrete surface on an earth road far exceeded that of using a gravel surface. A review of engineering literature indicates that Kipp’s article was the irst by an American engineer to provide details and actual igures for an economic roadway analysis. Although some other engineering professionals and academics wrote, during the late 1920s, about the need for economic highway design, their published articles were typically very general, and did not provide sample igures for use in an actual analysis. By the mid-1920s Frederick Lavis stood out as the most proliic proponent of the modern economic planning approach. His articles on highway planning appeared regularly in the engineering literature of the 1920s, and published works by other authors of the decade periodically cited Lavis’s writings or his ongoing work in New Jersey. During the late 1920s the major highway project in the United States was that being undertaken on the route between Jersey City and Newark, New Jersey. No other project utilizing economic analysis was cited within the engineering periodical literature of the decade. he Route 1 design, begun by Lavis and continued by Johannesson, helped clarify the methods of economic theory in highway planning, and both men used periodic published accounts of the work on the Route 1 Extension for the education of other engineers. By the start of the 1930s, a great deal of information had been presented in the various national engineering periodicals concerning the design of the Route 1 Corridor and the High-Level Viaduct. Authors continued to pay particular attention to districts of rapidly increasing traic volume, such as that encountered in northern New Jersey, in articles concerning highway design. he continued low of uninterrupted traic was of special concern due to the recognized inancial costs of delays and reduced road capacity. Means of separating traic and of reducing or eliminating traic intersections came under considerable scrutiny. Traic circles, ramps, medians, separation of highway access points, and grade separations began to be designed and constructed with greater frequency. Working engineers from state departments and educational 111

Chapter 4

institutions across the country began publishing information on the studies and work performed within their individual regions. hese design concerns became nearly universal across the nation prior to World War II. On practically every highway location, the engineer [was] . . . forced to evaluate distance. He [might] . . . not ind his value in dollars and cents, but . . . constantly balanc[ed] it against rise and fall, curvature and beneits to the community served. he problem [was] . . . constantly before the engineering profession.19

Highway location, while much spoken of, was still in its infancy in terms of detailed deinition and use in practice. “More and more, highway engineers [were] . . . considering the problems of highway location and construction from the point of view of the man behind the wheel. As a result, one [heard] . . . much of savings in the cost of motor vehicle operation.”20 Some engineers performed practical analyses, while others compiled known information in order that it might be more readily available to working engineers. Articles discussing the cost savings of reduction of roadway distance continued to appear as this area of study became more generally recognized as truly deserving of attention. Despite the increasing use of economic highway theory, many designers of the 1930s still did not feel that economic analysis along the lines of Wellington’s method was appropriate to roadway design. In 1931, in an article for Roads and Streets, Professor of Civil Engineering N.W. Dougherty of the University of Tennessee urged engineers to reevaluate this widespread negative position. He presented an analysis in which he compiled estimated fuel, lubricant, tire, maintenance, depreciation, garage, license, tax, interest, and insurance costs in terms of distance variations in much the same manner that Frederick Lavis had presented them during the previous decade. Indicating that many persons had presented similar analyses to his own during prior years, Dougherty stated that “many of the recent estimates ha[d] placed the savings as high as 10 cents per vehicle-mile” based on distance reduction.21 Dougherty’s model used a weighted average cost combining 90 percent passenger cars at 3.47 cents per mile and 10 percent trucks at 14.86 cents per mile. hus, his study derived a computed 4.43 cents per mile, which he felt should be reduced to account

112

T h e P u l a s k i S k y way

for the likelihood that probably more than 50 percent of passenger travel was undertaken for recreation rather than commerce. he cost per mile for these vehicles was, in his estimation, signiicantly less than that for business use. Despite the fact that Dougherty spoke in favor of economic planning for highway design, his discussion clearly indicated that he believed the generally published estimates of distance–time savings to be signiicantly exaggerated. He stated in his paper that a “recent author” who had based his assumption that a shorter route was preferable on a calculated monetary savings was, in fact, deliberately and unreasonably enlarging the signiicance of the savings by presenting it in terms of savings per year. Dougherty was of the opinion that savings should be provided for smaller periods of time, and that when this was done the savings were oten found to be of insuicient size to justify the expense required for a reduction in highway length. Dougherty proceeded thereater to address the consideration of loss of time. He believed that automobile traic was increasing in great part due to increased recreational use of cars rather than commercial use of trucks. his led him to the conclusion that distance savings determinations were not always warranted. He stated his concerns thus: If a person is driving upon the public highway for pleasure and recreation, he may be driving to spend time and no beneit will accrue to him if time is saved. . . . If a locating engineer has spent good money to decrease grades and shorten alignments, he may defeat the purposes of the party by carrying them to their destination in a short period of time rather than allowing them to spend the whole aternoon.22

Although more energy was being expended on the relatively new scientiic economic method of highway planning during the early 1930s, it is clear from the period literature that Dougherty was not the only engineer of the time who believed that the “modern” method was being endorsed and utilized to the point of frightening planners and the public unnecessarily. Dougherty spoke in favor of caution in using a method that could be useful but could also be misused or misinterpreted. He stated,

113

Chapter 4

there is no doubt that a saving of time for business vehicles has a monetary value but there are cases . . . where the money value is very small. he hauler of freight usually makes a certain number of trips during a working day. If the time saved does not allow an extra trip, it is of very doubtful value. We therefore come to the conclusion that the monetary value placed on business time is high and that the monetary value on pleasure and recreation does not exist. . . . he estimates thus far published are far above any expenditure an engineer is willing to make. When we take into account the small value that time saving may have, we arrive at conclusions more in keeping with our willingness to spend money.23

Clearly, not everyone saw the economic theory of highway location in the same way that leaders such as Lavis and Johannesson perceived it.

he Highway Location heory of Lavis and Johannesson he 1920s was a decade of great debate over and study of the economic analysis of highway design. In the midst of many authors of the time, Frederick Lavis stood out as the most proliic and most determined of the theory’s proponents. His statement of highway location theory was the most clearly and exhaustively deined, and his work was the most frequently referenced and lauded by his peers. Lavis applied his theory of location, as derived and revised from Wellington’s original analysis, to the Route 1 Extension and renamed Route 25 project in New Jersey. He proceeded in his analysis by irst determining those elements comprising the operating costs of the vehicles that he anticipated would be using the completed route. He provided the following information for use in highway planning: he costs of operation of a commercial motor vehicle may be deined into these items: 1. interest on capital invested 2. depreciation 3. insurance 4. drivers’ wages 5. license fee

114

T h e P u l a s k i S k y way

6. gasoline 7. oil 8. tires 9. repairs 10. washing and cleaning 11. battery 12. overhead It will be readily seen that Items 6, 7, 8, 9 and 11 are directly afected by and generally proportionate to the distance traveled by the vehicle, while the others are not.24

Lavis performed personal research and combined it with studies of vehicular costs for the region. hese studies were based on an existing observed vehicular makeup for the area roads comprised of 50 percent heavy trucks, 25 percent medium weight trucks and buses, and 25 percent passenger cars. As a result he estimated that, for the metropolitan New York City area, the combined cost of the ive indicated items equaled approximately 12 cents per mile. Other assumptions were made by Lavis and Pulaski Skyway co-designer Sigvald Johannesson in order that the project might be planned on a realistic cost basis. A speed of 15 to 20 miles per hour was used for traic in a total of four proposed traveled lanes (two in each direction), yielding a 5,455 vehicleper-hour capacity. In addition, “at times of peak traic in one direction it was assumed that the traic on the other two lanes would be at one half capacity.”25 hese determinations combined to yield an estimated 3,600 vehicle-per-hour maximum traic count, or a total estimated annual traic count of approximately 18,360,000 vehicles. With this yearly estimate in hand Lavis had the starting point on which to base all his alignment option analyses. “he economic problem [was] . . . the determination of the efect of the physical characteristics of the route and the cost of modifying or changing them.”26 Heavy trucks were determined to be those with a weight capacity of at least 2 to 2-1/2 tons, and this category of vehicle also included heavy buses. Light trucks included all other smaller commercial vehicles, and automobiles were intended to be primarily those used for recreational travel. he daily cost 115

Chapter 4

of operating a given truck was a combination of some items that were considered while the truck was actually in use and others that were spent regardless of usage. For example, research of area costs indicated that a 3-ton truck cost its owner $3,336 per year in payment for interest, depreciation, insurance, drivers’ wages, licensing fees, and overhead. Divided over an average 300 working days per year this sum yielded an operating cost of 2.3 cents per mile. For smaller commercial and business vehicles these costs added up to $3,120 per year or an equivalent 2.2 cents per mile. Lavis also noted that some vehicles would be using the highway for purposes of pleasure and, therefore, had no true commercial value. However, he also stated that, in his opinion, even those persons who drove pleasure vehicles would be upset by delays and would probably be willing to pay for the capability of avoiding such delays if they could be given the opportunity to do so. herefore, although payment for the option of moving ahead of others or of avoiding slowdowns altogether was not an actual possibility, Lavis believed that it was “proper to assign money value for delays at not less than 1 cent per minute.”27 By combining the average costs of all anticipated vehicles for his proposed route, Lavis could now determine what he termed the “Money Value of One CarMinute.” Assuming that the primary usage on the proposed Route 1 Extension would consist of 75 percent heavy truck, 20 percent light truck and passenger vehicle, and 5 percent pleasure car usage, Lavis’s preliminary analyses yielded an average calculated equivalent operating cost per car-minute of 2.2 cents. With this computation completed, the various factors to be considered for planning the highway location and alignment could be calculated in terms of their comparative expenditures. he goal of the analysis was to determine the amount of money that could be spent, at a proit, in shortening the proposed alignment. Distance was the irst element for which operating costs were computed. Studies were performed using combined time costs and operating costs. Lavis began with available area-wide information indicating costs for a vehicular makeup of 50 percent heavy trucks at a cost of 15 cents per mile, 25 percent light trucks and buses at 10 cents per mile, and 25 percent non-commercial vehicles at 6 cents per mile. From this information he computed an average 11.5 cent cost per mile, or 2.18 cent cost per 1,000 feet of distance traveled. Assuming 18,360,000 vehicles per year, a distance shortening of 100 feet was equivalent to 116

T h e P u l a s k i S k y way

a savings of just over $400,000, which, when capitalized at 6 percent, was equal to $6.67 million in savings. Assuming a rounded twenty-million-vehicle capacity, the equivalent expense for the extension of the route by a single mile cost an estimated $2.4 million per year. herefore, using this capitalized savings as being equal to the amount gained by reducing highway length by one mile, Lavis stated that is was justiiable to spend $48 million for each mile reduction of his route. he same ratio of cost to length could also be applied to other distance variations, and therefore, it was determined to be reasonable to spend $9,000 for the saving of every single foot of length on the highway. he Route 1 Extension being planned was designed using only curves with large radii, and it was assumed to be reasonable that the efects of curvature would be insigniicant in comparison with the other items under study. herefore, the efect of curvature was believed to have no signiicant impact on operational costs, and was not considered by Lavis and his design team. Rise and fall were “considered [in terms of ] how much power/fuel it took to move a vehicle up a grade and at what point it was more cost-eicient to keep a road at a higher grade for a consistent length when a higher grade was needed at a speciic point, than build a road with a series of grades.”28 Earlier experiments performed at Iowa State College had indicated that the work done to raise a weight to a height of one foot was approximately equal to the force required to propel the same weight over a level distance of 50 feet on good pavement. herefore, the cost of producing power alone could be approximated by the combined cost of fuel and oil, which was determined to be 3.22 cents per mile or 0.03 cents per 50 feet. hus a cost equivalent to the value of raising the average automobile to a height of one foot could be determined. On this basis Lavis attempted, in his planning studies, to reduce signiicant variations in grade for the sake of cost-eiciency. On Route 25, grades of between 2 and 3.5 percent were utilized with gradients varying for long distances of continuous rise. Lavis computed that for rises of less than 25 feet in elevation there was no cost efect. However, rises of between 18 and 20 feet, between 20 and 25 feet, and over 25 feet saw the proitable expenditures of $30,600, $61,200, and $91,800, respectively, for each reduction of a single foot of rise. 117

Chapter 4

Another major factor to be considered, which signiicantly impacted main highway traic, was the existence of crossings with other highways at grade. Wellington had not considered the cost of such delays because railroads were generally constructed with little opportunity for local crossings. However, in an area as congested with motor vehicles and motor-vehicle routes as northern New Jersey was, the cost of highway crossings was critical. For computational purposes “it was assumed that the traic conditions at the crossing were such that traic on the main highway could proceed for three minutes and then be interrupted for one minute.”29 he delays caused by at-grade crossings were assumed to occur randomly throughout any given day, and the predetermined average operational cost of 2.2 cents per minute was used for main highway traic. he daily loss on the main route was estimated at 14,532 car-minutes (or single minutes of vehicle travel time) per day, and the loss on the cross street at 9,000 car-minutes, providing a combined 23,532 car-minutes lost per day at one crossing. Using a conservative assumption of 7 million lost car-minutes per day at a conservative 2.0 cent cost per car-minute it was determined that every crossing would result in a loss of more than $140,000 per year. herefore, the capitalized equivalent that could reasonably be spent to eliminate a single at-grade highway crossing was found to be $2.33 million. Lavis explained that delay costs included not only actual loss of vehicular time but also reductions in highway capacity. herefore, capacity calculations had to be factored into the previously derived crossing delay costs. Assuming that delays occurred only at times of maximum volume demand, Lavis estimated a 12.2 percent reduction in the capacity of the main highway. Assuming a construction cost of $22 million for the Route 1 highway, the loss of eiciency was valued at 12 percent of the total, or $2.66 million. herefore, by adding the two elements of crossing delay cost together, one might proitably spend $5 million in order to avoid each crossing. “Of course, if there were a series of such crossings, this sum would not be multiplied by the number of crossings. . . . It was further calculated that, if there were several of such crossings, spaced at approximately equal distances apart and controlled by synchronously operated signals, the amounts which might be spent to avoid them might be assumed”30 to approximately equal $5 million for a single crossing, $5.3 million for two 118

T h e P u l a s k i S k y way

crossings, $5.6 million for three crossings, $5.84 million for four crossings, and $6.08 million for a total of six at-grade crossings. In consideration of the type of structure for use on the Passaic and Hackensack River spans, drawbridges like those existing at the adjacent Lincoln Highway route were considered in terms of vehicular operating costs. Prior to 1928, Lavis had compared tunnels and 35-foot-high movable bridges as his alternates for Section 3 of the proposed Route 25. He conirmed the existing number of openings, and the duration and the times of day at which openings occurred at the adjacent Lincoln Highway drawbridges. hese bridges had vertical clearances of only between 10 and 12 feet, and so the proposed Route 25 spans would require fewer openings than were currently being seen on the original route. Two factors were considered in the analysis of drawbridges and their efect on vehicle costs. hese elements were the same as those used for the atgrade crossing analysis. he actual cost of delays was combined with the cost of decreased overall highway capacity. he delay cost was determined to be equal to the average number of openings per hour multiplied by the number of vehicles using the highway per hour, providing the resultant estimated loss in car-minutes. Delays were estimated at 32,470,000 car-minutes per year. Using a conservative 2.0 cents value of a car-minute and a 6 percent capitalization rate, $10,823,000 could be saved through the elimination of one drawbridge. Eiciency was estimated to be reduced by 6.33 percent for the highway, based on the previously assumed construction cost of $22 million. herefore, an additional $1,392,600 could be saved through the gain in capacity, resulting in a total savings of $12,215,600 for the elimination of a single waterway crossing. Ater Sigvald Johannesson took over the primary work on the alternate studies following the departure of Lavis in 1928, he combined the economic efects of length, rise and fall, curvature, and construction costs for four options under consideration for the Route 25 Connecting Link. hese options consisted of the combined possibilities of two diferent structure types (high and low level). he lengths of the two routes proposed in late 1928, designated Lines A and B, measured 12,253 and 13,003 feet, respectively, regardless of the level at which they crossed the rivers. herefore, the analyses indicated that due to the relationship between the length of line A at either the high or the low level and that of line B, 119

Chapter 4

A was preferable to B because of the approximate 750-foot savings in length. At the 100-foot-length savings cost that had been previously determined by Lavis, this shortened distance resulted in an overall savings of $6,670,000. Johannesson eliminated Line B and utilized only Line A, comparing the diferences between the high-level and the low-level crossings. It was obvious that the high level required greater rises and falls, while the low level resulted in additional operational costs due to delays caused by openings of the drawbridges. Johannesson performed calculations through which the results of the two options might be made equivalent. To accomplish this, the cost of rises and falls was added to the overall low-level cost, and the cost of drawbridge delays was added to the overall high-level construction cost. hese computations indicated that the high-level option was preferable. he resultant construction cost combined with the cost of either delays or rise and fall was $21,474,223 for the low-level tunnel route, and $16,898,410 for the high-level bridge route. he inal studies for alternates for the proposed Connecting Link structure were undertaken by Johannesson in 1929, based upon three slightly diferent determined options. Comparison was made between one hundred 35-foothigh ixed bridges, 35-foot-high movable bridges, and tunnels. Estimated delay costs for the proposed drawbridges were derived from careful consideration of traic density and loss in highway eiciency based upon known igures for the existing Lincoln Highway structures. New lit bridges to be constructed above both the Hackensack and Passaic rivers were estimated to cost a total of $19,289,900. (See Table 4.5.) An estimated $20,454,000 would be required for an independent route for a new highway utilizing tunnels under both rivers. An independent route with high-level ixed bridges at both crossings was estimated at $18,915,000. he costs of all three of these schemes were total construction costs including the cost of real estate takings. (See Tables 4.6 and 4.7.) Annual operating costs were analyzed for all three options, as were the costs of diferences in distance, rise and fall in grade, and bridge operation delays. he total combined construction and operational costs for the three schemes were calculated to be $49,194,054, $48,512,333, and $42,259,167, for the movable bridges, tunnels, and ixed bridges, respectively. (See Table 4.8.) “he results obtained indicated that the most economical type of structure would be a highlevel ixed bridge. his type of construction was then adopted.”31 120

T h e P u l a s k i S k y way

Table 4.5 Cost Estimate for Scheme I – Lit Bridges. (Source: Sigvald Johannesson, Highway Economics [New York: McGraw-Hill Book Company, Inc., 1931]) Linear Feet

Item Description

250

Steel Structure, 24,760 Square Feet

500

Surface Roadway

520

Embankment Between Retaining Walls

250

Steel Structure, 24,760 Square Feet

240

Steel Structure

40 2,490 40

Unit Cost

Total Cost

$15

$371,000

60

30,000

800

416,000

15

371,000

660

158,000

Concrete Abutment

43,000

Concrete Arch Structure

650

1,618,500

Concrete Abutment

43,000

255

Steel Structure

660

168,000

179

Steel Superstructure

600

107,400

739

Lit and Tower Spans, Hackensack River

558

Steel Superstructure

997,000 600

Foundation, Hackensack River 40 2,780 40 2,280

312,000

Concrete Abutment

43,000

Concrete Arch Structure

660

1,834,000

660

1,504,800

Concrete Abutment

43,000

Steel Structure Ramp

435,400

Paving Jacobus Avenue 534

1,500 400

322,800

14,600

Lit and Tower Spans, Passaic River

601,000

Substructure, Passaic River

340,000

Steel Structure

700

1,050,000

Embankment

166

66,400 11,389,900

Lincoln Highway Tunnel, Hackensack River

6,000,000

Lincoln Highway Bridge, Passaic River

1,900,000

Total

$19,289,900

In the face of both promoters and detractors of the theory of economic highway location, Sigvald Johannesson, fresh from the completion of the Route 121

Chapter 4 Table 4.6 Cost Estimate of Scheme II – Tunnels. (Source: Sigvald Johannesson, Highway Economics [New York: McGraw-Hill Book Company, Inc., 1931]) Linear Feet

Item Description

400

Steel Structure

450

Embankment, Retaining Walls

3,000 250

Unit Cost

Ramps Embankment

Total Cost

$660

$264,000

600

270,000

60

180,000

200

50,000

1,000

2,000,000

2,000

Depressed Roadway

3,600

Tunnel, Hackensack River

1,800

6,480,000

2,900

Depressed Roadway

1,000

2,900,000

Bridges Over Roadway

120,000

Ramps 3,200 350

1,000,000

Tunnel, Passaic River

1,800

5,760,000

Depressed Roadway

1,000

350,000

Bridge Over Lincoln Highway 500

40,000

Embankment

120

60,000

Ramp to Avenue P

40,000

Reconstruction, Lincoln Avenue Ramp

40,000

Rock Excavation for Hackensack River Tunnels Total

900,000 $20,454,000

1 Extension project, continued into the 1930s to publish articles for the education of the engineering profession. He stated that “since the study of highway economics [was] . . . quite new, it [was] . . . natural that there should be a number of questions relating to it that [had] . . . not been fully answered as yet.”32 Ever concerned with teaching what he had learned and utilized in his New Jersey superhighway design, he published a paper for the New York publication Civil Engineer33 in 1933, addressing his primary concern relative to highway costs. Delays, according to Johannesson, had to be considered in terms of loss of travel time, which was a combination of vehicular operating costs per vehicle-minute and the value of the time being expended as it was perceived by the persons occupying those vehicles. Johannesson was well aware that not all highways of the type in question were then being designed to carry primarily large trucks. In addition, it was 122

T h e P u l a s k i S k y way

Table 4.7 Cost Estimate for Scheme III – High Bridges. (Source: Sigvald Johannesson, Highway Economics [New York: McGraw-Hill Book Company, Inc., 1931]) Linear Feet 800

Item Description

Unit Cost

Total Cost

Steel Structure

$800

$640,000

1,000

Steel Structure (with Ramp)

2,000

2,000,000

2,000

Steel Structure

1,000

2,000,000

1,500

Steel Structure (Hackensack River Bridge)

2,000

3,000,000

1,200

Steel Structure

1,000

1,200,000

1,000

Steel Structure (with Ramp)

2,000

2,000,000

2,600

Steel Structure

1,200

3,120,000

1,500

Steel Structure (Passaic River Bridge)

2,000

3,000,000

1,400

Steel Structure

800

1,120,000

300

Steel Structure

1,000

300,000

500

Steel Structure

800

400,000

500

Steel Structure

270

135,000

Total

$18,915,000

obvious by the early 1930s that the passenger-vehicle volume, within the northern New Jersey area with which Johannesson was familiar, was increasing even more rapidly than had previously been anticipated. Presenting his argument on the basis of passenger-vehicle, rather than commercial-truck usage, Johannesson ofered computed cost savings as they applied to more general usage than that assumed in the Route 25 design. He also did this because he believed that if operational highway costs were better understood by the general public, people would be more willing to pay for savings that they understood would beneit them directly. Johannesson shited his focus away from the Route 25 project study costs, wherein an approximate 2.0 cents per vehicle-minute had been utilized, in his analysis for savings based upon cars alone, stating that on Route 25 “the proportion of heavy trucks [was] . . . exceptionally high, and for that reason the estimated money value of a vehicle minute naturally [was] . . . materially greater than that for the average highway.”34 In approaching the cost of delays due to slowing and stopping Johannesson presented an equation of his own derivation 123

Chapter 4

through which it was possible to compute the amount of time lost under known conditions. his formula was stated thus:

T = a2p / 2 (d–t) Where: T represented the time lost in vehicle-minutes d is the hydraulic mean depth (inches) p represented the rate of speed a was the time traic was stopped d was the average distance between vehicles t equaled the minimum distance between them

Based upon the little available documentation on vehicle spacing at various speeds he “tentatively concluded that the spacing t, expressed in feet, [might] . . . be determined by the equation: t = 0.025q + 25, where q [was] the rate of speed in feet per minute at which the vehicles [were] . . . moving.”35 Johannesson went on to state that he was aware that some within his profession believed this latter equation to be inaccurate based upon varying diferences in current calculated estimates for vehicular capacities at diferent speeds. For this reason, he noted that suicient documentation did not yet exist from which to determine accurately the relationship between speed and vehicle spacing. In spite of the work still to be done in this area, Johannesson made a strong point in favor of economic roadway planning, stating without hesitation that economic theory was not only applicable but essential to roads engineering. he closing statement in his 1933 article entitled “Cost of Traic Delays” represented his feeling about the signiicance of economic consideration in highway design. He pronounced, the cost of highway improvement necessary to avoid delays may be found by standard engineering methods, and a comparison between this amount and the estimated cost of delays will indicate whether or not the improvement is economically justiied, unless there are other tangible or intangible beneits gained by the improvement. Generally speaking, an improvement is not economically justiied unless its estimated cost is materially less than the estimated cost of the delays eliminated.36

124

T h e P u l a s k i S k y way

Table 4.8 Comparative Estimate – New Jersey State Highway Route 25, Broadway, Jersey City, to Point West of Lincoln Highway, Newark. (Source: Sigvald Johannesson, Highway Economics [New York: McGraw-Hill Book Company, Inc., 1931])

Item

Scheme I

Scheme II

Scheme III

Lit Bridges

Tunnels

High Bridges

Physical Characteristics Length, in Feet

15,050

14,800

14,800

58

131

114

Highway Maintenance

$15,330

$10,240

$52,940

Highway Depreciation

265,679

292,180

214,690

Highway Operation

136,600

120,900

43,900

Total Annual Highway Costs

$416,609

$423,320

$311,530

Motor Vehicles, Distance

$100,000

$0

$0

Motor Vehicles, Rise and Fall

532,440

1,202,580

1,046,520

Motor Vehicles, Delays

798,000

0

0

Total Annual Motor Vehicle Costs

$1,430,400

$1,202,580

$1,046,520

Total Annual Costs

$1,847,049

$1,625,900

$1,358,050

Cost of Right-of-Way

$1,020,000

$960,000

$710

Cost of Construction, Route 25

11,389,900

20,454,000

18,915,000

Cost of Lincoln Highway Tunnel

6,000,000 $18,409,900

$21,414,000

$19,625,000

6,943,487

7,055,333

5,192,167

23,840,667

20,043,000

17,442,000

Total Annual Cost, Capitalized

$30,784,154

$27,098,333

$22,634,167

Total Comparative Costs

$49,194,054

$48,512,333

$42,259,167

Rise and Fall, in Feet Annual Costs

Cost of Construction and Capitalized Annual Costs:

Total Right-of-Way and Construction Annual Highway Cost Capitalized, 6% Annual Motor Vehicle Cost Capitalized, 6%

125

Chapter 4

he critical text in the ield of highway economics was also written by New Jersey designer Sigvald Johannesson and was published in 1931. Entitled Highway Economics,37 this book was generally recognized, throughout the literature of the early 1930s, as the leading published work in the ield. Johannesson was the recognized leader when speaking and working in defense of the theory, having taken the place of his predecessor and former coworker, Frederick Lavis. In spite of the advancement of economic design by 1931, it remained true that many engineers who could see the great signiicance of Wellington’s theory for use in railway location did not see its usefulness for the design of highways. Johannesson continued to attempt to convince his peers that this way of thinking was in error. Much of Johannesson’s text consisted of the basic concepts and items discussed by Frederick Lavis over the years leading up to 1931, and Johannesson’s prime example of the theory was the work on the Route 25 highway. Stating simply that any proposed highway improvement or construction project was economically justiiable if the derived beneits were worth the cost, he reiterated the importance of the analysis of distance, rise and fall, curvature, and elimination of delays. He noted that economic beneits were oten less evident in highway analysis than they had been when utilized in railway design, but the beneits were, nonetheless, similar in nature. In the case of the railroad the same body—the railroad company—spends the money for the improvement and derives the beneit therefrom in the shape of lessened operating expenses; while for the highway project the highway board or commission spends the money and the owners and users of the automobiles and trucks passing over the highway derive beneits from the improvement. But from where comes the money that the highway board uses for the construction . . .? It comes essentially from the taxes paid [by] . . . owners and users, . . . [and] therefore, the same body that pays for the improvement reaps the beneits therefrom.38

Johannesson stated that the tangible beneits of good economic management were reduced operational costs for vehicles and maintenance costs for the highway. In addition, intangible gains such as traveling comfort, improved business facilities, fewer road accidents, and increased real-estate values should 126

T h e P u l a s k i S k y way

not be overlooked. However, although Johannesson saw the need for engineers to weigh these latter beneits, he believed that “unless the intangible beneits [were] . . . of suicient importance to inluence the conditions, the money value of the tangible beneits should be materially in excess of the cost of the project, in order to make it worth while.”39 He explained the reason for the inclusion of each individual cost item in his and Lavis’s analysis, and indicated that the means of placing all items on a comparative basis was the reason for introducing some items as capitalized costs. For example, some items, such as construction cost, were single, non-recurrent costs, while other items of expense, such as those associated with maintenance, occurred, at varying levels, every year. “In order to place these annual expenditures on a basis which ma[de] it possible to compare them with the nonrecurring charges, it [was] . . . convenient to introduce the maintenance costs in the form of Capitalized Cost of Maintenance.”40 herefore, the yearly maintenance expense was considered to be equivalent to annual interest on a loan. he capitalized cost was, thereby, equal to that amount of money required for yearly maintenance costs at a predetermined interest rate. he “Highway Economics” text provided example problems throughout for the computation of the operational costs from various sources of increased expense. he assumptions were made, for these sample computations, that twothirds of passenger cars on the nation’s highways were traveling for noncommercial purposes, that 80 percent of trucks were of relatively light capacity (two tons or less), and that nationwide norms could be utilized for commercial drivers’ wages, vehicle maintenance and fuel costs, and number of working days per year. Johannesson used these determinations to present his case for the efective economic analysis of highway layout. He recognized the existence of criticism on the part of other engineers but continued in defense of his process. He stated within the book, it has been suggested by some that no money value can be assigned to the time lost on the road where it amounts to a few minutes only; in other words, that it is of no economic importance that additional minutes, or even additional half hours or more, are spent on the road, because if this time were not spent in traveling, the motor vehicle and the driver might be idling away the same time at

127

Chapter 4

the home garage. For commercial vehicles, this does not appear to be the proper viewpoint. . . . A commercial vehicle is, in fact, a commercial plant of which the cost of operation must be distributed according to the time spent on various jobs. It may be said that such distribution is purely a bookkeeping matter. While this may be true, it is true also that a deinite basis for establishing the value of a vehicle minute is obtained in this manner.41

Clearly, Johannesson said, money might not be lost if a delay occurred at the end of a day on the way back to one’s garage. However, actual money was truly spent and lost if the delay encountered resulted in the loss of an entire commercial trip or the wasting of a business journey. he book incorporated the methods previously deined in the Route 25 analysis for the calculation of the costs of added distance, of ascending or descending gradients, of maneuvering around curves, or of being slowed or stopped due to elements of delay. Johannesson also provided extensive detail about the analysis of vehicle speeds, vehicle spacing, and expended work, and included equations deining every element of his delay cost analysis. He provided his reader with the means for determining how many vehicles would be stopped and delayed during a given traic interruption, the length of time of the stoppage, the number of vehicles both originally stopped and later delayed by the earlier disruption, and the total number of vehicle-minutes lost by all vehicles involved in a given delay occurrence. He observed that “the total time lost is proportional to the square of the time of the stoppage, [and that] it is apparent that less time will be lost if the time of stoppage is decreased, even though the number of stoppages may have to be correspondingly increased.”42 he terms “traic density” and “maximum traic density” were also clariied by Johannesson. he designation of highway capacity was derived from density deinitions as being the greatest number of vehicles traveling past any point along a route within a set period of time. Although Johannesson’s capacity deinition difered from that which had recently been issued by the Highway Research Board, the engineer believed that his statement more accurately relected the fact that vehicles on a given highway were rarely spaced evenly or traveling at equal rates of speed. He stated that, although high-density urban traic was oten addressed through the construction of high-level routes, which oten resulted in 128

T h e P u l a s k i S k y way

increased construction costs, this type of construction could still be used to economic advantage based upon the needs of the region. In other words, although the structural costs for erection of a high-level highway were generally substantially greater than those for a surface route, the former alternate could be proven to be economically viable for heavy-volume routes passing through “built-up sections of cities and towns, intersecting many streets with considerable volume of traic, [with] the traic interference, [and] with all the crossings at grade.”43 he elimination of these sources of delay could, in heavily populated areas, prove to ultimately save more money than the reduction of initial construction costs. He “proved that in many cases, initial investments in road infrastructure would result in greater long term economic gain, if the design and location of the project enhanced traic low and reduced transportation costs.”44 Having said this, Johannesson presented, by way of illustration, the detailed analysis used in the design of the Pulaski Skyway. Few of the documents published in the 1920s and 1930s on economic highway theory referred to actual physical examples of its use, with the exception that the New Jersey superhighway was typically referenced as the prime example of the economic-design process. In response to Fred Lavis’s article “Highways as Elements of Transportation,”45 published by the American Society of Civil Engineers (ASCE) in 1931, comments were received from nine recognized and respected members of the society. W.W. Crosby indicated that in the utilization “of the principles suggested by Mr. Lavis . . . perhaps no single example of remarkable magnitude [could] . . . be cited.”46 Similarly, W.L. Webb, in discussing the same paper, stated that “the construction of Route 25 in New Jersey [was] . . . one of the irst of many . . . projects which [would] . . . undoubtedly arise in the solution of the fast growing motor-transport industry.”47 he New Jersey Route 25 highway construction provided, in the words of D.P. Krynine, “an introduction into the transportation system of a new kind of link that is something between ‘highway’ and ‘railway.’ his new member of the transportation family may be called a ‘superhighway’.”48 Engineer E.F. Wendt called Lavis a leader in the utilization of railway economics for highway planning, and D.M. Baker commended Lavis in the use of the economic theory of highways, which was still in the process of being fully deined. As noted in these responses to Fred Lavis’s ASCE article, Lavis’s work on the Route 25 highway 129

Chapter 4

was recognized by his peers as pioneering, as well as signiicant to future highway development. Routes 1&9 was designed in the era when predictive formulae for assessing future traic load was emerging and [was] . . . the irst highway in New Jersey to make use of this technology. Admittedly, these formulae were based on wellestablished railway engineering principles, but, nonetheless, this appears to be the irst time that these basic concepts were applied to a highway. In terms of its design it is signiicant as an early application of highway engineering principles in the sense that it was one of the irst roads in New Jersey to be speciically designed for high speed, automobile and heavy commercial truck use.49

he combined work of Lavis and Johannesson, during and ater the design of the Route 1 Extension, provides the clearest and most extensive deinition of the signiicance and means of utilizing economic theory in the layout of highways. In addition, both their published writings and the resultant physical highway project itself were seen by all period authors and publications reviewed in the preparation of this document as the most valuable examples of the theory put into practice. he Corridor and the Skyway, and the analysis undertaken in their planning, provided the basis for the widespread use of economic design of engineering works that was to follow. Additionally, the design and the structures resulted in signiicant time and, therefore, inancial savings for the operators of vehicles utilizing the highway during the years following completion.

Time Savings on the Pulaski Skyway When the Pulaski Skyway was nearing completion, the state of New Jersey requested that the Federal Bureau of Public Roads (BPR) conduct a traic study to compare the efects of the new route to conditions existing prior to its opening. he highway department had expended approximately $5.2 million per mile for the viaduct, and the most critical saving analyzed during the design phase of the project had been that of travel time. herefore, it was of great importance to evaluate the ultimate efectiveness of the project. In response to this request, “as 130

T h e P u l a s k i S k y way

part of a twelve month traic census . . . [the] Bureau of Public Roads include[d] a time study of delays, between Jersey City and Newark, before and ater the viaduct opened.”50 During the years 1932 and 1933 traic low and volume studies were performed on and in the vicinity of the viaduct by the BPR’s Division of Highway Transport. Although the results were compiled in terms of vehicleminutes saved, “no attempt . . . [was] made to value vehicle time saved in view of the general lack of agreement among engineers and economists” of the time.51 Due to the variation in drawbridge openings on the original route, it was of importance that the study incorporate wide variations in study time. he traic analysis of the old route was performed in late September and early October of 1932 over staggered eight-hour periods during each project day. Traic counts were made on the Skyway itself during May of 1933, six months ater the highway opening, “in order to allow traic to adjust itself to driving conditions.”52 he studies undertaken compared vehicular time spent in crossing the area via the old route (4.2 miles in length) versus that taken along the new, shorter route (approximately 3.7 miles long) between the same end stations. (See Figure 4.4.) he new highway permitted vehicles to enter and exit at limited locations. Figure 4.4. Map showing old route (stippled) and new viaduct route. (Lawrence S. Tuttle, “A Time Study of Traic Flow on the New Jersey High-Level Viaduct,” Public Roads 14, no. 2 [1934]) 131

Chapter 4

However, many persons using the old route made use of the existing underpassing roads and did not utilize the main highway for the entirety of its length. A method of recording vehicle license-plate numbers at each end of the route and at various observation stations along the way was developed. his enabled workers to count vehicles traveling the entire route as well as those leaving at various points. Each observation point consisted of two separate stations to permit the recording of traic information in both directions. Vehicles were noted and recorded between the set stations, and classiied by size and type. Punch cards were used to transfer information, which could then be tabulated mechanically. Weighted average trip times were established by compiling estimated hourly volumes for each vehicular class in each traic direction and combining it with existing information on bridge openings on the original route. Area traic counts for 1932, provided by the state, noted that increases in regional traic during the decade prior to the Skyway’s opening had averaged well over 300 percent. Port Authority traic records for the Holland Tunnel were analyzed and compared with equivalent 24-hour counts for the old Hackensack River span. he combined vehicle counts were then corrected for known variations due to seasonal driving habits. New Jersey records indicated a signiicant decrease, between 1926 and 1932, of drawbridge openings on the old highway. In spite of this, BPR and New Jersey state engineers still believed that “there [was] . . . every indication of an increasing number of openings in future years, and consequently more frequent interruptions to traic”53 and, therefore, that as traic volume increased and delays on the old route remained unabated, any time savings seen in 1933 would continue to be realized in future years. he trip times and traic counts for the longer route were divided into hourly periods to provide a realistic presentation of the variations seen both in overall volume and in number and duration of drawbridge openings. “he average trip time for each type of vehicle was obtained by multiplying trip time for each hour and for each direction by the number of vehicles for each corresponding hour and direction, totaling these products and dividing by the total number of vehicles of the given type for the day.”54 Vehicular volumes at various sections of the original route were plotted against the total computed trip times. hese mathematical 132

T h e P u l a s k i S k y way

curves represented the comparative volume versus time, dependent upon the number of bridge openings per hour. Once the curves for the original route were deined, “similar curves were itted to the data for trip time on the viaduct.”55 Although this method of analysis seems simplistic by today’s standards, it provided information of suicient accuracy to give the persons performing the analysis a good mathematical approximation of the encountered trip times. he literature of the time does not indicate when, in the course of the study, the method of analysis to be used was deined or developed. In the early 1930s little had yet been accomplished in the area of analyzing the movement of vehicles. However, during 1932 another vehicle-movement study was conducted to analyze traic low and delays at the intersection of Constitution Avenue and Seventeenth Street in Washington, D.C. For this study, which covered a length of only several hundred feet (unlike the approximate four-mile-long Skyway route), the required degree of accuracy was obtained through the use of a timerecording machine activated by telegraph connections to various points along the route. his apparatus was new to traic study and “probably no more accurate or inclusive information could [have been] . . . obtained with methods . . . available to traic investigators” at the time.56 he amount of precision obtained (to the nearest second) through the use of such machinery was needed for the Washington study because the roadway distance under consideration was relatively small. On the other hand, the manual documentation used for the Pulaski Skyway, wherein recording was made to the nearest minute, was of adequate precision for the purposes of the study. Both methods were recognized in 1934 by the BPR as being appropriate given their particular ranges of study. In fact, both methods were also reported, at the time, to be the most advanced means of performing traic-movement analysis then in existence in the country. he New Jersey time study results indicated a signiicant time savings due to the construction of the Route 1 Extension and Pulaski Skyway. he great diference in the volume of traic traveling between Tonnelle Circle and the west end of the viaduct, before and ater the opening of the viaduct, . . . [was] considered as a measure of the traic . . . diverted to the viaduct, and which unquestionably beneit[ed] by the saving in time between these points.57

133

Chapter 4

Additionally, a large amount of area traic bound for the newly opening George Washington Bridge had previously utilized roads within the municipalities north of the Tonnelle Avenue circle in Jersey City. he study indicated that time was saved in the area due to the reduced volume of traic on the old route resulting from the diversion of much of this volume onto the new Skyway. It was observed that the number of cars at the old river crossings exceeded the volume of a decade before by less than 60 percent, despite the greater than 300 percent overall regional increase in volume seen during the same decade under consideration. In addition, the vehicles using the new route were able to cross the region more rapidly than had been possible in earlier years. An estimated 15- to 45-minute savings per trip had been assumed in Johannesson’s original studies. During the year of the highway opening it was observed that the trip, which had previously taken approximately 55 minutes to complete, took only 20 minutes. Even at an estimated savings of one cent per minute per car, which was conservative Figure 4.5. General view of the Pulaski Skyway between the Passaic and Hackensack rivers. Jack Boucher, photographer, 1978. (Library of Congress, Prints and Photographs Division, Historic American Engineering Record, HAER NJ, 9-JERCI-10-4)

134

T h e P u l a s k i S k y way

Table 4.9 Minimum Estimate of Time Saved Per Year by Vehicles which Previously Traveled on the Ground-Level Route Between Tonnelle Circle and the West End of the Viaduct. Vehicle-Minutes Saved

Daily Vehicles

Minutes Saved Per Trip

Per Day

Number of Days

Per Year

13,695

6.3

86,278

254

21,915,000

Light Trucks

1,249

7.5

9,368

254

2,379,000

Heavy Trucks

2,316

7.4

9,738

254

2,474,000

Weekday: Cars

Saturday: Cars

21,276

6.4

136,166

52

7,081,000

Light Trucks

957

7.8

7,465

52

388,000

Heavy Trucks

711

7.7

5,475

52

285,000

Sundays/Holidays:

29,800

8.0

238,400

59

14,065,000

Light Trucks

Cars

670

7.3

4,891

59

289,000

Heavy Trucks

410

10.5

4,305

59

254,000

Total Vehicle-Minutes Saved Per Year Passenger Cars

Light Trucks

Heavy Trucks

Total

On the Viaduct

43,061,000

3,056,000

3,013,000

49,130,000

On the Ground-Level Route

4,346,000

465,000

731,000

5,542,000

Total

47,407,000

3,521,000

3,744,000

54,672,000

(Source: Lawrence S. Tuttle, “A Time Study of Traic Flow on the New Jersey High-Level Viaduct,” Public Roads 14, no. 12 [1934])

when compared with Lavis and Johannesson’s original approximation of two cents, it could be observed, in 1932, that $6.36 million per year would be saved in vehicle costs. As two means of time savings existed, two separate savings estimates were computed. he irst consisted only of the diference in traic time for those vehicles traveling between the Newark (west) end of the Skyway and the circle at the (east end) intersection of Route 25 with Tonnelle Avenue. he other consisted of all vehicles diverted from the Lincoln Highway Hackensack River Bridge to the Route 25 crossing. hese estimates were designated the minimum and maximum estimates, respectively. (See Tables 4.9 and 4.10.) New Jersey’s inal Skyway construction cost had been $19,300,000 and Johannesson had estimated a 6 percent 135

Chapter 4

rate of capitalization in his analysis. Using these igures, the BPR personnel determined that “in order to justify the construction of the viaduct on the basis of the . . . minimum and maximum estimates a vehicle-minute [had to] . . . be valued at 2.12 or 1.75 cents.”58 he viaduct cost was, therefore, justiied, as Lavis’s original estimated average 2.2 cent cost per vehicle-minute exceeded both of these igures. he BPR study considered only savings due to delays and trip time, and no accounting was made of the recognized likelihood that traic that had previously avoided the old route due to congestion and delays was now using the new route. However, the volume of traic on the original route had deinitely been substantially reduced over the course of less than one year, permitting “greater freedom of movement for local traic, particularly at the previously congested intersections.”59 It was also stated within the study results that there was “every reason to believe that the traic on the viaduct [would] . . . increase during the next few years to a volume which could not have been adequately served by the old route.”60 Ultimately the bureau’s study indicated both a clear savings in time on the old route due to traic diverted to the express highway and “a considerable time saving to those vehicles which . . . use[d] the viaduct.”61 he total computed time savings on both routes was estimated at 54,672,000 vehicleminutes per year. Regardless of how one might value this total, the time savings was undeniable.

Conclusion As public oicials and American citizens began to focus on road conditions during the closing decades of the nineteenth century, New Jersey led the nation in the establishment of funding for road construction and the development of a state highway planning agency and enacted the nation’s irst State Aid Road Act in 1891. By the early decades of the twentieth century, scientiic highway planning was seen by many as a means of dealing with the overwhelming transportation diiculties plaguing the densely populated urban districts of the country, such as those in northern New Jersey. State engineers began studying the improvement of the existing road network to accommodate the already taxing volumes of vehicular traic, focusing primarily on the development of a system 136

T h e P u l a s k i S k y way

Table 4.10 Maximum Estimate of Time Saved Per Year by Total Vehicles Diverted to the Viaduct From Communipaw Avenue Vehicle-Minutes Saved Daily Vehicles

Minutes Saved Per Trip

Per Day

Daily Vehicles

Saved Per Trip

17,680

6.3

111,384

254

28,292,000

Light Trucks

1,804

7.5

13,530

254

3,437,000

Heavy Trucks

1,306

7.4

9,738

254

2,473,000

21,880

6.4

140,032

52

7,282,000

Light Trucks

1,209

7.8

9,430

52

490,000

Heavy Trucks

711

7.7

5,475

52

285,000

Weekday: Cars

Saturday: Cars

Sundays/Holidays:

37,130

8.0

297,040

59

17,525,000

Light Trucks

Cars

1,025

7.3

7,482

59

441,000

Heavy Trucks

545

10.5

5,722

59

338,000

Total Vehicle-Minutes Saved Per Year Passenger Cars

Light Trucks

Heavy Trucks

Total

On the Viaduct

53,099,000

4,368,000

3,096,000

60,563,000

On the Ground-Level Route

4,346,000

465,000

731,000

5,542,000

Total

57,445,000

4,833,000

3,827,000

66,105,000

(Source: “A Time Study of Traic Flow on the New Jersey High-Level Viaduct” by Lawrence S. Tuttle, Public Roads, September 1934, Volume 14, Number 12)

incorporating the construction of commercial routes and the separation of through and local traic. he state’s earliest comprehensive highway system was a milestone in not only regional, but national highway planning. It combined the use of existing routes with the establishment of new corridors and alignments, and represented pioneering American work in the use of future projections of road needs. During the early 1920s American writers of engineering literature had begun presenting articles recommending a more modern, scientiic, economic approach to highway planning, although few engineers of the time knew how to put these new ideas into practice. A review of the engineering literature of 137

Chapter 4

the period shows that Frederick Lavis wrote regularly about the means of performing economic road analysis and that his peers in the engineering community recognized him to be the expert in the ield. When examples of procedures and project documentation were referenced by others, Lavis and the Route 1 (Pulaski Skyway) project were widely noted. Details of the design and construction of what was aterward called the Pulaski Skyway were featured in many professional articles. It was heralded by engineers as a leading public works project. Even more important, progress of the construction of the Skyway was widely noted in the popular press until its completion in 1932. Although largely forgotten today, the completion of the project can be claimed as the beginning of the superhighway with dual lanes and ramps for merging and existing traic. To many observers, it was the German autobahn system of 1935 that ranks as the beginning of the modern dual highway. Others would cite the Queen Elizabeth Way opened by King George VI and Queen Elizabeth in 1937 as the oldest example of a superhighway in North America, connecting as it did and now still does the Niagara Falls area with Toronto, Canada. Another North American candidate is the Pennsylvania Turnpike, inished in 1940. here is little doubt that the 1932 Pulaski Skyway was the beginning of this widely accepted modern superhighway. he Skyway is the surviving symbol of a signiicant step in highway planning in its transition from railway engineering practice. During its construction the Pulaski Skyway was termed the greatest highway project in the United States. When it was opened to traic this pioneer elevated highway was labeled, by the engineering community, as the outstanding achievement in its history. he Route 1 Extension project brought the use of economic analysis in highway design to the forefront of American planning practice by the time of its completion. he Pulaski Skyway was in 1932, and remains today, a prominent, unavoidable testament to the expansion of industry, the dramatic advances in transportation technology, and the inventiveness of engineers that dominated New Jersey and the remainder of the United States during the irst half of the twentieth century. It is, therefore, entirely itting that the importance of this skyway be reestablished in the minds of not only the profession but of the public in general and particularly those living in the northeast corridor of the United States. 138

T h e P u l a s k i S k y way

Chapter 4 Notes 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21

KFS Cultural Resources Group, Overview History of New Jersey Highway Development, 1997, 82. A. M. Wellington, Economic heory of the Location of Railways (New York: John Wiley & Sons and the Engineering News, 1902), 15. Wellington, Economic heory, 8. TAMS Consultants, Historical Narrative and Assessment of Signiicance and Integrity: Routes U.S. 1&9 Corridor Historic Engineering Survey (1991), 18. Wellington, Economic heory, 185. Ibid., 327. Ibid., 713. Ibid., 660. C. R. homas, “Notes on Economic Highway Location,” Highway Engineer and Contractor 35 (Sept 1929): 44. Ibid. J.A.L. Waddell, Economics of Bridgework (New York: John Wiley & Sons, Inc., 1921). Wellington, Economic heory, 116. J.A.L. Waddell, Bridge Engineering (New York: John Wiley & Sons, Inc., 1916). H. L. Brightman, “he Location of the Modern Highway,” Roads and Streets 65 (2 June 1926): 326. H. T. Tudsbery, “Economics of Highway Engineering,” Engineering 120, no. 3114 (4 Sept 1925): 287–88, and 120, no. 3115 (11 Sept 1925): 339–340. Ibid., 340. H. T. Tudsbery, “Economic Study of Highway Design and Location,” Engineering News Record 90 (19 April 1923). O. L. Kipp, “Economic Study of Highway Design and Location,” Engineering News Record 90 (19 April 1923): 699. N. W. Dougherty, “Evaluation of Distance and Time in Highway Location,” Roads and Streets 71(3 June 1931): 231. W. W. Zass, “Rudimentary Economics of Highway Design,” Civil Engineer (New York) 2 (Oct 1932): 619. Dougherty, “Evaluation of Distance,” 232.

139

Chapter 4

22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49

Ibid. Ibid., 232–233. Frederick Lavis, “Grade Crossings: he Money Value of a Car Minute,” he Annals of the American Academy 133 (Sept 1927): 173. Frederick Lavis, “Highways as Elements of Transportation,” Transactions of the American Society of Civil Engineers 95, no. 1783 (1931): 1024. Lavis, “Grade Crossings,” 174. Ibid., 177. TAMS Consultants, Historical Narrative, 19. Lavis, “Grade Crossings,” 173. Ibid., 174. H. W. Hudson, “he New Jersey High-Level Viaduct,” Civil Engineer (New York) 3 (March 1933): 149. Sigvald Johannesson, “Cost of Traic Delays,” Civil Engineer (New York) 3 (March 1933): 149. Ibid. Ibid., 151. Ibid. Ibid. Sigvald Johannesson, Highway Economics (New York: McGraw-Hill Book Company, Inc., 1931). Ibid., 4. Ibid., 5. Ibid., 31. Ibid., 53. Ibid., 86. Ibid., 113. KFS Cultural Resources Group, Overview History, 68–69. Lavis, “Highways as Elements of Transportation.” Ibid., 1041. Ibid., 1043. Ibid., 1047. TAMS Consultants, Historical Narrative, 42.

140

50 51 52 53 54 55 56 57 58 59 60 61

FHWA By Day–A Look at the History of the Federal Highway Administration, www.hwa.dot.gov/byday/hbd1124.htm. Lawrence S. Tuttle, “A Time Study of Traic Flow on the New Jersey High-Level Viaduct,” Public Roads 14, no. 12 (Feb 1934): 223. Ibid., 225. Ibid., 228. Ibid., 226. Ibid. E. H. Holmes, “he Efect of Control Methods on Traic Flow,” Public Roads 14, no. 12 (Feb 1934): 242. Ibid., 228. Ibid., 231. Ibid. Ibid. Ibid., 229.

141

5

James Finley and the Origins of the Modern Suspension Bridge “Look at the nations, and see! Be astonished! Be astounded! For a work is being done in your days that you would not believe if you were told.” –Habakkuk 1:5

S

ince the publication of the multi-volume History of Technology in Britain and its subset, History of Engineering, the ield has grown steadily as witnessed by numerous publications, the establishment of programs in the history of technology, and the formation of professional organizations.1 Equally important, new societies have been formed in the United States and elsewhere to explore the role of science and technology in society, especially during the Industrial Revolution. Born of World War II, the discipline of industrial archaeology arose to take leadership in the new discipline of documenting individual sites with measured drawings, archival photographs, and written histories of each site or district under investigation.2 During the growth of industrial archaeology as a discipline, the documentation of historic bridges has been a central focus in the ield work under the aegis of the Historic American Engineering Record of the United States Park Service. Many others have also participated in documenting our industrial past. An impressive collection of documents on bridges has been established in a special collection at the Library of Congress in Washington. Numerous bridges are listed on the National Register of Historic Places and in collections at the local and state levels. Ater three and more decades of recording, as of the irst decade of the twenty-irst century little has been done in the way of producing 142

Th e O r i g i n s o f t h e M o d e r n S u s p e n s i o n B r i d g e

contextual history. In a sense we will see that this may be construed as an example of antiquarianism without serious evaluation by historians of this rich facet of history. As a result there is only a slim possibility of this resource being integrated in a meaningful manner in American history. hus, such work must be developed, perhaps most conveniently by bridge type. he move away from essentially an antiquarian approach to bridge history is exempliied by the history of the beloved American covered timber bridge. Without delving into the origins of this bridge type, it has achieved a legendary status with the public in North America because it is cherished as a genuine American type. In fact, nearly all of the rapidly developing technical innovations of the nineteenth century can trace their origins to Europe, except possibly the timber covered bridge so common in nineteenth-century North America. An illustrated large-format publication from Leipzig, Germany, in 1735 shows Figure 5.1. he earliest depiction of a covered bridge is this one from Carl Christian Schramm’s 1735 work, Historischer Schauplatz, of a design for a covered bridge in Zwickau.

143

Chapter 5

a complete covered bridge framed with queen post trusses, a level deck to support wheeled vehicles, and all protected by a wooden roof and siding.3 hus German-language areas of Europe can take pride in developing the fully modern covered bridge, which appeared long before the American Revolution. Despite eighteenth-century details, little is known about the origins of the European timber covered bridge, except to say that the timber roof trusses for building date from Roman times. Hidden away out of sight, many historic roof structures are extant. Such architectural features were clearly the origins for the German timber bridge. It is important to note that in the hands of American bridge builders and engineers, the traditional timber truss bridge appeared by the middle of the nineteenth century as an all-iron, and later steel, truss bridge. his was a major transformation and can be credited to North American engineers at the time. During the nineteenth century American engineers can be credited with three signiicant structural developments. In addition, the all-metal truss enabled really long-span rail and road bridges to be built across major rivers in the United States and Canada.4 Skyscrapers emerged like a phoenix from the ashes of the great Chicago ire. Equally important, major suspension bridges irst appeared in America, which led to the erection of towering bridges, which attracted a loyal following from the public. As an example, the Wheeling Suspension Bridge of 1849, exceeding a clear span of 1,000 feet, and by 1883 the great Brooklyn Bridge, perhaps the world’s most famous bridge, was inished with a main span of 1,595 feet. American ascendancy in long-span suspension bridges dating from the Wheeling and Brooklyn examples lasted until a new generation of bridges was built overseas following World War II. A leading consultant in the late nineteenth century and early twentieth century, J.A.L. Waddell, was a forceful proponent of the modern suspension bridge, but in his book he shows a troop of monkeys linked hand and foot to form a suspension bridge.5 While clearly not factual, the legendary illustration is indicative of the primitive use of rope and other organic materials by prehistoric man on a world-wide basis. Many of these aboriginal bridges still grace the landscape in alpine areas. he origin of the long-span catenary bridge rests in the separate and distinct alpine regions in the world, namely the Andean mountains and the Himalayan 144

Th e O r i g i n s o f t h e M o d e r n S u s p e n s i o n B r i d g e

mountain range lying between the Indian subcontinent and Tibet. In the case of the Peruvian technology, many signiicant long-span bridges were built by the Incas to serve their extensive road system. hey were fashioned of vines and ropes by a stone-age culture devoid of metal components. Because of the fragility of the structure, the bridges were inspected regularly and cables replaced as the organic materials disintegrated. Housing at each end of the bridge was erected on top of the abutments. Not only did this protect the ends of the cables, but possibly served a religious purpose. In the charge of a priest, these structures protected the most vulnerable parts of the bridge. Because of the inherent stretching of the cables, these elements were wound around primitive capstans that allowed the cable sag to be adjusted, unlike modern leveldeck suspension bridges. hese Inca bridges had their deck laid directly upon the cables, as did the bridges of the Chinese. hus they could be designated as catenary bridges. Nevertheless, and despite the coniguration of the catenary bridge, pack animals were brought safely across such bridges. Unstifened decks without stifening trusses were very lexible under wind forces. An abbreviated Figure 5.2. he Penipe Bridge over the Chambo River in Latin America. his is a typical Inca-style rope bridge. Humboldt’s drawing was used by Charles Ellet in his Wheeling Suspension Bridge illustration. 145

Chapter 5

story tells of an army moving from Chile that encountered some diiculties when the soldiers attempted to pass a cannon on such a bridge. he artillery piece was ensnared on the deck and was in danger of depositing its load in the river. he story ends by noting that the artillery piece was completely disassembled and carried across piece by piece. A popular novel of such a bridge entitled he Bridge of San Luis Rey provides a graphic description of the diiculties of crossing an Inca bridge.6 In a well-drated account of Inca life, other primitive methods of crossing a gap are described, including a single-cable system that could transport both men and beasts in harnesses across dizzying heights. One wonders how the horses were induced to jump of the clif in a harness and be carried across a spectacular gorge. Many similar cableways have existed but few remain. Arial tramways called Téléphériques in French are still used on construction sites. Despite technical advances made by the Incas without using metal, development of the suspension bridge was severely curtailed. Especially important, this kind of technology played a negligible role in America. Inca suspension-bridge technology did not spread to other parts of Spanish America, although there is fragmentary evidence that such structures appeared in Maya societies in Meso Figure 5.3. he Chaksam-chori Lamasery Bridge, ca. 1420, across the Brahmaputra River. A very early example of a level deck suspension bridge.

146

Th e O r i g i n s o f t h e M o d e r n S u s p e n s i o n B r i d g e

America. Despite information available in Europe and America the lack of iron technology relegated the Inca technology to an antiquarian interest. It was not instrumental in the subsequent development of the modern suspension bridge. Nevertheless, later bridge designers may have been impressed by the Peruvian achievements in promoting modern French-style wire suspension bridges. Charles Ellet Jr. includes a view of the Penipe Bridge on his well-known drawing of the Wheeling Suspension Bridge of 1849. his leaves one to conclude that he was inspired by Inca achievements but this was not the case. Ellet relegated the work of James Finley together with the Peruvian achievements to the primitive stage of suspension-bridge development, and not related to the French technology of the time, which used both chains and wire. Ellet was a leading proponent of the new technology, having studied in France and examined a number of French pioneering suspension bridges. He was also clearly inspired by Navier’s published work on the mathematical approach to the subject.7 It was Navier who irst suggested that the suspension principle could be used for aqueducts. his idea was taken up by John Roebling in his aqueducts in Pittsburgh and on the Delaware and Hudson Canal in New York State. While there is no established link between the Andean suspension bridges and associated bridge-building technology, Needham lays out a detailed suggestion that the high culture of Meso America owed a signiicant debt to Chinese inluence resulting from Chinese voyages of discovery to the New World. his thesis has not been validated. Moving from the isolation of the high Andes to an equally isolated region of the Himalayas, it has been established that SinoTibetan bridge building enjoyed a direct inluence on later European engineering. With the origin of both suspension bridge centers in high alpine regions it is not surprising that the origins of the suspension bridge demonstrate a rather clouded historic situation. From the crude single-cable structures to the emergence of a fully developed catenary bridge, which was known to European travelers, very similar inventions occurred on the Indian side of the Himalayas and formed an important link in Buddhist pilgrimage trails, replete with suspension bridges as needed to link Buddhist centers in India and Tibet. At the most elementary level a single-cable bridge was used on the slope so that by suspending a ring lubricated with yak butter it was possible to pass easily across a chasm. A second cable sloped in the opposite direction allowed 147

Chapter 5

a person, cargo, or even an animal to be transported across a chasm. Although dangerous, it was used extensively. A number of these devices continued to be in operation in the twentieth century, and in selected remote areas are still operating in the twenty-irst century. By placing a pair of cables side by side, a light wooden deck could be installed, producing what has come to be known as a catenary suspension bridge. With this device, pedestrians, livestock, and pack animals could be induced to cross the bridge. Such structures were very lexible, with only rudimentary handrails and cables. It takes little imagination to envisage sets of cables supporting a level deck by means of secondary cables called suspenders passing from the chain cables to the loor beams of the deck, increasing in length as the structure approaches the towers. Although the beginning of the use of suspenders is not known, suspension bridges using exclusively organic cables date from well before the Christian era. During this era the cables were fashioned from vines and ropes, necessitating constant maintenance and adjustment of the cables by a series of capstans, used both in the Andean and Himalayas region. Bridge keepers were on duty permanently and charged with maintaining such bridges in good condition. Iron smelting and production of cast iron represent an early and outstanding achievement by the Chinese. his iron material with its high carbon content displays superior compression properties and, as its name suggests, can be cast in intricate forms in suitable moulds. he presence of carbon in the material greatly increases the strength of the iron at the molecular level. High concentrations of carbon also enhance the molten material’s ability to be cast. With this ancient technology one could consider this material to be “grey iron,” with notable compressive strength and lacking tensile capabilities. Nonetheless, cast iron proved to be a judicious choice for bearings supporting chain cables and in other uses. It was particularly used in the nineteenth century for a wide variety of structures, including cast-iron bridges. he other form of iron that displays both tenacity and compressive strength laid the foundation for Chinese iron suspension-bridge chain cables and is referred to as wrought iron, meaning that the carbon content of the iron furnace product must be reduced by working, which would involve hammering and later rolling to reduce the carbon content to nearly zero, giving an outstanding 148

Th e O r i g i n s o f t h e M o d e r n S u s p e n s i o n B r i d g e

structural material. Traditionally, wrought iron was produced in small quantities by a blacksmith on a Catalan forge, which was developed later in Europe. Wrought iron as a malleable metal could be worked into various shapes. he traditional horseshoe is a prominent example. Equally important, rods could be produced, but in limited lengths, since the iron forge could only yield batches of several hundred pounds. hus a chain could be made link by link. By the sixth century BCE chains were available and could be produced in any required length. A typical link would be about one foot long. It was a straightforward transfer from organic rope to iron chains. In the Chinese case, in the absence (Top) Figure 5.4. A Chinese catenary bridge supporting people and livestock. (Above) Figure 5.5. Chi-Hung catenary suspension bridge, 1470. An outstanding example of Chinese iron-chain suspension bridges. 149

Chapter 5

of wheeled vehicles for either commerce or military purposes there was little incentive to develop level-deck bridges. he catenary bridge served both man and beast very well. here were, however, a very limited number of level-deck pedestrian bridges that appeared, although they were the exception. he idea of the number and distribution of suspension bridges in China clearly establishes the signiicance of this notable technology. he Lu Ting bridge of 1705 is hallowed as a Chinese landmark in recognition of the passing of Mao’s revolutionary army in the twentieth century, which was a pivotal event in the establishment of the communist regime in China. Spanning more than 300 feet, the Lu Ting is an outstanding symbol of iron-chain suspension bridges, having been in service a century before the irst Finley bridge.

Figure 5.6. Chuka-Chazum iron-chain bridge in Bhutan showing details of the anchorage structure.

150

Th e O r i g i n s o f t h e M o d e r n S u s p e n s i o n B r i d g e

Bishop Faustus Verantius is a special case of a bishop without a diocese as a result of the Turkish invasion of the Balkans. He had lost all of his spiritual lock, leaving him free to pursue other activities. He was associated with the fortiication work at Vienna to stave of the invasion of the Turks. Especially important from our point of view, while on a tour of Italy he published an illustrated text in 1617 entitled Machinae Novae.8 In this text he illustrates an iron cable-stayed bridge but fails to cite any references to actual structures built according to this design. he idea was there, however, in print. (Top) Figure 5.7. One of the earliest descriptions of an iron suspension bridge is found in Machinae Novae by Faustus Verantius, 1617. here is no evidence such a bridge was ever built. (Above) Figure 5.8. Wellington’s famous catenary rope bridge at Alcantara, Spain, built using ship’s ropes in 1792 to replace a span of this Roman bridge destroyed by the French. Such rope military bridges were oten used by military engineers in Europe. 151

Chapter 5

he record is incomplete, but little appeared during the next two centuries to advance suspension-bridge technology. We have evidence of bridges using ropes for military purposes using suspension-bridge technology. hus it was military engineers who had experience building catenary bridges. One of the earliest iron-chain bridges of which we have details was the chain bridge of iron erected in the province of Kweichow, China, in 65 AD. Built by a military general, it supported its load by four chains. his bridge is not only noteworthy as an outstanding early example of a Chinese chain bridge, but was also the irst step toward a level-deck bridge centuries (Top) Figure 5.9. Carrick-a-rede Bridge, County Antrim, Northern Ireland. he catenary rope bridge existed from early times and may well have been known by James Finley. (Opposite) Figure 5.10. Wynch Bridge of 1791 over the Tees River, England. An early catenary chain bridge for pedestrian use.

152

Th e O r i g i n s o f t h e M o d e r n S u s p e n s i o n B r i d g e

before such a feature appeared in a design by Judge James Finley. he illustration from Tyrrell shows the design of such a bridge with vertical suspenders and a span of about 330 feet clear.

A Case of Technology Transfer While we are unaware of any Chinese-inspired suspension bridge designs appearing in Europe at the time that Verantius published his work, a number of primitive bridges appeared in Europe that may provide links to Finley’s later works. hey need to be at least mentioned. As an Ulsterman probably living in County Antrim, Finley knew about and may even have passed over the Carrick-a-Rede catenary bridge on the Antrim coast near Bally Castle. A stone stack rises out of the sea sixty feet from the Ulster shore. In order to gain access to the stack and to facilitate ishing, a rope catenary bridge appeared from ancient times to cross the ninety-foot-deep chasm. he wood deck rested directly on the rope cables. he rope has been renewed many times and the bridge is now supported on wire ropes. With the strong possibility of disintegration in the moist sea air, the ropes were replaced frequently—perhaps even annually. he only other suspension bridge in the British Isles was erected in County Durham two miles above the village of Middleton. he chain bridge with a span of seventy feet was only two feet wide and appeared in 1741. It was a pedestrian bridge to accommodate miners crossing the river Tyne. Apparently the bridge 153

Chapter 5

collapsed in 1802, killing several people. he replacement served until 1908 and possibly even later.

Western Pennsylvania Having established, albeit briely, the background for iron link-chain suspension bridges employed from ancient times, the stage has been set for examination of James Finley’s seminal work on suspension bridges. First, however, it is essential to establish the conditions under which the irst modern suspension bridge was erected. During the eighteenth century and early portions of the nineteenth century, western Pennsylvania represented a sparsely settled area based largely on agriculture. A signiicant portion of the land was claimed and settled by Ulster Presbyterians from Northern Ireland. Included were many Finley family members. Historical records mention the Reverend James Finley, who served as a leader of the Presbyterian community. Although largely undeveloped, there were notable ironworks and other select industries in the area. As we shall see, the ironworks made possible the pioneering work of James Finley. With a frontier society unfettered by controlling laws or even federal control, the population celebrated their freedom. In a landscape marked by very poor roads throughout the region, tracks were used by pack animals to move agricultural products. It was a diicult and expensive situation with little likelihood of major transport works, and limited because the principal water connections to unite people and goods in the area lowed westward instead of to the east, where the markets lay. his area was ater all west of the eastern continental divide. A proitable venture promoted by farmers in the late eighteenth century was to turn surplus grains into distilled spirits, which represented a value added to the product and made transport easier. he federal government became interested in controlling and taxing the production of whiskey. As a result, tensions rose, the local community pitted against the involvement of the government. Dubbed the Whiskey Rebellion, it put the federal government’s authority at stake in the new republic.9 Just as President Eisenhower was reluctantly forced to order federal troops to Little Rock, Arkansas, to quell unrest erupting regarding school integration, the 154

Th e O r i g i n s o f t h e M o d e r n S u s p e n s i o n B r i d g e

success of the government forces ordered to western Pennsylvania clearly established the authority of George Washington’s government and caused the rebellion to collapse. It was a triumph for Washington and the role of a weak national government at the time to take charge of both local and regional conlicts. he transformation of this Pennsylvania wilderness into a European-style civilization is presented in detail in the book he Planting of Civilization in Western Pennsylvania.10 he establishment of the iron industry in western Pennsylvania was essential to enable Finley to erect the world’s irst level-deck chain suspension bridge. Without an established iron industry Finley could not have erected his patented bridge in such a region. Even before the American Revolution, iron was discovered in western Pennsylvania. he ore deposits were widespread and of suicient quality to attract the development of furnaces and forges.11 hree of the earliest furnaces were Isaac Meason’s Union Furnace on Dunbar Creek, irst put into blast in 1791 but enlarged in 1793 in partnership with Moses Dillon and John Gibson. An advertisement in the Pittsburgh Gazette claimed that Meason could produce cast-iron stoves, end irons, and an assortment of cooking kettles. By 1801, the year of Finley’s irst chain bridge, Meason was the sole proprietor of the Union Iron Works. On the west side of the Allegheny River at Dunbar Creek the Union Furnace operated a sawmill, a forge, and a boatyard. he irst rolling mill in the region was founded by Jeremiah Pears and began production in 1800 on Redstone Creek. It rolled iron sheets, which replaced the earlier hand-forged sheets, had a slitting mill, and produced wrought iron for handmade nails. It was an achievement that should be celebrated by those interested in the history of technology. Pears had a mill on Redstone Creek as early as 1784. Situated on Mounts Creek, the Mt. Vernon furnace, erected circa 1798 by Isaac Meason, was reconstructed in 1881. It employed as many as sixty men and was operated by Isaac Meason, Jr. he irst modern iron-chain suspension bridge by Judge James Finley dates from 1801. In the year 1801, I erected the irst bridge on this construction over Jacob’s Creek, on a contract with Fayette and Westmoreland counties, to build a bridge of seventy feet span, twelve and a half feet wide, and warrant it for ity years (all but the looring) for the consideration of six hundred dollars. Nothing further

155

Chapter 5

of the kind was attempted for six years. he exclusive right was secured by patent in the year 1808.12

With a span of seventy feet, this diminutive bridge displayed all of the essential features of the modern suspension bridge. Crossing Jacob’s Creek on the Connellsville to Mt. Pleasant Turnpike, the bridge site lies close to Meason’s forge. As a leading ironmaster with interests in a number of industries that made him one of the wealthiest men in the region, Meason was also interested in and was a principal in the Union and Cumberland Turnpike road system established by the general assembly of Pennsylvania. Meason was at the irst meeting of the two county commissions for the Finley bridge. Although evidence is

Figure 5.11. Rendering of the Isaac Meason House near Uniontown, Pennsylvania.

156

Th e O r i g i n s o f t h e M o d e r n S u s p e n s i o n B r i d g e

circumstantial, the author believes that Meason provided the iron for the chain links and suspenders and very likely supplied the ironwork for other chain suspension bridges erected in the area. here is a dearth of artifacts regarding Finley’s early work; nothing remains at the Jacob’s Creek crossing. However, the Lehigh Gap Bridge of 1826 represents a ine example of an early work using the Finley system. While little survives of Meason’s, or indeed any of the other early iron furnaces, it is well to discuss his Mount Braddock mansion as an extant structure of considerable architectural merit. Relecting his wealth and prominence in the area, Isaac Meason Sr. decided to build a grand house, arguably the inest early high-style structure west of the Allegheny Mountains. he design represents the work of the architect/builder Adam Wilson from Scotland. It was also Wilson who turned his hand to landscape architecture and produced notable entrance grounds to this mansion. All of the work on the house and the gardens was under the direct control of Wilson. He was trained as a joiner (i.e., carpenter) at about the same time he was also involved in another high-style home for Isaac Meason, Jr. at New Haven. It is so modiied that the original structure is not discernable. Wilson, a close friend of Meason, never married and died in 1823.13

Internal Improvements Movement While many espousing the study of American history are familiar with the Turner thesis for interpreting American history in terms of the westward development of the country, and the related slogan “go west, young man” is well known, there is another powerful movement directly related to the building of the nation’s infrastructure. Despite Gallatin’s plea for the involvement of the federal government in this movement, the government really served as a bystander in the development of a transport network of roads, bridges, canals, and river improvements.14 Later, railways were added to this movement, and by the mid-nineteenth century had become the dominant technology for the rapid expansion of a transportation network, which ultimately stretched from coast to coast. In Canada, a transcontinental railway linking British Columbia with the rest of the nation ensured that the west coast province would play a signiicant role following the British 157

Chapter 5

North America Act of 1867 that formed the Dominion of Canada from the British colonies of North America. With little concern for Gallatin’s call to a more perfect union, states, cities, and local entities undertook the mandate of the Internal Improvements Movement in the promotion of great public works without the leadership or support of the federal government. A notable exception was the building of the National Road, which passed through western Pennsylvania. Let on their own, without strong and comprehensive national leadership, the movement in many cases became highly competitive. Perhaps the most indicative illustration of the competition was the construction by large east-coast cities of a link with the Great Lakes and western waters of the Ohio River. Following the great successes of the Erie Canal, Philadelphia promoted the Pennsylvania Mainline Canal. At the same time, Washington’s dream of improvements of the Potomac River resulted in a new venture, the Chesapeake and Ohio Canal. his elaborate scheme envisaged a towpath canal along the Potomac River followed by a towpath along the Casselman River, thence to Pittsburgh and the western waters. If completed, this large-scale scheme would have provided competition with the highly successful Erie Canal. he Pennsylvania Mainline Canal and later railway system ensured that Pittsburgh and western Pennsylvania counties were fully integrated into the social and economic fabric of the great Commonwealth of Pennsylvania. While canals and river navigations represent the most visible manifestation of the early Internal Improvements Movement, the construction of turnpike roads and toll bridges represented a leading development in early nineteenthcentury America.

Links in a Chain In two seminal papers, the Portfolio and the more obscure Chain Bridge of 1811, James Finley proclaims his new suspension bridge invention.15 hese concise papers presenting a wealth of engineering details reveal Finley’s knowledge of overseas bridge developments. He also lets the reading public know of his appreciation of the mechanical properties of wrought iron as well as his knowledge of the catenary curve and what we would call the funicular polygon. In an age when master builders produced designs with little or no engineering 158

Th e O r i g i n s o f t h e M o d e r n S u s p e n s i o n B r i d g e

sensibilities, Finley overcomes this lacuna in engineering design by producing a method for detailed analysis for determining the shape of the chain and the length of the links. Using an engineering graphical method Finley could determine the curve of the chain. He says, To ind the proportions of the several parts of a Bridge of one hundred and ity feet span, set of on a board—one hundred and ity inches for the length of the bridge, draw a horizontal line between these two points representing the underside on the lower tiers of joists—on this line mark of spaces for the number of joists intended in the lower tier, and raise perpendiculars from each, and from the two extreme points, then fasten the ends of a strong thread at these two perpendiculars, twenty three inches and one quarter above the horizontal line—the thread must be so slack that when loaded, the middle of it will sink to the horizontal line, then attach equal weights to the thread at each of the perpendiculars and mark carefully where the line intersects each of them.16

Because the chain links span panel point to panel point, their length varies with their position in the chain. he graphical method also provides a means of assessing the force in the chain as well as determining the maximum load Figure 5.12. he earliest depiction of the patented Finley chain bridge. It appears in Finley’s 1811 pamphlet; it is probably his patent drawing.

159

Chapter 5

at the top of the towers. he strength required of the link at any position is used to determine the cross section of the iron of the link. Finley was aware that the strength at the end of each link must be carefully considered. He recommends that the ends that are to be lap-forged should be “upset” by heat to provide additional bearing surface in an area of unusually high stresses. hus, the link is closed by forging the free ends together. It appears that a number of chain failures that are reported in the literature and described as design failures were caused by faulty forging. hese reports are without exception devoid of any engineering data. he failures may also be ascribed to a failure of the ittings such as the saddles at the top of the tower or even movement of the anchorages. It must be remembered that despite Finley’s cautionary advice, many of the earliest chain suspension bridges lacked adequate stifening trusses, which made them too lexible in cross winds. In addition, there was no control on the amount of load a bridge could safely support. he graphical method devised by Finley enabled him to study the forces in the chains for uniformly distributed deck loads as revealed in the quote above. He correctly establishes the relation of the sag/span ratio to the force in the chain at the tower top. Earlier scholars of the Portfolio article, such as Sir Samuel Brown and homas Telford, considered the 1/7 or 2/13 sag/span ratios to be excessive. Brown used a sag/span ratio of 1/10 in several designs. Although not at all obvious, the 1/7 ratio is very important because the maximum load at the tower exactly balances the load on the deck. hus, a coniguration of 1/7 ratio results in a rule-of-thumb method that eliminates Finley’s graphical and analytical method. He also predicts, based on his studies, that bridges of 1,000 feet clear span could be built. his prediction was manifest in the 1,008-foot span of the Wheeling Suspension Bridge of 1849. In addition to his studies of the shape and strength of bridge chains, Finley discusses the design of the tower, the anchorages of the chains, and the stifening of the deck. Well in advance of the debut of John Templeman’s patent on multi-span suspension bridges, Finley is concerned with the overturning of the towers on multi-span bridges with only one span fully loaded. Wrought-iron bars secured the timber towers to the abutments. Before leaving the subject of link chain, Finley had to solve the problem of the funicular polygon. He concisely states that with a heavy load on the deck, 160

Th e O r i g i n s o f t h e M o d e r n S u s p e n s i o n B r i d g e

the resulting shape of the chain is no longer a catenary curve. By loading the chain at each joist, the chain between the links assumes a funicular polygon that carries the deck load to the chain by a series of suspenders attached to the chain where two links intersect. hus, the chains support the deck at their ends only, avoiding any bending, which satisies the conditions for a funicular polygon. One can envisage such an arrangement as rather like a giant bicycle chain. he design must ensure that joints at the end of the links can rotate freely and not induce bending in the chain. Finley addresses the situation by ensuring that the joint at the end of each link performs as a smooth pin. He connects the chain to the deck at the joint of the links to avoid bending. In his paper, Finley quotes an encyclopedia reference on the strength of wrought iron.17 Using a factor of safety of two, he selects 60,000 psi in tension for designing the chains and other iron components. Accepting the stress values in the encyclopedia, he makes no attempt to proof-test each link as Telford did on the chains of the 1826 Menai Bridge. From the same encyclopedia he mentions the famous cast-iron bridge at Coalbrookdale, the Sunderland Bridge over the Wear, together with a cast-iron bridge over the Teme in Hertfordshire. He presents this information in his 1811 paper and says “I venture to pledge myself, that one third of the money should erect such a bridge and keep it open for ever.”18 It was a bold conception, in advance of later American bridges such as the Wheeling and Brooklyn suspension bridges. Finley was clearly a visionary concerning the future of his invention. Not content with developing a new bridge design, Finley undertook a study of the forces in the chains at various sags (i.e., delections) from the tower to the center of the span. he results are presented in his Portfolio paper. With his graphical method he had a means of determining the length of each link but also the force in the link by the loads on the deck. He shows conclusively how the maximum force varies with the sag/span ratio. In addition, the graphical method could determine the force needed in each link of the chain. hus, drawings and dimensions of the chain could be supplied to the fabricator. he main chain was attached to a special forging at the top of the tower as was the back chain that passed to the anchorage. In order to balance the forces exerted on the tower top, Finley states that the angle of the chain on both sides must be equal to avoid lateral loads. 161

Chapter 5

In an era before soil mechanics and foundation engineering, the design of foundations was based on traditional practice. In the absence of sound information, Finley suggests that the chain must be weighted down in the anchorage with stones to avoid slipping. Another signiicant contribution presented by Finley was his understanding of the need for a stifening truss to distribute live loads along the length of the deck. He proceeds to give details of the type of timber framing needed. An unstifened deck was the source of signiicant movement induced by the wind. Despite this cautionary advice, a number of Finley chain suspension bridges were built without adequate deck stifness.

John Templeman, Entrepreneur To end James Finley’s noteworthy paper in Portfolio and its derivative paper in 1811 it would appear that Finley would “go out” on a crescendo. He, however, produced the following paragraphs, which open another chapter in the saga of the early chain bridge: In March 1808 I entered into an agreement with Mr. John Templeman, of Georgetown, Maryland, by which he was to receive one half of all the monies arising from what permits or patent rights he could dispose of for and during the term of ive years. All contracts to be in my name, and the money payable only to my agent in the city of Washington, who should pay one moiety over to Mr. Templeman. But in delineating the principles of my bridge in the patent oice, I spoke of only one arch or space, and it seems that Mr. Templeman took it into his head that I should go no further; accordingly soon ater our agreement, he took a patent for all continuations, but he has thought better of it since, for I have gone on to receive the perquisites for all the spaces, with his knowledge, and without any complaint from him on that head. In the same article it is provided that the parties shall not grant permission to build bridges on this plan at less than one dollar per foot span, without any discrimination as to breadth. But gentlemen have proceeded to build with design to pay when the work should be completed, and have always paid on demand.19

162

Th e O r i g i n s o f t h e M o d e r n S u s p e n s i o n B r i d g e

he arrangement with Templeman indicates that with the exception of the earliest chain suspension bridges, Finley acted as a consultant rather than a builder. His role is in line with the modern use of consultants. Not only was Templeman engaged as Finley’s agent but he also shared in the royalties of bridges built according to Finley’s method. Templeman took it upon himself, apparently without the knowledge of Finley, to patent variations of the basic Finley invention. A patent was issued to Templeman on August 16, 1808. Templeman’s most obvious claim was for a multi-span bridge. Finley disagrees that this is a substantial improvement but thinks it is a logical extension of his pioneering work and hence, as he says, not worthy of patent protection. In the early days of the Republic patents were relatively easy to procure. In his statement on deck stifness, Templeman makes what is believed to be the irst mention of wire or bar chains. homas Telford considered the use of wire for cables as early as 1816. Josiah White and Erskine Hazard erected a wire pedestrian bridge across the Schuylkill River in Philadelphia in 1816 as the irst wire suspension bridge in North America. Although this pedestrian bridge was Figure 5.13. John Templeman’s Potomac River Suspension Bridge.

163

Chapter 5

only two feet wide, it spanned 408 feet. In his letters patent regarding wire or iron-bar chains, Templeman states, Besides the above mentioned improvements and applications I have invented an entire new Chain Bridge with a method of suspending ropes or cables of Hemp or of wire made fast into cords and then into rope and then into Cable—or iron chain crossways. he ropes of wire are intended to counteract or avoid the efect of Frost on the chains as it is possible in very cold climates the frost may render the chains unsafe. he wire ropes are to be well oiled and covered with oil-cloth and molded with spun yarn—and at the distance of 5 or ten feet they are to have two pieces of iron of six inches long and half an inch thick worked in at the time of laying the rope to keep the suspenders at proper distance.20

In the popular press it may appear that Finley was the builder of a number of chain link suspension bridges. One should, however, note that Templeman served as the sole agent for Finley. As well as acting as an agent it is not clear how Templeman was involved in each bridge erection. Although bridges over the Potomac near Washington and as far north as Newburyport, Massachusetts, and others are credited to him, there is no information on Templeman’s engineering work. In 1836, a disastrous ire destroyed the patent oice and all of the patent documents. Many patent documents were restored, including Templeman’s patent of 1808. Nothing was forthcoming regarding the Finley patent of June 17 of the same year. Now we must turn to Templeman’s biographical information. Extensive research has revealed that John Templeman emigrated from Europe to America in 1780, before the Treaty of Paris ended the American War of Independence. He was heralded by the dental profession as a leading dental surgeon. We ind him irst setting up his practice in Newport and later Providence, Rhode Island, before moving to Boston. He advertised extensively that he would implant live teeth in patients’ mouths for four guineas per tooth. He probably obtained his live teeth from poor members of society. He also treated scurvy of the gums and specialized in children’s teeth. It appears that part of this treatment of children resulted in straightening their teeth. In 1783, Templeman married Mehitabel 164

Th e O r i g i n s o f t h e M o d e r n S u s p e n s i o n B r i d g e

Loveless, who had been baptized in 1755 and was approximately 28 years of age. We have yet to establish John Templeman’s dates except to note that he probably died in the 1830s. Templeman acquired a coastal trading ship named Mehitabel and traded to east-coast cities. We know he sold axes, or perhaps tomahawks, in Delaware, but he notes they were not selling well. he Boston City Directory of 1789 lists John Templeman as a broker and dentist with an oice opposite the northeast corner of the State House and the south Latin School street. In 1791, his son George was born in Boston. In 1792, while still in Boston he bought lots in Washington and a year later he purchased additional Georgetown properties. Also in 1792 the Bank of Columbia was chartered and Templeman served on the board of directors of this and another bank. Research into Templeman’s real-estate dealings has resulted in eighteen pages of deed and purchase records.21 He clearly was a land speculator. John Templeman moved in the upper echelons of political and inancial society in Washington and environs. He knew President George Washington and other members of the cabinet, as well as being associated with banking interests in the District of Columbia. An especially important friend was Benjamin Stoddert, who became the irst Secretary of the Navy. It was through this contact that Templeman was commissioned to supply heavy timbers for the construction of warships in Washington. To accomplish his contract, signed August 1799, Templeman acquired property in Cumberland, Maryland, to secure the needed timber. He moved his slaves to cut timber and loat the logs to Georgetown, and it was there that the logs were sawn and prepared for the navy. It should be noted that at the time Benjamin Stoddert borrowed $1,000 from Templeman, who provided him with a bank check. In a career-long engagement with publications in the history of technology, it has been rare to ind any presentation of human interest for leading engineers. hese publications include manuscript collections, which are also devoid of material portraying the lives of engineers. hus, the following material gives details of an extraordinary event that took place in Bladensburg, Maryland. In this portrayal we witness John Templeman in action. Next comes that wonder of childhood, the Wire Dancer, with his balancings and other accomplishments. his was Mr. Templeman, a dancer on the slackwire.

165

Chapter 5

he exhibition was in Tattison’s dancing room. We got there at early candle light. he room was brilliantly lighted. A large wire fastened at each end of the room, near the ceiling, hung in a curve, the middle of it within twelve or iteen inches of the loor. I remember the pouring in of the company till the room was illed, as the phrase is, ‘with all the beauty and fashion of the place’. Still better do I remember, ater a note of preparation from another room, which bespoke and commanded silence, the entrée of Templeman—a tall man, superbly attired in a fanciful dress; of a military air, with a drum hung over his shoulder by a scarlet scarf. It was such a picture as I had never seen. Saluting the company with dignity, he placed himself upon the wire; then giving a hand to his attendant, he was drawn to one side of the room, and, being let go, swung at ease,—beating the drum like a professional performer. He performed all the usual exploits, balancing hoops, swords, etc.—and to crown the whole, danced what I had never seen before, a hornpipe, in superior style;—his spangled shoes, in the rapidity of his steps, producing upon me a most brilliant efect.22

Chapter 5 Notes 1 2 3 4 5 6 7 8 9 10

Charles Singer et al., A History of Technology, 5 vols. (New York and London: Oxford University Press, 1956), vol. 4. Historic American Engineering Record of the U.S. National Park Service, founded in 1969. Carl Christian Schramm, Historischer Schauplatz (Leipzig: Breitkopf, 1735). Carl W. Condit, American Building Art, he Nineteenth Century (New York: Oxford University Press, 1960), chapters 3, 4, 5, 8. John Alexander Low Waddell, Bridge Engineering (New York: John Wiley & Sons, 1916), vol. I, 2. hornton N. Wilder, he Bridge of San Luis Rey (1927). C.L.M.H. Navier, Memoire sur les Ponts Suspendus – Rapport à Monsieur Becquey (Paris: Carilian–Goeury, 1823). Verantius Faustus, Machinae Novae, 1617. Solon J. Buck & Elizabeth Hawthorn Buck, he Planting of Civilization in Western Pennsylvania (Pittsburgh: University of Pittsburgh Press, 1979), chapter 19. Ibid., chapters 11, 13.

166

Th e O r i g i n s o f t h e M o d e r n S u s p e n s i o n B r i d g e

11 12 13

Ibid., chapter 13. James Finley, in Portfolio, 3, no. 6 ( June 1810); Finley, A Description of the Chain Bridge (Uniontown, PA: William Campbell, printer, 1811). Christina A. Spyrakos, he Isaac Meason House: A Catalyst for an Historical and Architectural Study (Morgantown, WV.: MA thesis, West Virginia University, 1995).

167

6

French Movable Dams in America

A

mere recital of various nineteenth-century French movable dams and their inventors lacks an historical context in evaluating the signiicance of movable dam technology of the nineteenth century. No attempt is made here to produce a deinitive narrative, but rather to show the ancient origins of these types of dams. Traditionally, river men desired, and even in some cases demanded, open river navigation during periods of navigable river levels, while at other times they needed hydraulic structures such as locks and dams to create suitable slack-water pools, which would permit navigation even in times of low water. Traditional hydraulic works were unavailable to meet both these demands without the installation of movable dams. When erected, a movable dam created a slack-water pool like an ordinary ixed dam, and in each case a sluiceway was provided in the dam for the passage of vessels, but when the movable dam was lowered, open navigation permitted vessels to pass over the dam in an unrestricted fashion. By adapting French movable-dam technology, the United States Army Corps of Engineers proceeded to employ movable dams on numerous sites, principally along the 981-mile Ohio River and its tributaries. his adaptation of a foreign technology represents an outstanding example of technology transfer, a concept cherished by historians of technology. It is not just a recital of inventions that garners the interest of historians, but how these inventions are transferred from person to person, and indeed from nation to nation, that constitutes a fundamental aspect of any study of technology in the last few centuries. As we shall see, the transfer of technology gave rise to a number of modiications so there are families of movable dams that one would need to consider in adopting a particular system. 168

F r e n c h M o va b l e D a m s i n A m e r i c a

With a centralized bureaucracy established in Paris, the Bureau des Ponts et Chaussées led the world in the development of engineering, with applications throughout France and indeed the world. It is little wonder that the United States Army Corps of Engineers was modeled ater this French corps. Even the early Corps of Engineers’ uniforms were patterned ater the French corps.

Domesday and Beyond Without a detailed study of hydraulic works in Europe, the predecessors of the French movable dam cannot be presented, since this would require an extensive supplementary study. A better course of action appears to be to use the case-study method, a widely applied technique by legal professionals. he case for the background for French movable dams based upon published data is the Domesday Survey in Britain of 1086.1 Ater achieving control of England as a result of the Battle of Hastings in 1066, England’s new Norman French masters commissioned a nationwide survey of resources. he Domesday Survey of 1086 has become one of the most important documents in English history. For our purposes we should know that more than one thousand mills were in operation at the time. In an age before steam power these were driven by waterpower. In order to provide suicient “head” to drive waterwheels and associated gearing, one can assume that nearly every one of these mills was on a waterway. he usual means of providing suicient head was to throw a dam across the watercourse. While an eicient means of developing waterpower, mill dams precluded boats from using what might otherwise be a navigable waterway. he most straightforward means of accommodating both navigation and milling was to leave an opening in the dam, which could be closed by various types of gates to impound water. Gates hinged at the level of the foundation sill, called “drop gates,” could be dropped to pass vessels through the dam. here were other devices that could achieve the opening and closing of the dam. Movable portcullis gates, reminiscent of the entrance to an ancient castle, were also employed. By installing a sturdy beam at the top of the opening, the opening could be closed by a series of wooden members called needles in French. Each rested on the sill of the dam at the bottom of the channel and against the so-called balance beam at the top. hese needles placed side by side efectively held back the water to permit 169

Chapter 6

milling to continue. By means of control conduits with valves, water could be let onto the waterwheel located nearby. Mill dams with sluices were widely applied throughout continental Europe. hese single movable gate openings called lash locks were developed into a systematic procedure on the Yonne River, a tributary of the Seine that passes through Paris on the way to the sea. he valley of the Yonne produced large quantities of timber which, in the form of logs, was loated to Auxerre where the Yonne joins the Seine. A notable feature of the river was the large number of mill dams complete with lash locks. At a predetermined date, a great lotilla of logs was dumped into the river at the upper reaches near the Massif Central, and the logs would be rolled into the water to pass through the opening in the dam. A large crest of water moved downstream with the leet of logs. Operators at the next dam, seeing the lotilla approaching, would open their sluice gate and direct the freshet through the dam and on downstream. he sequence continued down the entire length of the river until the leet of logs reached the Seine and was loated downstream. his lashing technique, which occurred more than 400 years ago, was also a feature of certain navigations in the nineteenth century. he widespread use of the lashing system clearly needed to be improved if river navigation was to ind wide application.

Bear-Trap Gate he use of needles to close of an opening in a dam can trace its origins to the ancient lashing system used on the Yonne and elsewhere. he later nineteenthcentury development of the Poirée needle dam appears to be quite straightforward but nonetheless an important engineering development. Before we look at this technology, however, we must move in time and place to improvements on the Lehigh River in Pennsylvania by Hazard and White. As early as 1818 Josiah White, a well-known Philadelphia businessman, and Erskine Hazard served as managers of the Lehigh Navigation Company.2 hrough river improvements they sought a means of improving navigation on the Lehigh to facilitate the movement of anthracite coal to the Philadelphia market. Improvements on the Lehigh River began with clearing the channel of snags, stumps, boulders, and other obstructions, at the same time enhancing the low 170

F r e n c h M o va b l e D a m s i n A m e r i c a

by erecting wing dams protruding from the river bank to the main channel. he narrowed channel increased low and helped to move sediment downstream. hese measures were inadequate and it became clear that a slack-water system could achieve the goal of greatly improving the coal-carrying capacity of the navigation company. While White and Hazard were fabricating a prototype movable timber gate, members of the public inquired as to what was being built with solid timber “leaves.” he reply, which may well be legendary in its attempt to conceal its purposes, was that it was a bear-trap. To this day this moniker continues to designate a special type of movable gate. It reminds one of a similar situation during World War I when the curious public inquired as to what was being fabricated and they were told that they were water tanks. hus the word “tank” as an armored vehicle is still in use today. he original bear-trap gate was a simple arrangement of two solid timber leaves. When not impounding water, the leaves rested upon the ixed bottom sill of the dam. When water was injected under the leaves they would rise and form a hydraulic barrier. Although the width was limited, the bear-trap gate served the Lehigh Navigation well. he standard operating method of successive openings of the gates was reminiscent of the French lashing system as used on the Yonne. Width was always a problem with bear-trap gates, as oten a pass would be 120 to 130 feet in width. Leaves of this size would be diicult to construct and even more diicult to operate. he three principal problems to be solved in connection with bear-trap dams are irst, to secure the power to start the gates when they are to be raised, second, to prevent warping and twisting during the raising and lowering, and third, to construct the gate so that it can be used for passes of considerable width without having intermediate piers.3

Because the standard bear trap proved unsuitable with its inherent lack of width for a pass or a weir on a movable dam, it saw only limited application. In American experience the most signiicant was in the many Chanoine wicket dams in the 981-mile-long Ohio River improvement, where pairs of bear-trap gates were installed in each dam to provide a quick and eicient means of 171

Chapter 6

passing debris lodged behind the dam. hey were most efective in passing ice lows downstream during the winter. Operating the bear-trap gate required the injection of water under pressure beneath the leaves. his required a “head” of water to raise the gate. On the Ohio and Lehigh, suicient water pressure was available behind the dam. But in many cases, auxiliary reservoirs were required. In the case of the Neuville Dam on the Marne in France, a hénard shutter was used. his was apparently the only large-scale application in France. Subsequent gates saw improvements in which the modiications were associated with the inventor’s name. Prominent amongst this group were gates by DuBois in 1862, Carro in 1817, Girard in 1868, and Bruno in 1867. he sluice gates, such as those itted with needles and used on the Yonne and elsewhere, were not suitable to provide the wide-open passage clamored for by the rivermen when river conditions were favorable. he upper beam on which the needles rested in the sluice could span only a limited distance, precluding open navigation. Interior piers could not be used because they would block passage of vessels. Although using cables attached to the abutments of the riverbank was a reasonable theoretical solution, they were strong enough but too lexible to be used in practice.

Needle Dams he problem of open navigation coupled with the slack-water system was ingeniously solved by Ingénieur Poirée with the introduction of the needle dam. In 1834 the new system married the idea of closely spaced needles but this time supported not on a beam but on iron frames at a close spacing of approximately four feet. At the top of the iron frame, called a fermette by French engineers, a beam spanning from fermette to fermette supported closely spaced timber needles. At the bottom, the needles rested against a ixed sill at the base of the dam, closing of the low of water and creating a slack-water behind the dam. Vessels could pass from level to level through the construction of an associated lock. he fermettes rested lat on the sill of the dam, but when required, they were hoisted to a vertical position one ater the other. As each one assumed the vertical position, the upper beam was attached to adjacent frames providing lateral 172

F r e n c h M o va b l e D a m s i n A m e r i c a

stability. Once all the fermettes were raised, an iron walkway on top of the fermettes was used to install or remove the needles as necessary. Ingénieur Poirée erected the irst needle dam at Basseville, France, with the fermettes spaced one meter apart.4 Although the trestles, the term used in the United States, were two meters high, the water impounded had a maximum height of one meter. Scores of needle dams appeared in the French countryside on such rivers as the Seine and Yonne, the Nivernais Navigation. Most of these dams were of comparatively low head requiring slender needles, four to ive inches square, which could be handled manually. Greater hydraulic heads required larger needles, which were too heavy for manual work, and required mechanical handling devices such as a crane or a winch to maneuver. Many such dams are still in use in France. he author’s irst encounter with the needle dam system occurred at Auxerre, where the lit is less than four feet. As we shall see, the Chanoine wicket dam proved to be the choice of the U.S. Army Corps of Engineers, resulting in many applications on the Ohio and the Great Kanawha rivers and many other places. he only practicable application Figure 6.1. Poirée needle dam. (Edward Wegmann, he Design and Construction of Dams [New York: John Wiley & Sons,1900], 155)

173

Chapter 6

of a large needle dam was by Benjamin Franklin homas, a Corps of Engineers oicer, in his design on the Big Sandy River. his river and its tributaries provide a gateway to an extensive and yet nearly inaccessible region, with huge resources of coal coupled with timber spreading over the various mountains and hills. In the nineteenth century, the exploitation of these resources without roads or railways rested entirely on using the Big Sandy under favorable conditions. hus it appeared that improving the river would open economic development of the region and allow coal in particular to enter the Ohio system. he river represents a signiicant but yet minor tributary of the Ohio River, characterized by noticeably low low a number of times a year, particularly in late summer and early autumn.5 At such times, barges were tied up waiting for a rise in the river. To increase the ability of the Big Sandy and its tributaries to handle commercial traic, early attempts, extending into West Virginia Figure 6.2. Big Sandy needle dam. (B. F. homas and D. A. Watt, he Improvement of Rivers [New York: John Wiley & Sons, 1918], 566)

174

F r e n c h M o va b l e D a m s i n A m e r i c a

and Kentucky, were made to improve the river by removal of snags, rocks, and other impediments. he Chanoine wicket system was under consideration for improving navigation on the Big Sandy, but in order for the Chanoine system to operate without diiculty, the wickets had to be spaced with a gap of approximately four inches. On the Ohio and Great Kanawha rivers the resulting leakage was less than the low low of the river and was judged to be acceptable. On the Big Sandy, however, because of the problem of low low together with the large deposits of sand, the amount of leakage could not be tolerated. Enter Benjamin Franklin homas, who proposed to William Craighill, who was associated with the taming of the Great Kanawha River and who later became Chief of Engineers, the idea of building a needle dam or dams on the Big Sandy to counter low low in late summer, which oten reached a minimum of ity cubic feet per second.6 he preferred type would have been the self-acting Chanoine wicket, but because of very large movements of sand that could not be readily handled by wicket gates, an alternative was clearly needed. he French needle dam, which was not self-acting, had to be maneuvered by a loating maneuver boat with a steam derrick mounted for handling very large needles more than fourteen feet in length. Removing the needles meant that the current would lush out the sand and move it on downstream. A unique application of the needle dam was the Bonnet Carré spillway erected parallel to the bank of the Mississippi just north of New Orleans.7 When opened, the nearly mile-long dam diverted a copious supply of water away from the Mississippi and the city of New Orleans and moved it directly into Lake Pontchartrain. he needles were of such size that they were positioned by a railmounted crane. he Louisa Needle Dam, built from 1891 to 1897, featured a 52-foot-wide by 255-foot-long masonry bed, at one end of which was constructed a lock for the movement of vessels featuring steel-framed lock gates. Adjacent to the river wall of the lock, a 130-foot-wide pass was located, separated from the weir by a 12-foot-wide pier. he dam stretched across the entire river and reached 400 feet in total length, including a 17½-foot-wide abutment. his arrangement provided a slack-water pool 15 feet deep over the bottom sill of the dam. Placed on 4-foot centers, the fermettes measured 15 feet, 2 inches on the pass opening but only 9 feet, 8 inches on the weir.8 Compared to French precedent 175

Chapter 6

and considering the height of the pool the fermettes were lightly framed of steel channels. Each pass frame weighed 1,140 pounds, whereas the weir was only 940 pounds. Just above the Louisa Dam the river divides into two branches, namely the Tug Fork and the Levisa Fork. hose involved in promoting the river improvement envisioned a series of locks and dams, presumably employing needle dams, to continue the slack-water navigation up the Tug Fork as far as Williamson, West Virginia, and on the Levisa Fork to Pikeville, Kentucky. Such a system would open up the vast coal resources of the entire region. he proposed series of locks and dams would require at least ten sites. Considering the limited low on each fork it appears that the slack-water navigation would be the customary mode of operation with the mobile dams in the raised position. Open navigation would have to wait for the occasional freshets coming down either or both forks of the river. Despite optimism on the part of the promoters, the full system never reached reality. Nevertheless, two needle dams were completed from Louisa to Catlettsburg on the Ohio River. he two locks and dams on the Big Sandy provided fully controlled navigation on the main stem of the Big Sandy. Under adequate low rates shipping could proceed some distance up each fork. Details of the later locks and dams are illustrated for easy comparison with the Louisa movable needle dam. While complete control of navigation with a slack-water system would clearly facilitate the movement of coal to market, before such a system could be provided the railways took over as the primary means of coal transport.9

Boulé Gates More than four decades since the irst needle dam, in 1874 Auguste Boulé introduced the Boulé gate. Instead of needles Boulé substituted wooden panels framed in iron that would rest against trestles. hese iron-bound wooden plates closed the opening and were a substitute for needles. he thickness of the wood panels varied with the position on the fermettes; the thickest were at the bottom with progressively thinner planks for the upper panels. his provided an eicient movable dam system to use and maneuver with limited leakage. Perhaps 176

F r e n c h M o va b l e D a m s i n A m e r i c a

the most notable application was featured at the Louisville–Portland canal in Kentucky, adjacent to the falls of the Ohio River. In order to control looding on the upper reaches of the Muskingum navigation a Boulé gate was installed at lock and dam number eleven. In the event of a freshet the Boulé gates could be dropped, letting the lood pass without inundating land above the dam. he Boulé gate saw limited application in North America and few if any remain in service. Instead of individual panels, plates to close the dam were tried. Apart from the need to handle rather large and heavy plates, the water pressure against the channels developed signiicant friction when attempts were made to raise the plates. To reduce friction, suggestions were made to use a “Stoney plate” complete with rollers.10 While an attractive idea, no example is known to have been constructed in North America. he irst application was at the Mulatière Dam on the Soâne River near Lyon, France. Ingénieur Janicki constructed six Boulé dams on the Moscow River in Russia.11 Using the basic Poirée system of fermettes supporting a curtain wall, Ingénieur Caméré introduced the curtain dam in 1876, which represents a marked departure from previous mobile dams based upon the needle dam. While the Poirée fermette reappeared in the Caméré design, the closure consisted of a rollup wall rather like a common commercial roll-up garage door. In the Caméré system the roll-up curtain featured wood slats hinged together with brass Figure 6.3. Caméré curtain in rolled-up position. (Edward Wegmann, he Design and Construction of Dams [New York: John Wiley & Sons,1900], 162)

177

Chapter 6

ittings. A portable rail-mounted winch was used to both raise and lower the curtain. he Caméré type was introduced at the trestle dams already in existence at Suresnes and Villez and at the bridge dams at Meulan, Méricourt, PortMort, and Poses, all on the lower Soâne.12 Also, one was erected at the outlet of Lake Geneva on the Oder River in Germany and in general they gave excellent satisfaction. he movable dam at Poses with a lit of 16.4 feet, measured from the sill at the deepest elevation on the pass, was an outstanding example of the use of a Caméré curtain because the friction was much less when compared to the Boulé design. he Caméré dam was capable of application with high lits. he roll-up curtain represented a controlled means for adjusting the low rate on the pass and/or weir. he irst cost for a Caméré curtain was higher than a Boulé system and required considerably more maintenance in operation. Figure 6.4. Caméré curtain dam. (Edward Wegmann, he Design and Construction of Dams [New York: John Wiley & Sons,1900], 164) 178

F r e n c h M o va b l e D a m s i n A m e r i c a

In order to compare the Caméré and Boulé methods, at the extant Suresnes Dam, Boulé gates and Caméré curtains were installed side by side as a practical comparative study.13 he study began in 1885. By 1905 the engineers concluded that overall the Boulé system was superior. he Caméré curtains operated on a saving of 25 percent in operating time. he curtains, however, were found to be fragile and required constant maintenance since they tended to collect debris between the wooden members.

Drum Wickets he drum wicket represents an ingenious invention in the family of movable dams. Louiche Desfontaines, associated with the Marne Navigation, irst introduced the drum dam in 1857. Like spokes in a wheel the “leaves” could be rotated to open or close the gate with water let into the sluices under pressure, causing the upper leaf to rotate, releasing the head of water.14 he reverse procedure raised the leaf, closing of the low. he advantages of the drum gate feature were few working parts and operation by water pressure, which the Victorians Figure 6.5. Cross section of Desfontaines’ drum dam on the Marne River, France. (Edward Wegmann, he Design and Construction of Dams [New York: John Wiley & Sons,1900], 185) 179

Chapter 6

would have called “self-acting.” Finished in 1867, the Joinville Lock and Dam was the irst constructed example of the drum dam. Further examples, particularly in Germany, impounded as much as 9.2 feet of water. In the last decade of the nineteenth century, Captain Hiram Chittenden, United States Army Corps of Engineers, introduced an American drum wicket. Apparently the irst application was on the Osage River in Missouri in 1911. A more notable example was the installation of such a structure at dam number two on the Monongahela River in West Virginia as part of the Monongahela Navigation, installed in 1905. It appears that the Chittenden drum dam was a simple device and generally easy to operate, but as a result of mud and a great deal of debris these wickets oten got out of adjustment, necessitating frequent maintenance. As a result, the Chittenden drum wicket did not ind wide application.

Shutter Dams Shutter dams are of ancient origin, as indicated above, in providing openings in mill dams for the passage of vessels, rats, and other loating objects. he traditional shutter, hinged at the sill, would lie lat in the open portion, providing the maximum width of open water. Figure 6.6. hénard shutter dam. (Edward Wegmann, he Design and Construction of Dams [New York: John Wiley & Sons,1900], 174)

180

F r e n c h M o va b l e D a m s i n A m e r i c a

(Top) Figure 6.7. Betwa movable dam used on the Monongahela River. (Pittsburgh District, U.S. Army Corps of Engineers Archives). (Above) Figure 6.8. Chanoine wickets and props of navigation pass, Dam No. 6, Ohio River. (B. F. homas and D. A. Watt, he Improvement of Rivers [New York: John Wiley & Sons, 1903], 237).

181

Chapter 6

(Top) Figure 6.9. Erection of Louisa needle dam ca. 1848. (B. F. homas and D. A. Watt, he Improvement of Rivers [New York: John Wiley & Sons, 1918], 220) (Above) Figure 6.10. Needle dam at Louisa, Kentucky. (B. F. homas and D. A. Watt, he Improvement of Rivers [New York: John Wiley & Sons, 1918], 219)

182

F r e n c h M o va b l e D a m s i n A m e r i c a

It appears that the irst movable shutters were described by Delalande in 1778 when such a device appeared on the Orb River in France. Ingénieur hénard was familiar with these hydraulic works on the Orb. He applied movable shutters at several locations on the l’Isle River where the height of retained water measured 6.56 feet above low water. he dams were insuicient for navigation, but proved efective for controlling freshets. To obviate the lack of capacity hénard reduced the sill elevation of the dam to just four feet with the rest of the lit being controlled by a movable shutter.15 he shutters were held in place by props. hénard devised a chain tripping system whereby successive shutters could be dropped by moving a rack-and-pinion tripping bar. Ater lowering, the shutter ended up lying lat on the lower sill. Raising the shutter against the current proved to be a diicult act at best. Mermoyer, a French engineer, recommended to hénard that a counter-shutter upstream from the main gate be installed so that the current of the river would operate the counter-shutter and relieve the water pressure on the main shutter. With the counter-shutter raised by the current the principal shutter could be raised by hand with little diicultly. To assist in raising the main shutter a walkway was oten installed on the top of the counter-shutter. During the lourishing period of the hénard shutter dam several variants were introduced. In 1868 Ingénieur Krantz presented details of his pontoon (ponton in French) in which the shutter could be raised by introducing air under pressure in the chamber under the shutter, forcing the shutter into a vertical position, with a brace to support it against water pressure. Ingénieur Girard invented a movable shutter dam employing a hydraulic jack. Probably the irst application was at the Brulée Dam at Auxerre on the Yonne River near its conluence with the Seine. his hydraulic jack system provided a simple solution to raising and lowering a shutter while at the same time the hydraulic piston acted as a prop supporting the shutter against water pressure. Perhaps the most simple of these early shutter dams is the Betwa Gate Dam in India.16 In operation it functions like a hénard counter-shutter with the restraining chain replaced by a solid iron tension strut. he only examples known in America were applications on the lower reaches of the

183

Chapter 6

Monongahela Navigation near Pittsburgh. A study is needed of the extensive use of movable dams in the irrigation schemes in India, built under the direction of the Royal Engineers. he movable gates were not used for navigation on the Betwa River in India. hénard dams were at Mahanuddee and Cossie Rivers in India.

Wicket Dams While still employed on shutter-dam improvements, Ingenieur hénard retired from the French Corps of Ingénieurs. His place was taken by Jacques Chanoine.17 Chanoine was responsible for a seminal improvement of the shutter dam, by raising the pivot point of the hénard shutter to more nearly match the center of water pressure, which would be a third of the depth up from the bottom. he raised point of rotation necessitated the construction of an iron or steel frame called a horse. he great advantage of the modiication enabled the wicket (also known as a shutter) to swing in a horizontal position when the water reached the top of the wicket. With the wicket “on the swing” the entire movable dam could be lowered onto the sill. he operation of the Chanoine wicket dam began when the wicket assumed the horizontal position. To lower the wicket the prop is pushed sideways to the open groove in the hurter. To raise an individual wicket the horse and prop are lited into position from a chain windlass mounted on the fermettes. hese frames, together with the iron horse, are hinged to fold lat on the sill, clearing the way for open navigation through the pass when the water levels in the river were adequate. An alternative oten employed in America is to omit the service bridge and windlass using, instead, a maneuver boat to raise and lower all of the ironwork on the dam. Ingénieur Pasqueau introduced an improved hurter that eliminated the usual tripping bar, which allowed each wicket to be maneuvered separately.18 he U.S. Army Corps of Engineers adopted the Pasqueau hurter in the wake of satisfactory applications. Because of the lexibility and oten warping encountered, the wickets were mounted on a four-foot width with a four-inch space between wickets, allowing them to operate without interference.

184

F r e n c h M o va b l e D a m s i n A m e r i c a

Conclusions Although William Emery Merrill, USACE, began the construction of the irst Chanoine wicket dam on the Ohio Navigation in 1885, it was on the Great Kanawha Navigation in West Virginia that the irst river was canalized using ten Chanoine wicket dams and two ixed dams. his canalization of a major tributary of the Ohio River was opened for traic in 1898. Because of sporadic funding and the sheer size of the project the Ohio River navigation had to wait until 1929 to be completed. In the case of the Great Kanawha Navigation, the lits ranged from 6¼ to 8½ feet. he canalization of the Great Kanawha River was a triumph for the Corps of Engineers and a resounding economic success. he modern roller-gate lock and dam at Winield is the busiest in the entire Corps of Engineers’ system. In the face of increased traic, mainly coal tonnage, the Winield and Marmet dams enjoy new 800-foot-long locks. he history of French movable dams in America can be assessed on two levels. First, they represent a major hydraulic engineering triumph. As a result the technical history is worthy of study and evaluation. he movable dam made possible the riverman’s dream of a comprehensive system which could function as a traditional slack-water as well as open navigation. With the movable dam lowered, open navigation was available on the entire Ohio canalized system, together with improvements on selected Ohio River tributaries. By any measure the development represents a golden age in hydraulic engineering in America. he improvement of waterways east of the Mississippi River is an outstanding example of engineering development of a new technology. Not isolated to its country of origin, the movable dam appeared in America, Canada, India, and elsewhere, based on French development. It is an outstanding example of technology transfer, rivaled by few other engineering works from the mid-nineteenth century to the twenty-irst century. Chapter 6 Notes 1

George Macaulay Trevelyan, History of England (London: Longmans, Green & Co. Ltd., 1926), 125.

185

Chapter 6

2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18

Emory L. Kemp, he Great Kanawha Navigation (Pittsburgh: Pittsburgh University Press, 2000), 25–27. Edward Wegmann, he Design and Construction of Dams (New York: John Wiley & Sons, 1900), 193. Benjamin F. homas and D. A. Watt, he Improvement of Rivers (New York: John Wiley & Sons, 1913), 559. Kemp, Great Kanawha Navigation, 28. Ibid., 47. Ibid., 30. Wegmann, Design and Construction of Dams, 158. homas and Watt, Improvement of Rivers, 583. Wegmann, Design and Construction of Dams, 163. Ibid. homas and Watt, Improvement of Rivers, 616. Ibid., 617–618. Ibid., 646. Wegmann, Design and Construction of Dams, 172. homas and Watt, Improvement of Rivers, 522. Ibid., 585–607. Ibid., 574.

186

7

Building the Tennessee-Tombigbee Waterway

S

Water Surface (acres)

9t

Pool Elevation (t)

300 t

Lit (t)

148 mi

Locks 110 t wide by 600 t long

Channel Depth

River

Channel Width

Section

Total Length

tretching 234 miles, the Tennessee-Tombigbee Waterway connects the Tennessee River with the Gulf of Mexico via the Black Warrior and Tombigbee rivers and, thence, to the Alabama docks at Mobile on the Gulf of Mexico. Completion of the waterway in 1985 represented the largest project ever undertaken by the U.S. Army Corps of Engineers. he following lock and dam details give an idea of the magnitude of the project:

(river below Gainesville L&D) Gainesville L&D

36

73 t

5,000

Aliceville L&D

27

109

6,400

Columbus L&D

27

136

8,300

Aberdeen L&D

27

163

4,121

190 Canal

Divide Total

44

300

12

40

280

12

232 *





Lock A

30

220

914

Lock B

25

245

2,718

Lock C

25

270

1.642

Lock D

30

300

1.992

Lock E

30

0

851

Bay Springs L&D

84

414

7,645

– 341



43,483

*Oicially 234 miles or 376.58 kilometers

187

Figure 7.1. he Tennessee-Tombigbee waterway in relationship to other Corps of Engineers waterways in the Midwest. ( Jefrey K. Stine, Mixing the Waters: Environment, Politics, and the Building of the Tennessee-Tombigbee Waterway [Akron:University of Akron Press, 1993])

he Corps of Engineers reported the waterway, when completed, connected with 16,000 miles of navigable inland canals and waterways. Born amid great controversy, the project took twelve years to complete at a cost of $1,992 billion. It was during French control of Louisiana in 1770 that there appeared what was apparently the irst mention of a possible waterway. Later, in 1874, under the administration of Ulysses S. Grant, Congress authorized a survey of the projected civil work. Despite continuing interest in the project during the irst half of the twentieth century, nothing was accomplished on the ground until the 1946 River and Harbor Act of the U.S. Congress. he work, however, languished until 1970, when a twelve-year period of construction began. he history of this magniicent public work could be dealt with in several ways. he approach taken here is to stress the engineering aspects of building the Tennessee-Tombigbee Waterway.

Planning and Design Designing a major public works project is an enormous task requiring the balancing of a complex mixture of challenges, some technical, some political. Designers of the Tennessee-Tombigbee Waterway took into consideration a variety of factors, including geography, geology, hydrology, river hydraulics, environmental considerations, the balancing of multi-purpose objectives (for example, building the waterway for both navigation and recreation), inancial limitations, time constraints, court orders, and political interjections. Such is the nature of large civil-engineering works. With the Tenn-Tom, however, the technical considerations were magniied by the sheer size of the endeavor, while its political considerations were exaggerated by the project’s celebrated controversy. At its most basic level, the design of the Tenn-Tom was most signiicantly inluenced by the lay of the land. In conceptualizing the project, Corps designers divided the waterway into three sections—river, canal, and divide—each relecting the engineering approach taken to overcome the distinct geographical characteristics found along the path of the project. he river section comprised the southern end of the waterway, where engineers confronted the challenge of making this section of the Tombigbee River 189

Chapter 7

navigable by constructing a series of four locks and dams and by straightening, widening, and deepening the natural river channel. he canal section paralleled the upper reaches of the Tombigbee River, which Corps engineers determined was too narrow and winding to provide an economical navigation channel. hey proposed to solve this problem by constructing an artiicial canal upland from the river and containing ive locks. he northern section of the waterway posed the greatest earthmoving challenge; for here the Corps had to cross the mountainous divide separating the Tennessee and Tombigbee watersheds. he Corps proposed building a high-lit lock and dam at the southern terminus, and digging a 27-mile-long cut through the divide. Designers of the Tenn-Tom faced numerous substantive engineering challenges. From a geotechnical standpoint, the foundation conditions presented one of the greatest problems throughout the waterway. Nine out of the ten locks had to be built on alluvial soils rather than on bedrock, and the abundance of groundwater—much of it under artesian pressure—made dewatering these construction sites (so that the project could be built in the dry) a major technical consideration.1 Environmental considerations—prescribed by legislation enacted subsequent to the completion of the overall design memorandum—presented designers with another set of technical problems. Could this project, which was conceived during an earlier time, still be built while meeting the new environmental criteria? he combination of recently passed National Environmental Policy Act (NEPA) and NEPA-based litigation aimed at stopping the TennTom worked to ensure the Corps incorporate an interdisciplinary approach in designing the waterway. Although time pressure was not a speciic design requirement, it challenged the designers, as well as most other professionals associated with the waterway. Like most intellectual and skilled tasks, time is a critical variable in producing the best possible product. Rapid construction scheduling constantly pushed Corps designers, causing them to work with assumptions they might otherwise have tested more thoroughly through modeling or additional sampling. he time pressure came from several sources, but one of the driving factors was Congress. Project supporters worked hard to obtain maximum construction funds for the Tenn-Tom, and at times were able to 190

B u i l d i n g t h e T e n n e s s e e -To m b i g b e e Wat e r way

secure appropriations exceeding Corps requests. Under such circumstances, the Corps felt obliged to spend those funds. It did not want to be accused of dragging its feet on the project.

Organization he actual planning and design of Corps projects takes place at the district level, under guidelines established by the Oice of the Chief of Engineers and under review of the appropriate division, Board of Engineers for Rivers and Harbors, and Oice of the Chief of Engineers (OCE). he geographical boundaries of Corps districts are deined by regional watersheds, and since most water-resources projects modify a particular watershed, they usually fall within the jurisdiction of a single district. he Tenn-Tom was diferent and therefore posed a unique organizational problem for the Corps. Its purpose was to join the waters of two diferent river systems and, by cutting through the geographical divide separating them it also cut through the jurisdictional divide separating the Nashville and Mobile districts. his bureaucratic anomaly, along with the magnitude of the project, led senior Corps oicials to consider establishing a special, temporary district assigned the sole task of building the waterway. hey also considered giving the entire project to the Mobile District where rested the largest portion of the project. Protection of bureaucratic turf, however, is almost instinctive among divisions within a government agency. his fact was not lost on the Corps’s leadership, who opted, in 1946, for a compromise solution: the Nashville District would design and construct the divide and canal sections, while the Mobile District would design and construct the river section in addition to serving as the lead oice for the entire project. Although only 25 miles of the 253-mile-long waterway fell within the geographical boundaries of the Nashville District, Nashville gained responsibility for some 85 miles of the project. his division of workload gave Nashville about 60 percent of the dollar amount of the Tenn-Tom.2 Dividing the projects between Nashville and Mobile created another organizational problem in that the two districts belonged to diferent divisions, the Ohio River Division and the South Atlantic Division, respectively. he Corps resolved this potential conlict by giving SAD the lead role, which included 191

Chapter 7

responsibility for preparing the budget and justiication data and testifying before Congress. In 1967, the division of responsibility was reapportioned to resemble more closely the geographical boundaries of the two districts. Under the revised division, Nashville held responsibility for designing and constructing the divide section, including the Bay Springs Lock and Dam, while Mobile was responsible for the remainder of the project, including the canal section. he districts’ existing workloads and capabilities prompted this rearrangement. Overall responsibility for the waterway remained with Mobile, including accountability for all funds, maintenance of cost records, and issuance of inancial reports. Primary responsibility of planning and construction also rested with Mobile. Coordination between Mobile and Nashville and the South Atlantic and Ohio River divisions was a critical element in this organization.3 Design of the Tenn-Tom took place in many stages and over a period of years. Nearly all the design work was done “in house” by Corps personnel at the Mobile and Nashville districts, whose work was then reviewed by oicials at the South Atlantic Division and the Oice of the Chief of Engineers. (Construction, on the other hand, was undertaken entirely by private contractors under Corps supervision.) Major design concepts were usually explored in depth at conferences attended by Corps personnel from the districts, divisions, and headquarters. he Corps’s capability of designing a large and complex public-works project is unusual within the federal government; most federal agencies responsible for major engineering works must contract out much of the design work.4 he basic engineering concept of the project was presented in Design Memorandum No.1, labeled “General Design,” which the Mobile District completed in June 1960 and which the Chief of Engineers approved in April 1962. By focusing on the overall engineering concept of the waterway, this document contained little detailed design. his general design memorandum, or GDM, dealt with such broad design concepts as how many locks and dams the waterway would have, the general location of those navigation structures, rough channel alignments, and the like. he major segments of the waterway—such as the locks and dams—then received their own individual general design memorandum. GDMs focused on determining the deinite location and orientation of navigation structures and 192

B u i l d i n g t h e T e n n e s s e e -To m b i g b e e Wat e r way

such things as the number and dimensions of spillway gates. Alternative designs were evaluated at this stage to determine the soundest approach from engineering, economic, and environmental perspectives. At this stage of the design process, pertinent state and federal agencies were provided general information on the project item under consideration and were asked to provide their views. Corps designers would then take into consideration any existing or potential problems raised in the agency comments. Speciic agency recommendations would also be considered. hese agency views were both summarized in the printed design document and printed verbatim in the appendices. At this general design stage, preliminary plans for recreation were developed. Real-estate requirements were also determined at the same time. Realestate considerations included acreage estimates for the following: (1) acres in fee for the construction area, operational areas, and public access areas; and (2) acres of perpetual easements for the pool created by the dam, channel improvements, and dredged material disposal areas. In addition, Corps designers would determine the need for utility, road, and railroad relocations. Further reinement and speciicity of the design for the various segments of each GDM led to a feature design memorandum, or FDM. FDMs were written for such items as locks, dams, spillways, and navigation channels. “Plans and Speciications” followed, and provided the level of detail needed by the construction irms to bid on the project (and to construct it). In designing the irst structure to be built on the waterway—the Gainesville Lock and Dam at the southern end of the river section—engineers at the Mobile District relied heavily on earlier design work. he process was standard, following the traditional pattern of preparing the design, sending it to the division oice in Atlanta for comment, addressing comments and resubmitting it, ater which it was submitted to headquarters for inal approval. It was a time-consuming process that generated a lot of correspondence. It became apparent to members of the Mobile District that a large and complex project like the Tenn-Tom, with its tight deadlines, required a streamlined system of review. his became even more imperative as the Corps sped up its construction schedule in the mid-l970s, which meant that design also had to be quickened. Consequently, the Corps modiied its traditional design process. Mobile and Nashville submitted their drat designs to the South Atlantic Division with the 193

Chapter 7

understanding that the division staf would have one month to review and comment on the drat. At the end of the month, members of the division would meet with district personnel to resolve any disagreements. Diferences were settled in person and on the spot. his greatly sped up the process. Ater the meeting, district engineers could return to work on the design, knowing that the issues had been resolved.

Surveys and Lab Work he preparation of feature design memoranda for individual sections of the waterway required data obtained through extensive surface and subsurface surveys. Core borings were made at strategic locations and the resulting soil samples sent to the South Atlantic Division Laboratory for analysis. his testing of materials represented a small but essential step in the design process, since the Figure 7.2. Civil works in the Mobile District. (U.S. Army Corps of Engineers)

194

B u i l d i n g t h e T e n n e s s e e -To m b i g b e e Wat e r way

tests determined such things as the shear and compressive strength and permeability of the underlying foundations. he SAD Laboratory took entire responsibility for the Tenn-Tom testing: the main laboratory in Marietta, Georgia, tested over 30,000 soil samples from the Tenn-Tom, while the temporary ield laboratory in Fulton, Mississippi, conducted over 75,000 additional soil-sample tests during its two-year lifespan.5 Part of designing the navigation structures was locating sources of construction materials. Tests of sands and gravels allowed the Corps’s laboratory to specify the type of cement to be used for concrete mixtures. Laboratory tests also provided information on the availability of acceptable slope-protection stone from local sources. Economic considerations played a signiicant role in this process. If local construction materials did not meet Corps standards, then suitable materials would have to be transported from outside the area—thus raising costs. By shiting concrete speciications and other design criteria, it was oten possible to reach compromises allowing contractors to compensate for inferior local materials.6 Figure 7.3. he divide cut forms the link between the Tennessee Valley Navigation and the Tenn-Tom Waterway. (U.S. Army Corps of Engineers)

195

Chapter 7

he Corps’s Hydraulic Design Branch concerned itself with a variety of issues, not the least of which involved how cross currents near canal entrances might afect the entry of tows. Model tests were oten called for, and from the results, design modiications could be made to lessen or eliminate adverse currents.7 Designing the project necessarily preceded building it, and as the waterway drew increased criticism from various quarters, congressional supporters of the Tenn-Tom pressured the Corps to build faster. his in turn placed the Corps’s designers under tight deadlines. Designers asked those technicians performing lab work or model studies to accelerate their pace. As the pressure mounted, designers cut back on the amount of studies they normally would have requested. Hydraulic models for one lock and dam suiced for others. Corps designers were told in December 1973, for example, that “irst priority will be given to expediting the model studies in order to meet the project schedules.”8 While the increased pace of construction placed greater burdens on Corps designers, those same designers felt the pinch of time constraints. When Congress appropriated the initial construction funds in 1970, only the general design concepts for the waterway had been prepared. In order to request construction bids and proceed with building the project, detailed feature designs had to be prepared. When the Environmental Defense Fund sued the Corps in 1971 and won a temporary injunction against the Tenn-Tom, ultimately postponing the start of construction by a year, designers at Mobile privately welcomed the situation, since it allowed them time to collect needed data and designs for Gainesville and the surrounding navigation channel.9 he press to design quickly necessitated abandonment of certain routine procedures. One of the changes was to reduce the extent of surveying, sampling, testing, and monitoring, and instead to rely more fully on “normal assumptions.” Since these assumptions were generally conservative and based on broad experience, Corps designers moved ahead with adequate conidence. Yet without testing these assumptions—assumptions on lateral earth pressure, backill pressures, and uplit pressure, for example—reinements in the inal designs were limited.10 (he degree of acceptable uncertainty was increased.) At the design conference on Bay Springs Lock held in June 1976, Corps designers learned that “the completion schedule presented to Congress by the 196

B u i l d i n g t h e T e n n e s s e e -To m b i g b e e Wat e r way

Division Engineer assumes the proposed annual funding and early award of the Bay Springs project. Any delay caused by additional modeling and/or similar items would impact seriously on the completion date and optimum utilization of annual funding.” he engineers debated, among other things, the inal design of the hydraulic illing and emptying system, and although the Waterways Experiment station at Vicksburg, Mississippi, had already completed some model testing, certain further design modiications might necessitate additional model testing—a requirement that would take six months to complete and two months lead time to initiate.11

Instrumentation Corps policy required district oices to monitor and evaluate the behavior of civil structures both during construction and throughout their operation. his continued surveillance demanded the installation of instrumentation capable of measuring settlement, delections, movement, soil and pore pressure, and uplit pressure. Control monuments were installed on the structures and in the surrounding vicinity (usually anchored to rock) and triangulation surveys were made using laser distance instruments—thus determining set distances and any movement through periodical remeasurements. Alignment inserts placed in each monolith joint allowed technicians to monitor alignment and elevation. Uplit pressure cells and soil and pore pressure cells were installed. Concrete strain gauges monitor strains within the concrete monoliths and monitored cracking. Piezometers were widely used to measure artesian groundwater conditions. Corps designers indicated a schedule for periodic reading of the various instruments.12

Environmental Issues Objections by environmental critics in 1970 and 1971, followed by the NEPA-based litigation to stop the Tenn-Tom, forced Corps planners and engineers to consider seriously the environmental implications of the waterway and to work to minimize the project’s adverse efects. Mobile’s creation of a board of consultants for environmental concern relected that attention to environmental 197

Chapter 7

issues. Board members worked diligently in assisting the Corps with the litigation and helped the agency by advising and critiquing engineering designs and environmental reports. In April 1973, board members expressed their concern that the favorable court opinion liting the injunction against the project—an opinion, they argued, that was “based heavily on the environmental consideration indicated by the Corps and its consulting staf ”—had fostered “an apparent decrease in interest and support for these environmental concerns as the Tenn-Tom project moves into a period of construction.” hey complained that Mobile District had decreased its “ ‘in-house’ manpower and time spent on this aspect of the project.” Moreover, they argued, “there is a dwindling of support for contract studies related to environmental concerns, and a shit in attitudes of some of the Corps staf.”13 In its “First Supplemental Environmental Report,” dated August 1975, the Mobile District stated: “he cover of set of plans and speciications for construction of the waterway prominently displays the following statement, ‘Good Engineering Results in a Better Environment.’ his statement has become the motto of the Corps in designing and constructing the waterway.”14 he Board of Consultants for the Tennessee-Tombigbee Waterway environmental study remained active throughout the project design and construction. It continued to monitor and review the Corps’s environmental reports, and it commented on various design documents. During both court cases, the board played a signiicant role in assisting the Corps and the Department of Justice as advisors and expert witnesses. Daniel T. Nelson, the board’s ecologist member, died unexpectedly of a heart attack in August 1975. His replacement was Stanley I. Auerbach, Division Director, Environmental Sciences Division, Oak Ridge National Laboratory. Board members continued to meet quarterly to discuss various aspects of the waterway’s continuing design, the efects of construction, and other items deemed worthy of their attention. Spoil-disposal areas, revegation, and groundwater were special interests of the board. Following its inspection of the waterway in fall 1979, the board expressed its dismay with the inishing work around Gainesville Lock and Dam. “While it is still not understood by the Board why formed land around locks and dams have to have harsh geometric shapes,” they reported, “even so appropriate landscape architecture could at least have 198

B u i l d i n g t h e T e n n e s s e e -To m b i g b e e Wat e r way

sotened man’s intrusion into nature’s sphere.” heir recommendation: “Enough janitorial engineering! Develop a landscape plan with purpose and meaning. Buy plants by the thousands from the forestry service and ask children in the area to help plant them to clothe the embarrassing nudity of nature found in manmade forms.”15

Project Dimension Corps oicials held a conference in Mobile in September 1972 to address various design issues of the waterway. he channel commanded the most attention, as Corps designers debated the merits of constructing an initial channel width of 300 feet versus an initial width of 200 feet that could later be widened to 300 feet. With regard to the latter scenario, two plans were to be considered: “Plan 1—Completion of a 200-foot channel in 1983; commence widening in 1990 and complete in 1994” and “Plan 2—Completion of a 200-foot channel in 1983; commence widening in 2012 and complete in 2016.” Mobile District Engineer Colonel Harry A. Griith noted in March 1973 that “plan 1 is little more than a continuation of construction over a longer period of time” and that both plans would result in a reduction of beneits because of traic slowdowns during construction. In addition, later widening would incur the additional costs of replacing riprap and drainage structures, seeding the new cut slopes and disposal areas, and the higher unit price of excavation due to the diferent type of excavating machinery used in the widening versus that used in the initial construction. Griith therefore recommended to the South Atlantic Division that consideration of the two-stage development of the Tenn-Tom be discontinued.16 Channel alignment received a great deal of attention during the design phase. he Tombigbee River meanders, and its switchbacks become tighter and more numerous as you progress upstream. Following the natural river course demanded a great deal of dredging. An alternative was to excavate river cutofs to shorten the navigation route and reduce the amount of required excavation. Another factor involved bendway design criteria. In order to provide safe and eicient passage of barges, the Corps maintained guidelines calling for deeper and wider sections where the navigation channel curved. Straightening the channel would reduce the required amount of bendway 199

Chapter 7

widening. (For example, a nine-foot-deep channel was to be excavated to twelve feet in a bend.) In April 1975, Mobile District Engineer Colonel Drake Wilson responded to the request of Louisville and Nashville Railroad’s Leo Koster to explain the design changes in the channel depths of the waterway. he originally authorized plan for the Tenn-Tom provided for a nine-foot-deep navigation channel throughout the project. By 1975, Corps designers were calling for large sections of the waterway to include twelve-foot depths. Wilson explained, Model studies and experience have proven that, in restricted low areas such as in cut-ofs and in canals, navigation interests experience considerable drag on tows due to the close proximity of the channel boundaries. his phenomenon is experienced when the vessel takes up a relatively large portion of the low area and would be evident in the upper reaches of the River Section and in the Canal and Divide cut sections with depths of 9 feet. In order to reduce this impediment to navigation, the controlling depth has been increased from 9 to 12 feet in the Canal and Divide Cut Sections, in all cut-ofs, and in the Columbus and Aberdeen Lakes where excavation is required.17

Wilson said that the nine-foot depth would be maintained in the wider sections of the river section where the “drag problem” did not exist. In addition, he said that “all navigation locks have been designed to allow for possible future 12-foot project depths.” his modiication was made, he added, “so that the increased depth can be economically provided in the future if traic and tow sizes warrant the dredging expense to provide larger channels.”18 All ten locks on the waterway were of standard Corps design with chamber dimensions of 110 by 600 feet. With only two exceptions—Lock D and Bay Springs Lock—the locks were of the gravity-wall design. his meant that the structural strength of the lock—notably its ability to resist sliding and overturning—came from the gravity weight of the massive concrete walls. Each of the two lock walls at every structure acted as an independent monolith.19 he locks were designed so that they could be operated by a single person. Control booths were placed high above each lock and a closed-circuit television

200

B u i l d i n g t h e T e n n e s s e e -To m b i g b e e Wat e r way

system was installed to provide the lockmaster with views of all areas around the lock gates and spillways. Intake and discharge areas are also monitored by the camera to enable the lockmasters to ensure that no small boats are endangered by the opening of the miter gates. Nearly all lock operations are conducted from the control booth through electrical push-button controls.

River Section In the river section of the Tenn-Tom, the surrounding countryside is characterized by broad, lat pastureland, broken by gently sloping hills. he Tombigbee River is wider here, rarely narrowing to less than 200 feet. he river meanders through this area, at places creating blufs 20 to 30 feet high, but it is usually bordered by low riverbanks. he historic meandering has created a broad loodplain stretching one to two miles on both banks starting at the mouth of the Warrior River near Demopolis, Alabama. he river section extends north on the Tombigbee, widening, deepening, and straightening some 148 miles of the river.20 Four conventional locks and dams were to be constructed in the section. All four dams were to be earthill structures with concrete-gravity, gated spillways. he Gainesville Lock had a lit of 36 feet, while the locks at Aliceville, Columbus, and Aberdeen all had lits of 27 feet, for a total lit of 117 feet. he channel dimensions in the river section were to be 12 by 300 feet. Dredging would be required throughout most of the river section to provide the desired width and depth of the channel. Some widening of curves would be required in the lower reaches of this section, while the upper reaches would require major rectiication. his work would include the creation of several cutof channels. Canalization of the Tombigbee River in the river section relected the traditional Corps of Engineers approach to navigation projects. Little bank protection was provided in the river section. Contractors dredged the speciied width of the channel and the banks were let to ind their own angle of repose. If erosion problems later developed, the Corps would then return to ix them—this in lieu of providing bank protection throughout the waterway.21

201

Chapter 7

Disposal Areas Aside from converting the free-lowing Tombigbee River into a series of impoundments, the most signiicant environmental alteration in the river section involved the disposal of dredged and excavated materials. he Corps’s traditional method of disposa1 involved spreading the dredged material (typically about 80 percent “carriage” water and 20 percent solid material) along the banks of the river. he pumped water would then wash back into the river along with a certain amount of suspended material (signiicantly increasing the turbidity downstream), while the heavier dredged material remained on or near the riverbanks. Oten, the dredged material was used to ill a swamp or slough, but the wetlands represented some of the most important ecological sites in the river valley. Destruction of wetlands and bottomland hardwoods adversely afected the river’s water quality and robbed the region of prime ish and wildlife habitat. It also threatened archaeological sites and the area’s aesthetic qualities. Moreover, this traditional method of dredged material disposal eventually created maintenance problems as the solid materials eroded back into the river. he tremendous amount of dredging required in this section made the traditional approach all the more environmentally unacceptable.22 Corps planners had not developed a speciic plan for discharging this material by the time of the original environmental impact statement, but they promised to devote signiicant attention to determining the optimal methods for disposal.23 As required by NEPA, the Corps approached this problem in an interdisciplinary fashion by assigning its own engineering and environmental staf to the task and by consulting with its own Board of Environmental Consultants and representatives from the U.S. Fish and Wildlife Service, the Environmental Protection Agency, and state ish and game agencies. he Corps addressed this problem by developing new design concepts for the disposal of dredged material. Rather than dumping the material along the riverbanks, it was pumped away from the river to concentrated disposal sites. A continuous dike, typically nine feet tall and ten feet wide at the crest and built from material borrowed from the interior of the disposal site, encircled each site. he sites were internally divided by a transverse dike into two cells. 202

B u i l d i n g t h e T e n n e s s e e -To m b i g b e e Wat e r way

he primary cell—the larger of the two—was designed to accommodate all the materials excavated during construction, while the secondary cell provided storage space for ity years of maintenance dredging. Contractors pumped dredged material into one end of the primary cell, where it was allowed to stand—or “pond”—before it overlowed at the other end into the secondary cell. During the time the water stood in the primary cell, the coarser suspended materials settled, which meant the primary cell received the bulk of the solid material. As the water let the primary cell and entered the secondary cell, it still contained suspended material—usually of a much iner nature. he water was allowed to pond in this secondary cell and release more of its suspended material before it was then passed back into the river. his method substantially improved the quality of the water reentering the river.24 Dredged-material disposal generated widespread concern on aesthetic grounds because of the initial barren, scarred appearance it let. Corps designers addressed this problem by leaving an uncleared bufer zone around the diked disposal sites. he bufer zones, while not mentioned in the environmental impact statement (EIS), were discussed in Mobile’s First “Supplemental Environmental Report” for the Tenn-Tom, although its dimensions were not stated.25 Initially, Corps planners considered a bufer zone of 500 feet between the disposal sites and the river, but that was subsequently scaled down to 300 feet for channel frontage and 150 feet for the other sides—dimensions approved by EPA, FWS, and state agencies from Mississippi and Alabama.26 he bufer zones provided visual screens for the disposal areas and also served as seed sources for the natural re-vegetation of the sites. he interiors of the disposal sites were initially designed to have all trees and brush cleared, since the dredged materials would eventually kill most trees let standing. By the mid-1970s, however, biologists working for Mobile District argued successfully to limit the clearing of trees within the disposal areas in order to retain wildlife habitat and accelerate re-vegetation. he decaying trees would help rebuild the soil and would provide valuable habitat for a variety of arboreal animals and birds. Moreover, the design change ofered signiicant cost savings, as the Corps could forgo contracts for land clearing.27 he interdisciplinary approach mandated by NEPA found its fullest expression in the location of the disposal areas in the river section. Mobile’s 203

Chapter 7

Engineering Division formed an interdisciplinary team to select the site locations and design the dike alignments. he team consisted of a hydraulic engineer (Kenneth Underwood), an ecologist ( Jack Mallory), an agronomist/landscape architect (Bud Horter), and a real-estate specialist (Bill Williams). Ater 1975, Jerry Nielsen joined the team as an archaeologist. he small group worked together throughout the design of the river section, thus maintaining continuity. Occasionally, the team was joined in its ieldwork by a Corps surveyor and/or a representative from the U.S. Fish and Wildlife Service. Ater examining the land and studying maps and aerial photographs, the team reined the disposal area alignments to produce—in Underwood’s words—“what we felt like was the least environmentally damaging disposal area.” hey attempted to avoid destroying signiicant environmental and/or cultural-resources sites and tried to keep costs down by considering property lines and keeping real-estate Figure 7.4. Aerial view of the waterway showing the Aberdeen Lock and Dam together with two new bridges. (U.S. Army Corps of Engineers)

204

B u i l d i n g t h e T e n n e s s e e -To m b i g b e e Wat e r way

acquisitions to a minimum through minor realignments. hese considerations were made in light of engineering requirements and design guidelines, so the inal design was not compromised from the engineers’ standpoint.28 For Underwood and the other professionals at Mobile, they had little experience working together on an interdisciplinary basis prior to the Tenn-Tom. Each specialist tended to work independently, and there was little cross-fertilization of ideas in the design process. NEPA and the other environmental legislation passed during the early 1970s demanded otherwise, and for the Mobile District, Tenn-Tom served as the project to bring the various disciplines together, thus helping to change their approach to engineering design.29 Location of the various disposal sites became an important consideration for environmental, engineering, and economic reasons. Many of the lands surrounding the river held high environmental values; wetlands and bottomland hardwood forests, for example, provided prime wildlife habitat. To help reduce such impact, Corps engineers worked with their own biologists and biologists with other state and federal agencies when choosing sites. Economically, the further the dredged material had to be pumped, the higher the costs. Where possible, therefore, Corps designers tried to keep the pumping distances under one-half mile. Corps designers also tried to avoid siting the disposal areas in developed or settled areas—for example, they tried not to use productive farmland. Where practical, Corps designers located the disposal sites on the low bank or cut bank. hey also tried to avoid blocking natural drainage patterns with the disposal sites, yet, at the same time, they attempted to utilize the existing drainage patterns to manage the return of the pumped water back into the river. In all, there were over 100 disposal areas in the river section, many of them covering 150 to 180 acres.30

Cutofs Like many old rivers in the Southeast, the Tombigbee meandered in a serpentine fashion. Boats traveling this section of the river oten had to follow two miles of river channel to advance one mile. To shorten the length of the navigation channel in the river section, and to reduce the number of curves (which required

205

Chapter 7

additional widening for safety and ease of navigation), the Corps planned to construct several “cutofs,” or by-pass canals excavated across the necks of river loops. he mileage saved varied on each cutof, as did the size of the abandoned river loops created by the excavation. hese severed river loops, known as bendways, were analogous to oxbow lakes. Since barge traic would be diverted to the cutof channel, the Corps initially planned to allow these severed loops to silt over gradually. Because they remained connected to the navigational channel, their shallow backwaters and broad standing timber areas provided important ish and wildlife habitats, as well as recreational boating and ishing areas free of commercial barge traic. heir long-term existence was doubtful, however, as the river’s suspended sedimentation load would eventually ill in the ends of these loops, thus creating true oxbow lakes. his would occur because the main volume of water, along with its scouring properties, would low through the main channel, while the water that passed through the bendways would move much more slowly, causing it to deposit a portion of its suspended sediments.31 Elimination of these stretches of the river threatened ish populations in the Tombigbee River, and the U.S. Fish and Wildlife Service and several environmental groups pressured the Corps to preserve the severed sections of the river.32 By September 1981, the ongoing litigation, the question of endangered species, and the requirement to prepare a supplemental environmental impact statement convinced senior oicials at Mobile of the need for further studies of channel cutofs. he problems were several: Which bendways should be maintained for continued low from the natural river? What was the most practical engineering solution for providing the continued low? What would this efort cost, and how should the Corps budget for it? To answer these questions, Mobile’s Engineering Division Chief, James R. Covey, proposed creating an interdisciplinary “Cutof Study Task Force” comprised of Corps specialists in water quality, ish and wildlife, navigation channel design, sediment transport, operations, and recreation.33 Mobile’s leadership concurred, and a nine-member Bendway Task Force gathered for its irst meeting in mid-September. One of the principal concerns involved maintenance of the bendways, especially how much dredging would be required to keep them open to continuous river low. he task force sought to study the issue and recommend a management policy for bendways within 206

B u i l d i n g t h e T e n n e s s e e -To m b i g b e e Wat e r way

the Mobile District.34 Of the thirty-ive severed bendways, the task force identiied twelve requiring structural management measures in the river section.35 he Rattlesnake Bend Cutof in the southern section of the waterway between Demopolis and Gainesville represented one of the earliest, and one of the largest, cutofs in the waterway. Located about forty miles downstream from Gainesville Lock and Dam, the 6,000-foot-long Rattlesnake Bend Cutof eliminated about 10 miles of meandering river and created a 2,000-acre island surrounded by an oxbow lake. he Alabama Department of Conservation later managed this island as a wildlife preserve. In creating this major realignment of the river, the Corps disposed of some 2.5 million cubic yards of excavated material.36

Gainesville Lock and Dam Gainesville Lock and Dam was the irst navigational structure to be built on the Tenn-Tom. he construction site was a 250-foot-wide section of the Tombigbee River bordered by 30-foot-high blufs some 53 miles upriver from Demopolis Lock and Dam and about one mile northeast of the town of Gainesville in extreme west-central Alabama. An earth dam across the old river channel created a reservoir covering 7,200 acres and extending 45 miles upriver to the Aliceville Lock and Dam site. To reduce the navigational distance by about 4.23 miles, the Corps excavated a 9,200-foot-long canal through a large meander near the dam site. he 36-foot-lit lock sat in the downstream end of that cutof canal.37 he Federal Water Pollution Control Administration (FWPCA), the predecessor of the Environmental Protection Agency, reviewed Mobile’s drat design memorandum for Gainesville Lock and Dam in 1968. Regional Director John R. homan expressed the FWPCA’s concern about the possible efects of the structure on water quality, especially in lowering the levels of dissolved oxygen. He recommended the Corps install “re-aeration devices . . . to raise the dissolved oxygen to as near saturation value as practicable.”38 Although Mobile followed agency guidelines by publishing homan’s letter in its inal general design memorandum in January 1969, Corps designers dismissed his recommendation, stating: “Since water quality is not an authorized purpose of the TennesseeTombigbee Waterway and no tangible beneits were furnished to establish the 207

Chapter 7

economic justiication of such improvements, there is no basis for the inclusion of reaeration devices as part of the features proposed by FWPCA.”39 he subsequent passage of the National Environmental Policy Act and the mounting environmental criticism of the Tenn-Tom led Corps planners to reverse themselves on the Gainesville water-quality issue. As recommended by the FWPCA, the spillway at Gainesville (located adjacent to the lock in the excavated channel) received special attention, as it remained the major structural element afecting water quality. Water impounded behind the Gainesville Dam would become deicient in dissolved oxygen, and the Corps’s original spillway design—which consisted of a ixed-crest spillway that passed water over a smooth sheet of concrete—did little to improve the quality of water leaving the reservoir. In order to increase dissolved oxygen in the water downstream of Gainesville, Corps personnel redesigned the project’s minimum low ixedcrest spillway to include a lip—or “reaeration ramp”—about eight feet below the spillway crest that would cause the discharging water to fall free nearly thirty feet before it hit a delector bucket and splashed into the air. By maximizing turbulence and air-to-water interface in this way, reaeration was greatly increased. As Mobile’s planners happily discovered, the incorporation of these “environmental aspects” at Gainesville did not “increase the cost of the project.”40

Aliceville Lock and Dam he Aliceville Lock and Dam comprised the second navigational structure to be built on the Tenn-Tom. It was located on the Tombigbee River approximately one mile southwest of Pickensville in west-central Alabama, about forty miles west of Tuscaloosa, Alabama, and twenty-three miles southeast of Columbus, Mississippi. he lock and dam were connected by a concrete abutment wall. he dam had a combination gated and ixed-crest spillway: the gated spillway spanned the natural river channel and was topped by ive tainter gates (each 60 feet wide and 26 feet high) to control water low, while the 200-foot-long ixedcrest spillway extended from the gated spillway to the right abutment and sat 6 inches below the upper pool level.41 Engineers at the Mobile District requested that the Corps’s Waterways Experiment Station (WES) assist in the design of the structure at Aliceville. 208

B u i l d i n g t h e T e n n e s s e e -To m b i g b e e Wat e r way

he Hydraulics Laboratory at WES built a model of the proposed structure and the river in the Aliceville vicinity in order to run hydraulic tests. Based upon these modeling investigations, several design changes were recommended. hese included modifying the excavation of the let bank of the approach channel and the right bank of the bend near the upper end of the approach channel to improve navigational conditions in the upper approach to the lock, and improving navigational conditions below the lock by placing a dike on the right side of the approach channel.42

Columbus Lock and Dam Columbus, Mississippi, was the site of the third navigation structure on the Tenn-Tom. he Corps located the lock within a 5,700-foot-long canal to be excavated through the east river bank. To provide the old river channel a continuous low of water, designers incorporated a 200-foot-long ixed-crest spillway across the main river channel. Within the excavated canal, the lock joined a 416-foot-long gated spillway, which was controlled by six 60-by-26-foot-long tainter gates. As originally designed, Columbus Dam threatened Plymouth Bluf, a regionally renowned Cretaceous fossil bed overlooking the Tombigbee River about seven miles northeast of Columbus, Mississippi, with a twenty-nine-foot rise in the low-water level. Such a rise in the water level would inundate the bench at Plymouth Bluf and cover more than half of the exposed Eutaw Formation and two fossiliferous strata. he Corps described this impact in its environmental impact statement: “he bluf would lose its attraction for visiting study groups from other areas as well as most of its value as a teaching-research site for local universities and its natural beauty and recreational usefulness would be drastically reduced.” Having stated this, the Corps promised to evaluate design alternatives to avoid harming this resource.43 Subsurface investigations for the Columbus Lock and Dam undertaken in August 1971, however, were conined to the original site near Plymouth Bluf. he proposed site, which was near a campsite belonging to the Mississippi State College for Women, was found to be underlain by the Eutaw Formation, which produced excellent foundation conditions.44 Yet Corps oicials felt a strong 209

Chapter 7

need to demonstrate their agency’s environmental sensitivity and commitment to complying with NEPA. Plymouth Bluf, the most popular geological formation within the entire project area, seemed eminently worth saving—both for its own sake as well as for its public-relations value. And since the feature design for Columbus Lock and Dam had yet to be prepared, resiting the project would cost the Corps little from either an engineering or an economic standpoint. By November 1971, Mobile District designers had resited the lock and dam 3,800 feet upstream, thus sparing the unique paleontological site from looding.45 Even before the design change was formally adopted, Corps oicials used the resiting of Columbus Lock and Dam to refute the growing chorus of environmental criticism directed at the waterway.46 When the Environmental Defense Fund iled its lawsuit against the Corps in July, Daniel Nelson, one of the three members of Mobile’s Tenn-Tom environmental advisory board, recommended that the defense attorneys “make the most out of saving Plymouth Bluf. Use this as an example of modern Corps policy.”47 he relocation of the Columbus Lock and Dam was really not as signiicant from an engineering and design point of view as many of the waterway polemicists implied. Plans for the lock and dam were in a preliminary stage in the early 1970s, and resiting the structure was a relatively easy task. he Aliceville Lock, for example, was designed before the structure at Columbus, and as late as March 1971 Corps planner James D. Wall could report that “here is no good foundation data yet available and the inally selected site could be moved perhaps a mile or two if better foundation conditions would obtain.”48

Slurry Trenches Groundwater conditions posed a signiicant problem for engineers working on the Tenn-Tom because the waterway contained so many major navigation structures in the highly pervious soils of the Tombigbee River loodplain. Initially, the engineers followed the traditional approach for dewatering, calling for the contractors to intercept the groundwater and pump what remained (or what later entered) from the protected construction site. his design incorporated a series of dewatering wells encircling the construction site; and since these systems had to be maintained and operated 210

B u i l d i n g t h e T e n n e s s e e -To m b i g b e e Wat e r way

throughout the three- to four-year construction period, dewatering became a major construction cost item. he long periods of construction required at the various projects on the Tenn-Tom favored a system that would actually prevent groundwater from entering the sites.49 he high costs associated with conventional methods of dewatering construction sites led Corps designers to consider alternative solutions. he relatively new concept of earth- illed slurry trench construction drew their attention. he use of slurry methods as an engineering solution for underground construction was developed primarily in Italy and France. Prior to World War II, the use of slurry methods extended only to water-well and oil-well drilling operations. Gradually, during the 1950s and 1960s, the slurry method found use in underground construction in urban areas—primarily with large building foundations, subways, underground parking facilities, and large underground utilities. Its use in large water projects for construction-site dewatering, however, had gained only minor attention in the United States by the early 1970s.50 Gainesville Lock and Dam became the irst structure on the Tenn-Tom to receive a detailed design. In March 1970, Mobile District Engineer Colonel Robert E. Snetzer reported his staf ’s studies of ways to prevent “underseepage” at the Gainesville lock. Slurry trenches represented one of the alternatives considered. In this report, concern was not with keeping seepage out of the construction site while the structure was being built, but with installing a permanent barrier to prevent seepage along the upper and lower lock approaches. he considered plan involved utilizing two 500-foot slurry trenches connected to the guard walls and extending upstream. Mobile discarded the plan, however, “because of the probability of undesirable settlement occurring in the overlying ills.”51 In the spring of 1971, Corps oicials undertook “alternative studies for Gainesville Dam” as a means of reducing costs, incorporating environmental considerations, and ensuring safety. George H. Mittendorf, chief of SAD’s Engineering Division, reported in May that “based upon costs of installing the slurry- pile wall at Millers Ferry, it may be possible to reduce the cost of the cutof walls by about 35%.”52 However, at a summary conference at SAD on the Gainesville Dam held in July, the impervious-core trench received greater enthusiasm than did the slurry-pile wall.53 211

Chapter 7

he slurry-trench method of construction was tried as an alternative and was found to be both eicient and cost-efective. In October 1975, Mobile District programmers estimated that $4,379,500 had been saved through the employment of the slurry-trench method at the Gainesville, Aliceville, and Columbus construction sites.54 As the Corps explained to the Army Audit Agency in January 1976, “Very little specialized equipment is required, construction is relatively fast and allows the partial utilization of on-site materials without extensive skilled labor.”55 At this point the Corps was in the process of adding the slurry-trench concept to the remaining general design memoranda. Large backhoes and draglines dug three-foot-wide trenches down to the impervious Eutaw clay (usually thirty to forty feet deep, but at some locations extending down seventy feet). Contractors then mixed the excavated material with natural clay, sand, gravel, and bentonite slurry (bentonite is a gray Figure 7.5. Aerial view of Aliceville lock and dam. (U.S. Army Corps of Engineers)

212

B u i l d i n g t h e T e n n e s s e e -To m b i g b e e Wat e r way

powder made from natural clays containing montmorillonite) and pushed it back into the trench to displace the slurry and to form an impermeable barrier. he displaced slurry was then reprocessed and reused as the trenching process progressed. he bentonite slurry is similar to the material used by water- and oilwell drillers to prevent their holes from collapsing. he process proved far less costly than the traditional methods of dewatering, but designers and contractors alike held reservations about its efectiveness since it had not been used in a large-scale project in the United States. he performance of the slurry trench at Aliceville Lock and Dam surpassed all expectations, and its use spread throughout the project.56 In discussing slurry trench construction in the canal section, Mobile District engineers recommended in 1977 that “since the slurry trench construction is relatively new to the District as well as most contractors, it is required that the successful bidder will sublet the slurry trench to a irm who specializes in a Figure 7.6. Aerial view of Columbus Lock and Dam. (U.S. Army Corps of Engineers)

213

Chapter 7

slurry trench construction. Most of these irms are equipped to conduct their own quality control and for this reason the contractor quality control was let in for this feature.”57 he success of slurry trenches on the Tenn-Tom led designers throughout the Corps to utilize the system in most of its major structures. And its use spread outside the agency as well, to other water-resources developers and to engineers facing other technical challenges, such as the Environmental Protection Agency’s hazardous-waste cleanups and coninements. As the Mobile Chapter of the American Society of Civil Engineers boasted in 1986, “It would be fair to say that the slurry trench came of age on the Tenn-Tom.”58

Canal Section he canal section began just south of Amory, Mississippi, and extended forty-four miles north through the wide loodplain of the upper Tombigbee River to the Bay Springs Lock and Dam in Mackey’s Creek.59 his run of the Tombigbee was smaller and more twisting than its southern reaches. Early in the planning for the Tenn-Tom, Corps designers determined that it was uneconomical to widen and straighten the river in this section. Instead, they proposed building a lateral “perched canal” for this stretch of the waterway that would run parallel to the upper reaches of the East Fork of the Tombigbee River. It would be perched in the sense that it would sit up from the river on the hillsides of the east loodplain. he navigation channel consisted of an excavated canal contained between parallel levees. he west levee separated the canal and the river; the east levee prevented streams and creeks originating in the uplands from entering the canal. his local drainage fed into an excavated ditch running parallel to the canal and would be channeled under the waterway through a number of reinforced-concrete culverts. Water from storms exceeding the capacity of the culverts would be allowed into the canal by way of concrete spillways. Excess water within the canal would be discharged into the river by means of ixed-crest spillways. Five canal locks—identiied as locks A, B, C, D, and E—would overcome the 140-foot diference in elevation from end to end of the canal section. hree of the locks—A, D, and E—had 30-foot lits, while the lits at locks B and C were 25 feet. he Corps reported on this design in its environmental impact statement of March 1971.60 214

B u i l d i n g t h e T e n n e s s e e -To m b i g b e e Wat e r way

he perched-canal concept represented the traditional design approach. During the early 1970s, however, Corps oicials began to question this plan as they prepared the project’s detailed designs. Corps technicians conducted hydraulic studies in the summer of 1971 to evaluate the problems associated with changing the design of this section. Members of Mobile District’s hydrology section and Design Branch met in September to discuss the indings.61 From a hydrologic standpoint, it became clear to many Corps engineers that maintenance on the perched canal would be extremely high, especially for those areas where the culverts passed under the navigable channel. Small, infrequent loods would tend to silt up the culverts, while big loods threatened to block these culverts with debris, thus potentially looding large areas above the waterway and possibly penetrating the canal itself. In. addition, the relatively narrow conines of the parallel levees subjected the canal to potential surge problems from the Bay Springs lockages and also made the canal almost entirely dependent on the Tennessee River for its water supply, since all local drainage would be passed under the waterway. From an aesthetic viewpoint, the perched canal would present an unsightly scar on the land—a straight ditch stretching for miles above the Tombigbee River. As early as July 1971, Acting Mobile District Engineer Lieutenant Colonel Paul D. Sontag could claim, in defending the Corps’s environmental planning for the waterway, that “we are currently considering design changes for the entire canal section to make the canal a productive ish and wildlife area.”62 Arthur M. (Bud) Cronenberg, a forty-year veteran of the Corps and chief of the Mobile District’s Hydraulic Structures Design Branch, became the leading proponent for the alternative and novel “chain-of-lakes” design concept. He was the irst person to present the idea, and ater he retired in 1971, several of his colleagues elaborated and reined the idea.63 he chain of lakes followed essentially the same alignment as the old design. he levee would remain on the west bank—the riverside of the canal. Its dike would be built from material excavated for the channel. he east-bank levee, however, would be removed, thus creating a series of reservoirs with their eastern boundary determined by the natural lay of the land. Near the levee, maintenance dredging would maintain a navigable channel, while the rest of the impounded water would serve as a series of lakes for aesthetic and recreational purposes. he streams and creeks coming of the 215

Chapter 7

hills would low directly into the waterway. his inlow would be passed into the natural river to the west by means of strategically located spillways. he ive locks—whose locations and lits remained essentially unchanged from the perched-canal design—would be incorporated into earthen dams that tied the levee into higher ground. hus, the project would not require the elaborate system of culverts running under the channel, and the broad surface area of the lakes would absorb the surges produced by the lockages at Bay Springs.64 In a status report on the Tenn-Tom issued by the Mobile District’s Engineering Division in March 1973, Corps designers already strongly favored the chain-of-lakes concept. “Considering the various factors that require evaluation today,” read the report, “this drainage ditch [the one intercepting the local streams] would be very undesirable environmentally and would result in wildlife and recreational beneits being practically nonexistent.” he designers stated that the chain of lakes “will provide numerous environmental and ecological beneits not available in the authorized plan and would eliminate the need for a drainage ditch. he appearance of the waterway will be more natural and the turbidity of the water resulting from traic will be less. he lakes will provide a desirable habitat for ishes and waterfowl and provide various recreational beneits.” Although 8,300 acres would be looded rather than the 4,300 acres contemplated under the original design, the chain of lakes did not necessitate the acquisition of additional land, the perched canal required the same amount of acreage because the uplying land outside the levees was subject to periodic looding.65 In November 1973, senior planners at the South Atlantic Division met with their counterparts from Mobile and Nashville districts to review design questions. he SAD oicials informed their commander that Corps designers had yet to resolve all their diferences with regard to the canal section. One of the problems, they reported, was that the perched canal remained “the authorized scheme discussed in previous reports.” hey explained, however, that oicials testifying before the District Court in the Tenn-Tom litigation emphasized “the fact that the chain-of-lakes approach was environmentally more acceptable and more beneicial than the perched canal approach.”66 Despite Mobile’s enthusiasm for the chain of lakes, SAD hesitated to approve the new design. Colonel Drake Wilson, the Mobile District engineer, 216

B u i l d i n g t h e T e n n e s s e e -To m b i g b e e Wat e r way

submitted an evaluation of the alternative plans for the canal section to the South Atlantic Division engineer in January 1974. “In view of the present tight schedule for planning and construction of the canal section,” Wilson asked for an expeditious review of the Mobile District report so that future planning could be conined to but one design.67 Mobile found the perched canal to have no major advantages over the chain of lakes, although its disadvantages were many: “(a) diicult hydraulic problems in water handling, (b) excessive excavation with non-economical spoil utilization, (c) limited potential for recreation development, and (d) least appealing aesthetically.” Construction costs for the two designs varied little. he major structures, canal excavation, and land requirements were the same. Land-clearing requirements were less for the perched canal, while the chain of lakes ofered a cost savings in that it required only a single levee. he cost of the drainage ditches and culverts in the irst design were ofset by the cost of the spillway structures in the second design. Mobile strongly recommended adoption of the chain-of-lakes design, claiming that it (a) conforms to the authorized project plan; (b) is within the project concept limits which have been ruled on by the courts; (c) will not necessitate revision to the Environmental Impact statement; (d) has been planned, coordinated and accepted by the local sector; (e) is clearly advantageous over the authorized plan; and (f ) can be pursued with the least disruption of the presently adopted planning and construction schedule for the waterway.68

Mobile’s arguments prevailed, and the Corps oicially adopted the chainof-lakes design in June 1974. Corps designers favored the chain-of-lakes concept primarily because it ofered a better engineering solution. Mobile’s Board of Environmental Consultants fully supported the change because it added signiicant opportunities for lat-water recreation, in addition to being more aesthetically pleasing.69 Like the resiting of Plymouth Bluf, the redesign of the canal section became a banner of the Corps’s environmental sensitivity. Corps spokesmen and other Tenn-Tom enthusiasts touted this change, and stressed the environmental considerations over those of engineering. Yet the canal section had been used as 217

Chapter 7

an example of environmental sensitivity even under the old design. he river section involved signiicant alterations to the Tombigbee River and its environment, and the divide section involved a massive earth-moving efort. he canal section, in comparison, appeared far more benign, whether it be built under the perched design or chain-of-lakes design. As Glover Wilkins of the Tennessee-Tombigbee Waterway Development Authority proudly told a Time/Life photographer who was ilming the proposed waterway route in spring 1970, in the canal section “we will leave the rambling river stream and hardwood forest intact while we go up on the hillside to build the canal.”70 When coaching Representative homas G. Abernethy of Mississippi in January 1971 about dealing with environmental criticism of the Tenn-Tom, Wilkins wrote, “Another bit of phraseology we are using is that our canal will be a ‘dry cut’ meaning we will not disturb the wet lands [sic]. I wish you would consider using this also as it has a good connotation.”71

Foundation Conditions and the Design of Lock D he abandonment of the perched canal for the chain of lakes represented a major design change. Although less grand in scale, Lock D also led Corps oicials to consider making a signiicant deviation from the original design concept. Unlike the chain of lakes, however, political considerations outweighed the favored engineering approach at Lock D, thus exemplifying the technical and political complexities of civil-engineering design. he issue at Lock D involved unusually poor foundation conditions and high artesian pressures. Foundation conditions throughout the canal section were less than ideal for the construction of massive navigation locks, but the Corps’s initial surveys and subsurface testing supported the designer’s belief that low-lit locks of the conventional gravity design could be built at all ive sites. he ground at the Lock D site, however, was particularly swampy. When Corps oicials began taking extensive core samples from the area in spring 1975, they found seams of sot clay and high artesian pressures that provided inadequate foundation strengths to support the bearing pressures of the gravitywall design.72 218

B u i l d i n g t h e T e n n e s s e e -To m b i g b e e Wat e r way

Corps oicials began to investigate alternative lock designs. heir detailed engineering and economic evaluations were presented at a series of design conferences held in 1975. By September, two alternative plans for Lock D made it past the technical debate. he overall original plan for the canal section called for lits of 25 feet at Lock C and 30 feet at locks D and E. Design alternative “A” called for lits of 25 feet at Lock C, 25 feet at Lock D, and 35 feet at Lock E, while alternative “B” listed lits of 25 feet at Lock C, the elimination of Lock D, and 60 feet at Lock E.73 At this same meeting, however, Mobile District Counsel Alfred Holmes advised against major changes in the overall design of the waterway. Fred hompson, chief of Mobile’s Project Engineering and Conference Section, reported that “Holmes was of the opinion that elimination of Lock D would be a changed condition which would probably invite litigation. . . .”74 Given the fact that the Corps was once again involved in litigation over the Tenn-Tom, and that this litigation challenged the Corps’s discretionary authority to adopt the chain-of-lakes design without Congressional authorization, the Mobile District’s Oice of Counsel took a special interest in the proposed changes in Lock D. David H. Webb, the assistant district counsel who was taking the lead in the Tenn-Tom suit, explained to the district engineer and chiefs of the Planning and Engineering Division that NEPA required that environmental considerations be taken into account by federal decision makers along with economic and technical considerations. Webb advised, As the District considers the possibility of eliminating Lock D in the Tenn-Tom Waterway or otherwise altering the canal section, foremost thought must be directed to the guiding principle that ultimate engineering (technical) decisions must be made subsequent to economic, environmental, or other studies rather than studies being performed because of an engineering decision. Furthermore, the District must avoid the appearance of studies being undertaken as a consequence of predetermined decisions.75

Corps oicials from the various divisions and branches within Mobile District continued to study the design alternatives. Planning Division underscored the loss of recreational development in the canal section if Lock D were eliminated. Operations and Maintenance estimated an annual savings of $95,000 219

Chapter 7

Figures 7.7 & 7.8. Lock A (top) and Lock B (bottom). Modern locks A through E were erected between Bay Springs Lock and Dam and Aberdeen Lock and Dam. (U.S. Army Corps of Engineers) 220

B u i l d i n g t h e T e n n e s s e e -To m b i g b e e Wat e r way

without the lock. Mobile’s economists estimated $46.6 million in additional navigation beneits over the ity-year life of the project without Lock D because of faster overall towing time through the waterway. he Foundations Branch strongly favored the elimination of Lock D and the resiting of Lock E to a more favorable location where the heightened lock could be built on sandstone. he Real Estate Division also favored the elimination of Lock D as it would reduce the acreage requirements of the canal section by 660 acres in fee and 600 acres in easements, producing a cost savings of $922,000.76 Commenting outside its area of expertise, Real Estate Division suggested that the substantial navigation beneits ofered by the elimination of Lock D should weigh heavily in the Corps’s inal decision since the waterway’s primary purpose was navigation. . . . we feel that the overall beneits to be gained from deleting Lock “D” dictate that we assume a positive position and proceed with our planning on the basis of its elimination. If we are successful in our present rationale that a postauthorization change is not necessary for the Chain of Lakes concept in lieu of a perched waterway, this additional design change would appear to fall well within the purview of such rationalization. Should we opt to proceed for a PAC on the Chain of Lakes design then the Lock “D” elimination would merely be added as a feature of that action.77

In a background report given to the Board of Consultants, Mobile’s Planning Division summarized the pros and cons: Elimination of Lock D posed signiicant losses for recreation in the canal section. Without Lock D, the canal section would be missing its largest planned recreation site—the Beaver Lake Recreation Area. Corps planners estimated that Beaver Lake would draw as large a visitation as the canal section’s other three planned recreation areas combined. he canal section would lose 1,200 acres of lake surface without Lock D, and its elimination would lessen the real estate required for the project by 1,350 acres. However, the absence of Lock D would require the excavation of an additional 10 million cubic yards in building the navigation channel. Navigational beneits stood to improve without Lock D as the travel time would be reduced by the time originally allotted for locking through Lock D—an estimated savings of $46.6 million over the 50-year life 221

Chapter 7

(Top) Figure 7.9. Section of the Tenn-Tom Waterway showing the chain of lakes. he lakes result from a conventional control structure along the lower side of the canal that allows lakes to form above the canal, extending to nearby hills. (U.S. Army Corps of Engineers). (Above) Figure 7.10. Lock C. (U.S. Army Corps of Engineers).

222

B u i l d i n g t h e T e n n e s s e e -To m b i g b e e Wat e r way

of the project. Annual operation and maintenance costs would be reduced, as salaries for personnel to operate Lock D could be eliminated as could the maintenance costs of keeping up the facility. Actual construction costs would be $1,776,000 less for building the project without Lock D. Finally, without Lock D, Lock E could be resited to a location with better foundation conditions—from the original site consisting of Gordo sand and gravel to a site founded on Hartselle sandstone.78 Mobile engineers reported in April 1977 that all three plans were “feasible as to engineering integrity and safety,” that the “costs of the 3 alternative plans are nearly equal,” that the “environmental impacts of each plan difer but none should pose insurmountable problems,” and that the elimination of Lock D would be less “conducive to the development of recreation areas as the other alternatives.”79 Further studies were called for, and it was noted that elimination of Lock D would require approval from the Secretary of the Army and any alteration from the general design memorandum would require a supplement to the environmental impact statement. here was also concern that any major changes might require Congressional authorization.80 Despite the strong engineering and economic arguments in favor of eliminating Lock D, the political and legal considerations led senior Corps oicials to retain the lock and pursue an otherwise less-than-ideal technical solution. Mobile therefore adopted a special rigid U-frame—or “bathtub”—design for Lock D, a design that allowed the entire structure to loat. To produce this rigid structure, Corps engineers designed a thick, heavily reinforced concrete loor. his lowered the bearing pressure from the standard gravity design of four tons per square foot to about two tons per square foot, guarding against diferential settlement. his design cost more than the others, but it was determined by Corps decision makers that it was worth the extra cost.81 his was a novel approach for the agency. Earlier, Corps designers questioned the strength of the foundation material at the Aliceville Lock site. In summer 1971, the Civil Engineering Section of the Mobile District conducted an analysis of a U-frame design as an alternative to the traditional gravity design.82 he basis of this consideration was Rolf Bobe’s “Base Pressure Distribution under Half Frame-Like Locks,” and in the early 1970s, Corps designers knew of no other 110-foot-wide lock designed as a U-frame.83 Ultimately, Aliceville was 223

Chapter 7

designed with the gravity lock, and Lock D became the irst large-scale lock built by the Corps with the U-frame design.

Divide Section he 40-mile-long divide section extended north from Bay Springs Lock and Dam in Mackay’s Creek, through a 27-mile-long cut in the ridge separating the Tennessee and Tombigbee river basins, to the Yellow Creek embayment, an arm of Pickwick Lake on the Tennessee River. he lock at Bay Springs was the largest on the Tenn-Tom with a lit of 84 feet. he elevation of the divide itself was about 570 feet above mean sea level, while the deepest part of the divide cut was 175 feet. Nearly 70 percent—or approximately 145 million cubic yards—of the earth excavated along the entire waterway came from the divide-cut section. “he earth removed from the divide cut is enough to build a two-lane highway from the Earth to the Moon,” declared the General Accounting Oice in 1981.84 As Nashville District geologist Marvin Simmons explained in 1971, the topography of the divide-cut area “ranges from smoothly rounded hills of low relief separated by broad valleys, to hills and ridges of 200 feet of relief with steep slopes, narrow crests and narrow separating valleys.”85 Nashville designers considered such factors as geology, soil conditions, groundwater, rainfall patterns, drainage patterns, and the like. he soils design for the cut area drew special attention. To determine slope stability, engineers calculated the efect of artesian pressures, efects of excavation methods, and the need for subsequent slope protection. Subsurface exploration and laboratory tests aided both the geologists and soils specialists in formulating the project design. Instrumentation was installed to monitor movements within the cut slopes.

Nuclear Excavation Excavating the divide cut threatened to be the most costly aspect of the waterway—certainly it was the most massive earthmoving challenge—and Corps designers correspondingly explored alternative construction techniques. 224

B u i l d i n g t h e T e n n e s s e e -To m b i g b e e Wat e r way

Spurred by Project Plowshare, which had been started in 1957 to explore the feasibility of building harbors and canals using nuclear explosives, the U.S. Army Engineer Nuclear Cratering Group studied excavating the Tenn-Tom by means of nuclear explosions. he proposal to build a sea-level canal through Nicaragua generated political support of the Tenn-Tom consideration, since promoters of the Tenn-Tom had oten likened the project to the Panama Canal. Economic and engineering studies were made. U.S. News and World Report covered these developments in its May 20, 1963, issue. “he irst use of a nuclear explosion for peacetime construction,” it reported, “may be to help build a canal in the U.S. South, instead of a harbor in Alaska or another Panama Canal.” It noted that oicials from Mississippi, Alabama, Tennessee, and Kentucky had launched a campaign “to ‘sell’ nuclear excavation of part of a planned 253-mile waterway to connect the Tennessee River with Alabama’s Tombigbee River.” Nuclear excavation, the proponents argued, could reduce the costs of excavating the divide-cut section of the waterway.86 he Tennessee-Tombigbee Waterway Development Authority retained the engineering consulting irm Brown and Root to study the possibilities of using nuclear explosives in building the waterway.87 In September 1963, the authority reported, he AEC is moving ahead with preparations for underground nuclear tests at the Tatum salt dome near Purvis, Miss., to establish the feasibility of using small nuclear devices on such projects as the Tennessee-Tombigbee Waterway. Signiicantly, underground testing is not banned by the nuclear test ban treaty. Dr. Edward Teller, “father of the H-bomb” and a principal witness against the treaty, has visited Mississippi and believes the Tennessee-Tombigbee divide cut is “ideally suited” for excavation by nuclear energy.88

he Nuclear Test Ban Treaty of 1963 ultimately ended this lirtation with nuclear excavation, despite the authority’s statements to the contrary. he treaty outlawed atmospheric testing, and nuclear charges used in excavation bordered too closely on this, as the likelihood of escape into the atmosphere was high. Nevertheless, in the environmental impact statement iled by the Corps in 1971, nuclear excavation was listed as an “alternate method by which the 225

Chapter 7

waterway connection could be provided”—a method that had been rejected by the agency. “Such construction,” the Corps charged, “would be complicated by possible increases in background radiation levels and atmospheric contamination along with associated environmental problems.”89

Design Design of the divide cut depended heavily on the types of material to be excavated. Very little rock had to be removed, which greatly enhanced the project’s economic feasibility since construction prices would have skyrocketed had the Corps been forced to excavate through rock. Instead, the Corps found a fairly consistent erodable silty sand, primarily made up of the Eutaw formation. his would allow contractors to utilize giant earth-moving equipment—of the type oten employed in open-pit mining—to excavate the material, thus reducing the unit cost of earth removal. his type of material, however, required gentle slopes to avoid erosion, and a series of benches had to be built. his increased the total amount of material to be excavated—at the deepest sections of the project, for example, the cut would be a half-mile wide. he estimated amount of earth to be excavated from the divide cut varied depending on the alignment, slope design, and other factors. December 1973 estimates placed excavation of the 27-mile-long divide cut at 161 million cubic yards.90 Minimizing this amount became a major design challenge since the excavation volume directly inluenced the costs, which were typically calculated on a per-cubic-yard basis. Reducing the excavation requirements would also lessen the problem of disposing of the material—a problem related not only to environmental considerations, but also to real-estate considerations, as more disposal land would be required. Safety and future maintenance costs (say, from slides and slumping) had to be weighed against excavation costs, the environmental consequences of dumping enormous quantities of excavated material, and other considerations. he 1966 general design memorandum used $0.27 per cubic yard as the unit cost for excavating the divide cut, a igure that translated the estimated 144 million cubic yards of common excavation to a cost of $38,928,000. Environmental 226

B u i l d i n g t h e T e n n e s s e e -To m b i g b e e Wat e r way

legislation, passed ater this estimate, meant that care and treatment of this material would have to minimize environmental damage. By May 1973, members of the OCE’s Directorate of Civil Works worried that “the cost of spoil area treatment on the Tenn-Tom Project could generate near prohibitive costs.” heir concern stemmed from the Corps’s recent experience on the Trinity River Navigation project, where “the estimated cost of extra haul to areas not directly along the banks, acquisition of land for spoil areas and landscape treatment to satisfy environmental requirements was in excess of $150,000,000.” Reduction of required excavation would thus help reduce construction costs at several levels. In considering ways to accomplish this, the Directorate of Civil Works requested that SAD study the engineering pros and cons of limiting the depth of the divide cut by excavating a summit-level canal with a set of locks (each forty to sixty feet high) at either end. he OCE was interested in knowing how much excavation would be saved in such a plan, how much the two additional locks would cost, and how much water would have to be pumped up to the summit-level canal from Pickwick Lake.91 Oicials at SAD had already pursued this idea. In September 1972, they asked Nashville District’s Engineering Division to study this alternative design. Nashville’s recommendation reached SAD in June 1973. Despite the reduction in excavation required by the summit-level canal, the additional two locks brought the initial construction costs $32 million over the original design, Nashville reported. In addition, the summit-level canal and two additional locks would require greater operating and maintenance costs and would include the extra cost of pumping water. Moreover, project beneits would be substantially reduced by this plan because of the increased lockage time imposed upon barge traic. hese negative economic factors led Nashville to recommend against the design alternative.92 Seeking other ways to reduce substantially the required volume of excavation, the OCE’s Directorate of Engineering sponsored a design conference on the divide-cut slope criteria in December 1973. Cost and environmental issues were supplemented as motivating factors in this case by the evolving dispute between the Corps and the Illinois Central and Gulf Railroad, whose line had to be relocated because of the project. he general design memorandum for the Tenn-Tom called for the deepest portions of the divide cut (some 200 feet) 227

Chapter 7

to have rather lat slopes, rising one vertical foot for each four horizontal feet. his required massive excavations and a relocation alignment of the Illinois Central’s line, which the railroad would not accept. If that excavation could be reduced, the railroad could then be routed through a valley that had been originally reserved for spoil disposal, and this alternative relocation the railroad would accept.93 Stabilization of the cut slopes along the divide cut became one of the major points of diference between engineers at the Nashville District and engineers at the Mobile District. Nashville’s design approach was to protect the slopes with riprap along the entire course of the divide cut. his was a more conservative approach, and District designers argued that it was simpler and more eicient to place the rock “in the dry” along the entire section, than to come in later to repair sections and be forced to lay the rock from loating platforms. Mobile and the South Atlantic Division, however, adopted a wait-and-see approach— leaving cut slopes unprotected, and rip rap only those sections that later slid. Although the point was debated, Nashville was not overruled, and the divide cut was built as the only section of the waterway to be lined fully with riprap. To prevent the erosion of soils underlying the placed rock, ilter blankets were to cover the cut slopes.94

Spoil Disposal Disposing of the huge amounts of material from the divide cut was perhaps the most environmentally controversial aspect of the project, and the Corps was placed under extreme pressure to produce a viable and environmentally acceptable solution. As early as March 1967, James T. Tozzi, a senior staf analyst with the oice of the Secretary of the Army, addressed the “scenic consideration” of disposing of the excavated material from the divide cut. In criticizing the Corps’s general design memorandum for the project, Tozzi stated: It appears that the Corps is assuming that the excavated material is to be displaced laterally, i.e., adjacent to the sides of the proposed waterway. I cannot determine from the report if the Corps intends to pile the excavation material in the form of a bank or to spread it over the adjacent land; in any case, such

228

B u i l d i n g t h e T e n n e s s e e -To m b i g b e e Wat e r way

an operation will decrease the attractiveness of the area by creating either large mounds of debris or a large expanse of land without vegetation.95

Tozzi said that the Corps had “deposited the excavation materials of the Cross Florida Barge Canal near its banks” and was “now in the process of correcting the situation by spending additional funds to improve the attractiveness of the area.” He recommended a close evaluation of the Corps’s plans with regard to the divide-cut excavation.96 Disposal of the huge amounts of excavated material in the divide-cut section was a monumental task, and one that involved immense environmental considerations. he 1971 Environmental Impact Statement estimated that in the divide cut alone, 140 million cubic yards of excavated material would have to be disposed of.97 he Board of Consultants took a special interest in this aspect of the waterway, arguing that the disposal areas should be shaped, blended, and contoured into the natural topography.98 Great eforts were also made in revegetating the disposal sites, which meant neutralizing the high acid content of the excavated soils. To help compensate for the loss of wildlife habitat in the dividecut area, ponds were constructed within the disposal areas. he environmental impact statement simply stated that the excavated material would be “wasted adjacent to the cuts.” he report added, however, that inal designs for the disposal areas were still under consideration, and the agency anticipated making signiicant changes for environmental reasons. Disposal methods under consideration included “use of some of the material as ill for roads and highways, rehabilitation of blighted areas, such as abandoned gravel pits and eroded gullies, illing and site preparation of industrial and commercial sites” and even the disposal of excavated material in Bay Springs Lake to reduce its depth and to cover mercury-laden sediments.99 In August 1971, senior Corps planners met with members of the Board of Consultants to discuss the environmental problems associated with the divide cut. Prior to the meeting, the board members were lown over the entire project area in a helicopter, then driven through the divide-cut area for an on-ground inspection. Disposal of the excavated material was a central topic of discussion, and Gerald McLindon, the board’s landscape architect, took a special interest in this problem. He recommended that the Corps take “a positive attitude” 229

Chapter 7

toward this task and attempt “to improve the environment” rather than simply disposing of the material as economically as possible. McLindon called for a “brainstorming” session on this issue but presented some potential uses of the material. hese included building a site for a city of 30,000 people and building sites for golf courses, industrial parks, a nuclear generating plant, and an airport.100 In July 1974, Nashville District Engineer William Brandes announced the Corps’s plans for disposing of the excavated material from the divide cut. According to Brandes, the estimated 145 million cubic yards of earth would be dumped in 51 valleys along the waterway. Following the deposit of this material, the Corps promised to revegetate the poor soil—poor because it came from deep underground and thus was low in organic content—and eventually turn it over to the Mississippi State Park Commission for use as park and game-management lands. To assist the Corps in developing a revegetation plan, Nashville awarded a contract to Mississippi State University.101 Treatment of the excavated materials became an important consideration in the divide cut. Early test samples failed to indicate the high acidic content of the underlying soils, and Corps oicials found that although the newly seeded disposal sites sported healthy-looking grasses for the irst few months, oxidation of the exposed earth soon lowered the pH levels, thus killing all their plantings.102

Modeling Within the divide cut, Corps designers faced the challenge of passing the waters of the area’s tributary streams into the waterway without eroding the side slopes, scouring the bottom of the channel, or endangering navigation through forceful lateral surges. Runof from the surrounding area would be channeled into the waterway by means of ive major drainage structures and over one hundred small drainage structures—and each of them had to be designed to handle the drainage from heavy rainfall. he engineering challenges were magniied by the fact that several tributary streams originated at substantially higher elevations than the navigation channel. he smaller streams feeding the waterway naturally posed the least concern, and Corps personnel utilized the traditional spillway design, allowing the water to low down the slopes in a shoot with its 230

B u i l d i n g t h e T e n n e s s e e -To m b i g b e e Wat e r way

energy being absorbed through the “hydraulic jump” created when it entered the main body of water. he larger streams, however, required a diferent approach. Because the hydraulic characteristics of the larger structures could not be reliably determined by mathematical analyses, in February 1974 the Nashville District requested the Hydraulics Laboratory at WES to model-test the structures. Between March 1974 and May 1975, the laboratory constructed and tested models of the drainage structures and surrounding basin. he laboratory borrowed from control structures developed by the Bureau of Reclamation for use in the American West. As a result of these tests, WES recommended a baled spillway design to dissipate the energy of the incoming water. Large, staggered concrete “teeth” were implanted in the spillway to absorb the energy of the falling water, which successfully eliminated erosion, scouring, and surges. Moreover, it aerated the incoming water, thereby improving the water quality of the tributary streams.103

Bay Spring Lock and Dam Bay Spring Lock and Dam is located at the southern end of the divide cut, twenty-eight miles northeast of Tupelo, Mississippi. It is the northernmost navigational structure on the Tenn-Tom and is also the most impressive. It consists of a rockill dam (with an impervious earth core), 120 feet high and 2,750 feet long. he 7-mile-long impoundment covered 6,700 acres and remained at the same elevation as Pickwick Lake, which it joined through the divide cut and the Yellow Creek embayment. Its 84-foot lit makes it the third-highest lit lock east of the Mississippi River. he rock foundation at the Bay Springs site allowed Corps engineers to design a high-lit lock, which a soil foundation would not support. he size of Bay Spring prompted Corps designers to study the surges associated with the 45 million gallons used in illing and emptying the lock. he Corps’s Hydraulic Design oice instigated a surge-model study for the lock to explore design alternatives. he Waterways Experiment Station conducted a model study to determine the efects of emptying the lock on tows and to explore methods of reducing the surge either by deepening the channel or by providing a surge basin. WES also explored the efects of various speeds 231

Chapter 7

of illing and emptying the lock. In addition to cost and safety, lockage time received prime consideration, especially since Bay Springs’s 84-foot lit meant that it would be the slowest lockage on the entire waterway.104 Nashville designers remained uncertain about the downstream turbulence caused by the rapid emptying of 6.19 million cubic feet of water from the lock chamber. In 1974, they requested WES to construct a hydraulic model to test the operational efects of their lock-discharge system design. From October 1974 to 1976, WES’s Hydraulics Laboratory carried out the model test and found that Nashville’s original design produced dangerous surges in the downstream approach and recommended modifying the design of the discharge system and lower approach geometry to difuse the surge and slow the upwelling of the water surface.105 Between November 1974 and May 1976, WES’s Hydraulics Laboratory built and tested a 1:25-scale model of Bay Springs Lock to study the longitudinal Figure 7.11. Waterway construction necessitated the relocation of numerous transportation systems such as roads and railways. his aerial view shows the new alignment of the Illinois Central–Gulf Railroad together with a new bridge spanning the waterway. (U.S. Army Corps of Engineers) 232

B u i l d i n g t h e T e n n e s s e e -To m b i g b e e Wat e r way

loor culvert illing-and-emptying system proposed by the Nashville District. his system featured “tuning fork”-shaped culverts within the lock chamber. Safety, reliability, and speed of operation were primary considerations. Safety issues posed a greater technical challenge in high-lit locks than in low-lit locks, as there existed greater potential for dangerous surges within the lock chamber during illing. WES found the Nashville proposal satisfactory.106 he illing and emptying system of the Tenn-Tom’s other locks utilized longitudinal loor laterals within the lock chamber. his meant that water would enter and leave the lock chamber through loor culverts arranged symmetrically along the center of the chamber. It took 8.5 minutes, for example, to empty the 36-foot-lit Gainesville Lock. Although eicient for low-lit locks, this system held less appeal for high-lit locks like Bay Springs, where the 84-foot lit would take about 30 minutes to ill using this system.107 he speed of illing and emptying Bay Springs Lock became such an important issue because of its economic ramiications. Navigation savings represented the lion’s share of beneits in the Tenn-Tom economic justiication, and those savings were measured in part by the annual tonnage passing through the system. Several factors restricted the maximum potential tonnage. hese factors included tow sizes limited to eight barges, the numerous curves in the watercourse, and the time spent passing through the ten locks. he more traic that could be passed through the system—at least in theory—the more beneits the Corps could claim for the project. he quest for the highest possible beneit/cost ratio therefore led Corps designers to streamline whatever traic bottlenecks they could. he lockage time at Bay Springs—which was the system’s tallest lock by a factor greater than two—represented an obvious target. Ater receiving A.T. Kearney’s economic re-analysis of the waterway in 1976, for example, Corps economists calculated that a four-minute increase in the lockage time at Bay Springs could create over $4.5 million in annual navigation beneits loss (assuming maximum traic on the waterway).108 Bay Springs, therefore, became the limiting factor in the Tenn-Tom’s traic capacity, and Corps designers found great economic incentive to design the fastest hydraulic illing and emptying system possible within safety limits. he hydraulic illing and emptying system of the huge lock represented a major design consideration. he side port illing and emptying system, which 233

Chapter 7

was standard design for most of the Corps’s low-lit locks, had a 32-minute illing time for Bay Springs, which Corps planners deemed “unacceptable.” In studying alternative approaches, Corps oicials found the bottom split lateral system incorporated in the Wilson Lock (100-foot lit) boasted a 14-minute illing time, while the interlaced lateral system used in the Little Goose Lock (also having a 100-foot lit) took 14 minutes to empty. Corps designers observed in September 1975 that “the bottom lateral system is not as safe as the bottom longitudinal system and is slower.” he optimal design would combine speed of illing and emptying with minimal turbulence. he designers argued that “since the Tenn-Tom project is a irst class system connecting two irst class waterways and since Bay springs is already the slowest on the waterway, the present good design [i.e., the bottom longitudinal hydraulic system] should be retained.”109 Nashville District engineers remained strongly in favor of the longitudinal illing and emptying system, even though the district had yet to build one. he design concept was not new to the Corps, however, and Nashville designers patterned the Bay Springs Lock hydraulic system ater the illing and emptying system on the Bankhead Lock on Alabama’s Black Warrior River. In June 1976, Nashville’s engineers described the system’s advantages as “its speed of illing, minimal disturbance in the lock chambers, favorable hawser stresses and lack of sensitivity to valve mis-operation.” hey noted that the Waterways Experiment Station conducted model tests and found the system the most favorable yet developed “from a technical standpoint.” Prototypes for the other principal design alternative, the bottom lateral system, indicated an additional four minutes in illing and emptying times over the longitudinal system, which for Bay Springs boasted a illing time of 8.5 minutes and an emptying time of 12 minutes. (he emptying time was kept slower to reduce downstream surges.)110 Water quality posed another problem for designers of Bay Springs Lock and Dam. he impounded water behind the dam would be released in surges into the canal with each lockage. he Ohio River Division’s Water Quality Section determined, through its physical and mathematical modeling of the system, that lock discharges during the spring would be cooler than the natural regime of Mackeys Creek during that same period, which would have a long-term detrimental efect on the creek’s aquatic organisms. During the late summer and 234

B u i l d i n g t h e T e n n e s s e e -To m b i g b e e Wat e r way

early fall, releases would be warmer than normal and would likewise be detrimental to native organisms in Mackeys Creek. Corps designers could only compensate in part for this phenomenon by drawing the lock intake water from diferent levels of the lake, because if they drew water too deeply, the dissolved oxygen content would be unacceptably low (as the content of dissolved oxygen diminishes with depth).111 Floods on the Tennessee River concerned Corps designers, for if the Bay Springs Lock or Dam were breached, then the Tennessee River—which was connected to Bay Springs Lake through Pickwick Lake—might be diverted into the Tombigbee River. he lock would have to be closed in the event of a major lood. An emergency closure system had to be installed in the lock, should for any reason the lock fail. he solution was the installation of a permanent stileg derrick (a large crane) capable of placing prefabricated steel stop logs upstream of the miter gates to form a bulkhead. (his closure system could also be used to dewater the lock chamber for periodic inspections and maintenance.)112 As a backup measure, the Corps stockpiled rock material on both sides of the divide cut near Paden so that front-end loaders and dump trucks could block the channel with this material to prevent the siphoning of the Tennessee River into the Tombigbee.113 he structural design of the lock walls also presented a major issue to Corps engineers. Cost was a key factor, as Bay Springs represented the largest structure on the Tenn-Tom. he general design memorandum called for a gravity design, which entailed high costs because of the large amount of concrete required. In 1975, a value engineering team within the Corps studied the possibility of using either a U-frame design or a hang-on wall design, which would be anchored into the shale face. he anchored design, while more daring, sported the advantage of having a relatively thin lock wall. It also required less excavating within the rocky site. he lock features themselves remained the same for all three designs.114 Discussion of the inal design for the lock walls continued into April 1976. Nashville District estimated that the gravity wall design would cost $35.9 million to build, while the U-frame design would cost $35.2 million and the anchored wall design $34.7 million. Nashville’s engineers argued that “the gravity type wall was simpler and more predictable to construct and therefore had 235

Chapter 7

the least likelihood of generating additional construction costs,” and they urged their colleagues to adopt that design.115 However, Richard C. Armstrong, chief of Ohio River Division’s Engineering Division, disagreed. He believed the $1.2 million saved by the anchored wall design to be substantial, and felt that “this cost diference should have been suicient justiication to select the anchored wall as the preferred plan.” He told the South Atlantic Division engineer that his oice believed that “the anchor plan and associated excavation plan was unduly conservative and that reinements in design, utilizing the services of a consultant experienced in the ield of anchor wall construction, would result in a substantial savings to the government.”116 By June, Nashville District engineers had reined their cost estimates of the Bay Springs Lock wall: $33.7 million for either the gravity design or the combined anchored and V-frame wall design; and $35.2 million for the V-frame design. he engineers determined that the rock at the construction site—which consisted of thick-bedded sandstone over moderately sot shale—was of suicient strength to allow construction of all three lock-wall designs. Ohio River Division engineers continued to favor the anchored wall design, while Nashville backed the gravity design. Although the dispute remained unresolved ater the June 1976 design conference, Corps oicials agreed that the gravity system “should be recommended for construction,” but that further cost estimates should also be developed for the anchored wall design.117 Nashville prepared feature design memoranda and cost estimates for the two principal alternatives for the lock chamber walls: the gravity and hang-on wall designs. Going into this detailed study, Nashville’s designers continued to back the gravity-type design—it was the standard design, and the Corps had extensive experience in building such structures. he hang-on wall was a unique design concept for lock walls, and the conservative nature of the district engineers led them to be cautious of it. Although the cost of building Bay Springs Lock with the gravity design was higher because of the large amount of concrete involved, preparation of the site was much easier. To build the lock with a hang-on wall, the rock face had to be carefully excavated to provide the required surface. hen deep holes eight inches in diameter had to be drilled into the rock face so that bundles of prestressed cables—known to the designers as multistrand tendons—could be inserted and anchored to the rock. Twenty-four of these holes 236

B u i l d i n g t h e T e n n e s s e e -To m b i g b e e Wat e r way

extended 120 feet into the bedrock. hese tendons would then be placed under tension and attached to the lock wall. Stringent requirements for the alignment of the holes were maintained throughout construction. his required far more detailed work and the participation of more engineering/construction specialists than did the more straightforward gravity design, which led members of the Nashville Construction Division also to prefer the gravity design during the initial consideration. Yet, engineers at the South Atlantic Division kept encouraging additional studies of the more daring hang-on wall design. Nashville Design Branch Chief Herman Gray recalled that ater this staf prepared the feature design memorandum, SAD asked for their recommendation. “We soul searched quite a bit,” he said, “and recommended the hang-on wall over the gravity wall.” Despite the greater complexities of construction for the hang-on wall, both designs provided ample safety, and the Nashville designers ultimately found the cost savings of the hang-on design compelling.118

Value Engineering Value engineering attempts to reconsider a project’s design procedure to determine if the project could be built diferently—such as using diferent materials or incorporating a modiied design—and provide the same result (including safety and reliability) for less money. Within the Corps, value engineering began in the 1960s but gained currency during the early 1970s—the same time that much of the Tenn-Tom design work was done. Career engineers within the agency initially expressed skepticism toward value engineering, believing that all good engineers automatically practiced value engineering when designing a project. Nevertheless, senior oicials within the agency endorsed the concept, and separate oices of value engineering were established throughout the Corps. hese oices attempted to incorporate various disciplines in their review of projects, and because they were not wedded to particular projects, their review was seen to be impartial.119 Perhaps no other single project ofered the Corps’s value-engineering staf such a rich opportunity to demonstrate the merit of their approach. In many ways, the Tenn-Tom was the ideal subject for a value-engineering evaluation. It was the agency’s largest and most expensive project: it was a project conceived 237

Chapter 7

at an earlier time and redesigned to comply with new environmental and safety standards; its construction costs were skyrocketing; and it was subject to intense criticism, much of it based on a marginal beneit/cost analysis. he South Atlantic Division hosted a Value Engineering Conference in September 1972. “It was determined at the conference,” reported Mobile District’s Assistant Chief of Engineering Division Powell Williams, Jr., “that the valueengineering possibilities presented should be reassessed to determine if savings can be made and if additional studies are warranted.” Mobile’s Engineering Division conducted those reassessments and found that “the dewatering plan for construction of the lock and spillway at the Aliceville Project indicates that the groundwater inlow during construction can be controlled by the use of a slurry cutof constructed underneath the coferdams.” his system would substantially reduce the need for sumping and pumping, and would result in a savings of $1,874,000 in dewatering the construction site.120 In December 1972, Major General D. A. Raymond, the South Atlantic Division engineer, sent a strongly worded memorandum to the SAD Engineering Division chief on the subject of value engineering. Raymond explained that he attended the Tenn-Tom groundbreaking ceremony and he and the Chief of Engineers spoke with Colonel Brandes, the Nashville District Engineer. Raymond quoted Brandes as saying that the Nashville District “hadn’t come up with much on the divide cut” and that “they hadn’t looked at the economics, only the engineering, and that it amounted to only $12 million or so, which was too small a saving, proportionately, to bother with.” Raymond said both he and the chief “were bothered by his comment.” He stressed that all value-engineering decisions were to be made at SAD or OCE, not at the district level. “No possibility is to be summarily dismissed” by the districts, he stated. He mentioned that the Chief of Engineers expected to see “real cost reductions” on the TennTom, and was willing to accept “some risk of slides” on the divide-cut slopes “if need be.” Raymond expressed his own belief “that a reduction in channel width, certainly in the divide cut, and probably elsewhere, will be required for the most signiicant cost reduction savings.”121 he Corps initiated three separate value-engineering studies for the Bay Springs Lock and Dam. he irst examined alternative designs for the lock’s illing and emptying system. he rigid criteria for maximum lock-chamber 238

B u i l d i n g t h e T e n n e s s e e -To m b i g b e e Wat e r way

turbulence and the need to keep the lock illing and emptying times to a minimum led the value-engineering team to endorse the system proposed in the general design memorandum without change. he second study examined the heights of the lock walls and the crest of the dam. Here, the value-engineering team recommended the dam elevation remain unchanged, but that the lock walls be lowered by 6 feet, thus saving the government an estimated $2,247,400. he third study examined alternative designs for the lock walls. As a result of the economic comparisons made in this study, the original gravity-type design was abandoned for a combination anchored walls and gravity-type gate design.122 By January 1976, the Mobile District reported the following actions based upon value engineering: “change from concrete parapet walls to metal hand railing for locks, steeper channel side slopes for Rattlesnake Bend, deletion of riprap on public use mounds, change in tainter value seals, and reduction in length of spillway abutment walls.”123 Joe J. Waits, the value engineering oicer at the Mobile District, proposed in March 1977 that the Corps conduct a value-engineering workshop for the Tennessee-Tombigbee Waterway. In presenting this suggestion to the Mobile District engineer, Waits said, In my view of the continuing scrutiny of the Tennessee-Tombigbee Waterway project, I would like to propose a unique VE project that could reduce the cost considerably. As the cost/beneit ratio may determine the ultimate destiny of the project, we should use one of the Corps’ most valuable resources to improve our position: Value Engineering.124

He recommended studying at least six major projects on the waterway, assigning a special six-person team to each project. he teams were to be composed of “three from Engineering Division, one from Construction Division, and two from other Corps of Engineers Districts throughout the country.”125 A value-engineering workshop on the Tenn-Tom was held at the Mobile District between June 20 and 24, 1977. Issues considered at this meeting included Lock A, the spillway at Lock A, dredging and clearing practices, relocations, public use facilities, and grade stabilization. It was estimated that changes in these six areas could save the government $9,662,126. Members of the District’s 239

Chapter 7

Engineering Division were asked to confer with the value-engineering evaluators to work out the details.126

Conclusion Designing public works projects is a dynamic process requiring many skills. he Tenn-Tom exempliies the complexities of this civil-engineering design process—of the need to balance carefully a wide range of technical factors with a variety of political considerations, many of which change over time. he general design called for a waterway divided into three sections. While this decision was dictated in large part by the geographical characteristics encountered, the feature designs of particular elements of the project relected the numerous options and limitations imposed by political forces. he size and controversial nature of the Tenn-Tom may have made this balancing act more diicult for the designers, but such balancing acts nevertheless remain at the heart of most major civil works projects. Chapter 7 Notes 1 2

3

4

5

Interview, author with Jack H. Bryan, 17 Nov 1986. Maj. Gen. George H. Walker (South Atlantic Div. Engineer) to Ohio River Division Engineer, 12 May 1967 (Tenn-Tom microilm, set I, reel 66, frames 578–580); and James H. Kitchens III, “An Outlet to the Gulf: he TennesseeTombigbee Waterway, 1571–1971,” unpublished manuscript, Oice of History, U.S. Army Corps of Engineers, 1985, 458–459. Brig. Gen. H. G. Woodbury, Jr. (Director of Civil Works) to South Atlantic and Ohio River Division Engineers, 28 June 1967 (Tenn-Tom microilm, set I, reel 66, frames 581–582). Interview, author with Benton Wayne Odom, Jr., 20 May 1986; and Emory Kemp, “he Engineering Aspects of the History of the Tennessee-Tombigbee Waterway,” part I, 2. Unpublished. Interview, author with Robert J. Stephenson (director, South Atlantic Div. Laboratory), 7 Oct 1986; and Col. Patrick J. Kelly (Mobile District engineer) to Robert J. Stephenson, 31 July 1984 (Tenn-Tom iles, South Atlantic Div. Laboratory,

240

B u i l d i n g t h e T e n n e s s e e -To m b i g b e e Wat e r way

6 7

8

9 10

11

12

13

14

Marietta, Georgia). A description of the laboratory and its mission is given in U.S. Army Corps of Engineers, South Atlantic Division, Summary Of Capabilities: South Atlantic Division Laboratory (Atlanta: U.S. Army Corps of Engineers, South Atlantic Div., 1973). Interview, author with Robert J. Stephenson, 7 Oct 1986. See, for example, Bobby P. Fletcher, “Typical Tennessee-Tombigbee Canal section spillways, Spillways A and B: Hydraulic Model Investigation, Technical Report H–78–21” (Vicksburg: U.S. Army Engineer Waterways Experiment Station, Hydraulics Laboratory, Nov 1978). “Summary of Discussions and Decisions, Conference on Aberdeen, Aliceville, and Columbus Projects, Columbus, Mississippi,” 10–11 December 1973 (Columbus L&D Conference folder, Conferences/Review Comments subile, ile 1518–01, CWPF, WPS, OCE). Interview, author with George H. Atkins, 18 Nov 1986. See “Summary of Discussion and Decisions, Conference on Lock 0, DM No. 19, SAD and OCE Comments, Mobile, Alabama,” 9–11 May 1979 (Canal Section Conference folder, Conferences/Review Comments subile, ile 1518–01, CWPF, WPS, OCE); and interview, author with Kenneth D. Underwood, 21 May 1986. “Nashville District Engineering Div., memorandum for the record,” 22 June 1976 (Bay Springs L&D Conference folder, Conferences/Review Comments subile, ile 1518–01, CWPF, WPS, OCE). See, for example, U.S. Army Corps of Engineers, Mobile District, Gainesville Lock and Dam, Tennessee-Tombigbee Waterway Alabama and Mississippi: Design Memorandum No. 12, Instrumentation (Mobile: U.S. Army Corps of Engineers, Mobile District, June 1972). Philip E. LaMoreaux, Daniel J. Nelson, and Gerald J. McLindon, “TennesseeTombigbee Waterway, Board of Consultants for Environmental Concern: Review and Recommendations following the meeting of April 9–10th, 1973,” n.d. [late April 1973] (Board of Environmental Consultants/April 1973 Meeting folder, ile 1501–03, SAMPD-EI). U.S. Army Corps of Engineers, Mobile District, First Supplemental Environmental Report: Continuing Environmental Studies, Tennessee-Tombigbee Waterway,

241

Chapter 7

15

16

17 18 19 20

21 22 23

24

Alabama and Mississippi, Volume I, Overall Study (Mobile: U.S. Army Corps of Engineers, Mobile District, August 1975), 29. Gerald J. McLindon, Stanley I. Auerbach, and Philip E. LaMoreaux, “Report of the 25th Meeting of the Board of Environmental Consultants, TennesseeTombigbee Waterway Annual Field Inspection, 24–26 October 1979” (Board of Environmental Consultants, Oct 1979 Meeting folder, ile 1501–03, SAMPD–EI). Col. Harry A. Griith, memo to South Atlantic Div. Engineer, 19 March 1973 (Tenn-Tom microilm, set I, reel 66, frames 81–83). he economics of this debate are also discussed in F.P. Gaines (chief, Nashville District Engineering Div.), memorandum to Nashville District Engineer, 19 Dec 1972 (TTW Spoil Disposal Areas/1971–1975 folder, ile 1501–07, SAMPD-EI). Col. Drake Wilson to E. Leo Koester, 14 April 1975 (TTWW litigation ile, Ofice of Counsel, Mobile District). Ibid. See Kemp, “Engineering Aspects,” part I, 6–8. Initially, the Corps reported the length of the river section as 168 miles. By 1977, the Corps had shortened the watercourse in the river section by 20 miles, primarily through channel realignments and the inclusion of addition cutofs. See U.S. Army Corps of Engineers, Mobile District, Second Supplemental Environmental Report: Continuing Environmental Studies, Tennessee-Tombigbee Waterway, Alabama-Mississippi, vol I, Overall Study (Mobile: U.S. Army Corps of Engineers, Mobile District, October 1977), 1. Interviews, author with Kenneth D. Underwood, 21 May 1986, and George H. Atkins, 18 Nov 1986. For a detailed discussion of dredging, see Margaret S. Petersen, River Engineering (Englewood Clifs, NJ: Prentice-Hall, 1986), 225–263. See U.S. Army Corps of Engineers, Mobile District, Environmental Statement: Tennessee-Tombigbee Waterway, Alabama and Mississippi Navigation (Mobile: U.S. Army Corps of Engineers, Mobile District, March 1971). he design concept is discussed in Frank C. Deming (chief, Mobile Engineering Division) to South Atlantic Division Engineer, 1 June 1976 (Tenn-Tom microilm, set I, reel 1, frames 1772– 1775); interview, author with Kenneth D.

242

B u i l d i n g t h e T e n n e s s e e -To m b i g b e e Wat e r way

25

26 27 28

29 30 31 32

33

34 35

Underwood, 21 May 1986; and Gerald J. McLindon, “Creative Spoil: Design Concepts, Construction Techniques, and Disposal of Excavated Materials,” Environmental Geology and Water Sciences 7, nos. 1/2 (1985): 91–108. U.S. Army Corps of Engineers, Mobile District, First Supplemental Environmental Report: Continuing Environmental Studies, Tennessee-Tombigbee Waterway (Mobile: U.S. Army Corps of Engineers, Mobile District, August 1975). Tommie Pierce, memorandum for the iles, 29 Jan 1976 (TTW Project/Jan. 1975– Feb. 1976 folder, ile 1503–03, SAMRE). Jack C. Mallory, memorandum for the ile, 12 Oct 1977 (Tenn-Tom microilm, set I, reel 1, frame 56). Interview, author with K.D. Underwood, 21 May 1986; and Willis E. Ruland (chief, Mobile District Environment and Resources Branch) to Don Sherman (consulting engineer to the Waterways Experiment Station), 4 Oct 1976 (TennTom microilm, set I, reel 1, frame 1420). Underwood interview, 21 May 1986. Ibid. Jack C. Mallory, memorandum for the iles, 30 Aug 1976 (TTW cutofs folder, ile 1501–07, SAMPD-EI). U.S. Fish and Wildlife Service, Division of Ecological Services, Bendway Management Study, Tennessee-Tombigbee Waterway Alabama Mississippi: A Fish and Wildlife Coordination Act Report (Daphne, AL: U.S. Fish and Wildlife Service, Division of Ecological Services, March 1984). James R. Couey, “Disposition form to chiefs of Mobile District operations, real estate, and planning divisions,” 1 Sept 1981 (Bendway Task Force/1981–1982 folder, ile 1518–01, SAMPD-EI). homas A. Lightcap, memorandum for the record, 22 Sept 1981 (Bendway Task Force/1981–1982 folder, ile 1518–01, SAMPD-EI). Lawrence R. Green (chief, Mobile Planning Division) to Larry Goldman (ield supervisor, F&WS), 25 Jan 1983 (Bendway Task Force Study/1983 folder, ile 1518– 01, SAMPD-EI). See also John A. Baker, Carolyn L. Bond, and C.H. Pennington, Tennessee-Tombigbee Waterway Bendway Study Biological Impact Assessment: A Letter Report to the U.S. Engineer District, Mobile (Vicksburg: U.S. Army Engineer Waterways Experiment Station, Aquatic Habitat Group, Jan 1983).

243

Chapter 7

36

37

38 39 40

41

42

43

44

45

U.S. Army Corps of Engineers, Mobile District, Design Memorandum No. 3, General Design: Rattlesnake Bend Cutof, Tennessee-Tombigbee Waterway, Alabama and Mississippi (Mobile: U.S. Army Corps of Engineers, Mobile District, March 1972). U.S. Army Corps of Engineers, Mobile District, Gainesville Lock and Dam, Tombigbee River, Alabama and Mississippi: Design (Mobile: U.S. Army Corps of Engineers, Mobile District, January 1969). John R. homan to B. J. Christiansen (chief, Mobile Planning and Reports Branch), 25 Nov 1968, reprinted in ibid., IV–2, 3. Mobile District, Gainesville Lock and Dam . . . General Design, 10. U.S. Army Corps of Engineers, Mobile District, Gainesville Lock and Dam, Tennessee-Tombigbee Waterway, Alabama and Mississippi: Design Memorandum No. 11 Spillway (Mobile: U.S. Army Corps of Engineers, Mobile District, Dec 1971), 1–13; and interview, author with Jack Mallory, 16 May 1986. U.S. Army Corps of Engineers, Mobile District, Tennessee-Tombigbee Waterway, Alabama and Mississippi, Aliceville Lock and Dam Text, Design Data, and Charts for Design Conference (Mobile: U.S. Army Corps of Engineers, Mobile District, March 1972), section I, 1–2. Louis J. Shows and John J. Franco, Navigation Conditions at Aliceville Lock and Dam, Tombigbee River, Mississippi and Alabama: Hydraulic Model Investigation, Technical Report H–78–2 (Vicksburg: U.S. Army Engineer Waterways Experiment Station, Hydraulics Laboratory, April 1978). U.S. Army Corps of Engineers, Mobile District, Environmental Statement: Tennessee-Tombigbee Waterway, Alabama and Mississippi Navigation (Mobile: U.S. Army Corps of Engineers, Mobile District, March 1971), 19. See also interview, author with Jack Mallory, 16 May 1986. Ralph R. W. Beene (Soil Mechanics Branch, OCE), “Trip Report–Inspection of Sites for Columbus (Miss.) and Aliceville (Ala.) Locks and Dams,” 2 Sept 1971 (Tenn-Tom Waterway, General/1970–1971 folder, ile 1518–01, CW–601, SAD). See James W. Erwin (SAD geologist), “Trip Report–Tennessee Tombigbee Project,” 2 Dec 1971 (Tenn-Tom Waterway, General/1970–1971 folder, ile 1518– 01, CW– 601, SAD).

244

B u i l d i n g t h e T e n n e s s e e -To m b i g b e e Wat e r way

46

47

48

49

50

51

52

53

54

55

Col. Harry A. Griith (Mobile District Engineer) to Clark Hubbs (Professor of Zoology, University of Texas at Austin), 28 June 1971 (ALA/MISS–COE: TTWW folder, box 4, ile 429–81–37, CEQ records, WNRC, Suitland, Maryland). D. J. Nelson, “Analysis of Complaint,” n.d. [submitted to the Mobile District on 4 Aug 1971], 2 (Board of Consultants/Nelson Correspondence folder, ile 1501– 07, SAMPD-EI). James D. Wall (chief, SAD Structural Section) to Chief of Engineering Division (SAD), 19 March 1971 (Tenn-Tom Waterway, General/1970–1971 folder, ile 1518– 01, CW–601, SAD). Mobile District, “Use of Earth-Filled Slurry Trenches for Control of Groundwater on the Tennessee-Tombigbee Waterway Project,” unpublished abstract, n.d. [autumn 1977] (Board of Environmental Consultants/January 1978 folder, ile 1501– 03, SAMPD-EI). See also interview, author with Jack H. Bryan, 17 Nov 1986. Petros P. Xanthakos, Slurry Walls (New York: McGraw-Hill Book Company, 1979), 1–17. For a detailed technical description of slurry trench walls, see 194– 221. “Supplement to Design Memorandum No. 7, Gainesville Lock and Dam, Ala., Underseepage Studies,” attached to letter, Col. Robert E. Snetzer to Chief of Engineers, 26 March 1970 (Gainesville L&D folder, Correspondence subile, ile 1518–01, CWPF, WPS, OCE). George H. Mittendorf, memorandum to Mobile District Engineer, 28 June 1971 (Gainesville L&D Conference folder, Conferences/Review Comments subile, ile 1518–01, CWPF, WPS, OCE). “Summary of Discussions and Decisions, Conference on Alternative Studies for Gainesville Dam, South Atlantic Division,” 22 July 1971 (Gainesville L&D Conference folder Conferences/Review Comments subile, ile 1518–01, CWPF, WPS, OCE). U.S. Army Corps of Engineers, Mobile District, Studies of Project Costs and Beneits: Tennessee-Tombigbee Waterway (Mobile: U.S. Army Corps of Engineers, Mobile District, Oct 1975), 14–15. U.S. Army Corps of Engineers, Mobile District, Brieing Data for Army Audit Agency: Tennessee-Tombigbee Waterway (Mobile: U.S. Army Corps of Engineers, Mobile District, January 1976), V–45.

245

Chapter 7

56

57

58 59

60

61 62

63

64

65

Mobile District, “Use of Earth-Filled Slurry Trenches”; Bryan interview; and American Society of Civil Engineers, Mobile Chapter, “Tennessee-Tombigbee Waterway: Nomination for Outstanding Civil Engineering Award (OCEA)– 1986,” unpublished nomination report, 1986, 15. Engineering Division, Mobile District, “Information Document and Instructions on Design Concepts, Lock A and Spillway,” 10 August 1977 (Tenn-Tom microilm, set I, reel 1, frames 301–302). ASCE, Mobile Chapter, “Tennessee-Tombigbee Waterway,” 15. Prior to 1976, the Corps listed the canal section as 45 miles in length. Channel realignments made during the design phase reduced the overall length of this section by one mile. See Mobile District, Second Supplemental Environmental Report, vol. I, 1. Mobile District, Environmental Statement, 20–21. See also, U.S. Army Corps of Engineers, Mobile District, Environmental Study, Tennessee-Tombigbee Waterway, Ala., Miss., Tenn. (Mobile: U.S. Army Corps of Engineers, Mobile District, Aug 1970). Mobile District Engineering Division, “Status Report, Tennessee-Tombigbee Waterway,” 13 Aug 1971 (Williams papers). Lt. Col. Paul D. Sontag to R. Dale Caldwell, 13 July 1971 (Inquiries and Replies, TTW 1971 folder, ile 1501–07, SAMPD-EI). Additional information is provided in interview, author with George H. Atkins, 18 Nov 1986; and Kemp, “Engineering Aspects.” Interviews, author with Jack Mallory, 16 May 1986; Benton Wayne Odom, Jr., 20 May 1986; and Richard T. Kimberl, 21 Nov 1986. For additional background on Cronenberg, see William D. Jones (Head, Microilm Dept., Samford University Library) to Congressman Jack Edwards, 5 Oct 1971 (ile 1–C, carton 34, Jack Edwards papers, University of South Alabama Archives). Background of this design evolution and details of the inal chain-of-lakes design are provided in U.S. Army Corps of Engineers, Mobile District, TennesseeTombigbee Waterway, Alabama and Mississippi, Supplement to the Project Design Memorandum, GDM for Canal Section: Design Memorandum No. 5 General Design (Mobile: U.S. Army Corps of Engineers, Mobile District, Sept 1976). Mobile District, Engineering Division, “Status Report: Tennessee-Tombigbee Waterway,” 16 March 1973 (Tenn-Tom microilm, set I, reel 14, frame 229).

246

B u i l d i n g t h e T e n n e s s e e -To m b i g b e e Wat e r way

66 67 68

69

70

71 72

73

74 75 76

77 78

Walter E. Mussell et al., “Memorandum for [South Atlantic] Division Engineer,” 30 Nov 1973 (Tenn-Tom iles, Executive Oice, Nashville District). Col. Drake Wilson to SAD Engineer, 25 Jan 1974 (TTWW Wildlife Mitigation Study folder, ile 1517–01, SAMPD-ER). Mobile District, “Alternative Plan Studies for Canal Section, Tennessee-Tombigbee Waterway,” n.d. [ Jan 1974] (TTWW Wildlife Mitigation Study folder, ile 1517–01, SAMPD-ER). See discussion in Memorandum of Support of Federal Defendants’ Motion for Judgment on the Pleadings . . . Environmental Defense Fund, et al. v. Cliford Alexander, et al., U. S. District Court for the Northern District of Mississippi, 2 July 1980; and interview, author with George H. Atkins, 18 Nov 1986. Glover Wilkins, “Administrator’s Report, April–June 1970” (1970 Working Correspondence folder, ile 2–1000–18, GWTTW Archives). Wilkins elaborated on this theme to Mobile’s environmental planners: see Glover Wilkins to E.A. Drago, 31 March 1970 (ile 2–1000–932, GWTTW Archives). Glover Wilkins to homas Abernethy, 25 Jan 1971 (ile 2–1000–651, GWTTW Archives). See Engineering Division, OCE, disposition form, 18 July 1975 (Canal section Conference folder, Conferences/Review Comments subile, ile 1518–01, CWPF, WPS, OCE). See Robert L. Crisp, Jr., (Tennessee-Tombigbee Coordinator at SAD), “Minutes of Lock D Conference, Tennessee-Tombigbee Waterway,” 26 May 1977 (TennTom microilm, set I, reel 1, frame 452). Fred hompson, memorandum for chief of Mobile District Engineering Design Branch, 16 Sept 1975 (Elimination of Lock D folder, ile 401–07, SAMPD-EI). David H. Webb to Mobile District Engineer et al., 22 April 1977 (Tenn-Tom microilm, set I, reel 1, frame 776). Mobile District Engineering Division, Project Engineering and Coniguration section, memorandum, 7 Dec 1976 (Canal Section/August 1976–May 1977 folder, ile 1503–03, SAMRE). Mobile District Real Estate Division, memorandum, 1 Feb 1977 (Canal Section/ August 1976—May 1977 folder, ile 1503–03, SAMRE). Mobile District, “Tennessee-Tombigbee Waterway, Preliminary Evaluation of Design Concepts for Canal section Involving Lock D,” Jan 1977 [background

247

Chapter 7

79 80 81 82

83

84

85

86 87

88

89

paper prepared for Board of Environmental Consultants] (Board of Environmental Consultants/Jan 1977 Meeting folder, ile 1501–03, SAMPD-EI). Robert L. Crisp, Jr., “Minutes of Lock D Conference, Tennessee-Tombigbee Waterway,” 26 May 1977 (Tenn-Tom microilm, set I, reel 1, frame 452). See ibid., frame 453; and Robert L. Crisp, Jr., memorandum for record, 8 Feb 1977 (Tenn-Tom microilm, set I, reel 1, frame 1070). Interview, author with Jack H. Bryan, 17 Nov 1986; and Kemp, “Engineering Aspects,” part I, 8–9. Mobile District Engineering Division, “Status Report, Tennessee-Tombigbee Waterway,” 13 Aug 1971 ( James D. Williams papers, University of Alabama Archives). See “Minutes of Conference on Aliceville Lock and Dam Project, TennesseeTombigbee Waterway,” 15–16 March 1972 (Aliceville L&D Conference folder, Conferences/Review Comments subile, ile 1518–01, CWPF, WPS, OCE). U.S. General Accounting Oice, To Continue or Halt the Tenn-Tom Waterway? Information to Help the Congress Resolve the Controversy (Washington: General Accounting Oice, 15 May 1981), 2–3. For a brief description of this section of the waterway, see Cynthia A. Drew, “he Tennessee-Tombigbee Waterway: Divide section,” Military Engineer 77 (May–June 1985): 170–174. In Jimmy Bates, memorandum for record (Conference on Tennessee-Tombigbee Waterway, 14 July 1971), 29 July 1971, 3 (Other Agencies Coordination, Nashville District/TTW folder, ile 1501–07, SAMPD-EI). “An Atomic Blast to Help Build a U.S. Canal,” U.S. News and World Report 54 (20 May 1963): 14. William H. Stewart, Jr., he Tennessee-Tombigbee Waterway: A Case Study in the Politics of Water Transportation (Tuscaloosa: Bureau of Public Administration, University of Alabama, 1971), 466. Tennessee-Tombigbee Waterway Development Authority, Newsletter, 1 (26 Sept. 1963): 2 (copy in Greenough/Mobile Harbor folder, ile 18081, Mobile Municipal Archives). U.S. Army Corps of Engineers, Mobile District, Environmental Statement: Tennessee-Tombigbee Waterway, Alabama and Mississippi Navigation (Mobile: U.S. Army Corps of Engineers, Mobile District, March 1971), 29.

248

B u i l d i n g t h e T e n n e s s e e -To m b i g b e e Wat e r way

90

91

92

93 94 95 96 97 98 99

100

101 102 103

Geotechnical and Civil Branch, OCE, disposition form to Directorate of Engineering and Construction et al., 8 January 1974 (Divide Cut Conference folder, conferences/Review Comments subile, ile 1518–01, CWPF, WPS, OCE). Joseph M. Caldwell (chief, Engineering Division, Directorate of Civil Works), memorandum for South Atlantic Division Engineers, 21 May 1973 (Bay Springs/ 1971–1976 folder, ile 1501–07, SAMPD-EI). Lt. Col. Terrence J. Connell (Deputy Nashville District Engineer), memorandum to South Atlantic Division Engineer, 8 June 1973 (Bay Springs/1971–1976 folder, ile 1501–07, SAMPD-EI). Geotechnical and Civic Branch, disposition form to Directorate of Engineering and Construction, 8 Jan 1974. Interview, author with James G. Goad, 19 Sept 1986. Jim J. Tozzi, memorandum for the record, 14 March 1967 (folder 2, box 34, James J. Tozzi papers, Oice of History, OCE). Ibid. Mobile District, Environmental Statement, 23. Interviews, author with H. Joe Cathey, 15 Sept 1986, and Jack C. Mallory, 16 May 1986. Mobile District, Environmental Statement, 23. Under-Secretary of the Army haddeus R. Beal elaborated on these considerations in a letter to CEQ Chairman Russell E. Train dated 23 July 1971 (ALA/MISS–COE: TTWW folder, box 4, ile 429–81–37, CEQ records, WNRC, Suitland, MD). James W. Erwin (SAD Geologist), “Trip Report–Tennessee-Tombigbee Meeting with Environmental Consulting Board,” 19 Aug 1971 (Tenn-Tom Waterway, General/ 1970–1971 folder, ile 1518–01, CW–601, SAD). Bob Bryan, “Evacuated Dirt in Waterway Construction to be Placed in Valleys along the Route,” Tupelo (MS) Journal, 9 July 1974. Interview, author with H. Joe Cathey, 15 Sept 1986. See Jackson H. Ables, Jr., Divide Cut Drainage Structures, Tennessee-Tombigbee Waterway, Mississippi and Alabama: Hydraulic Model Investigation, Technical Report H–76– 18 (Vicksburg: U.S. Army Engineer Waterways Experiment Station, Hydraulics Laboratory, Oct 1976); and Kemp, “Engineering Aspects,” part II, 12–13.

249

Chapter 7

104 See “Oicial Minutes, Bay Springs Lock and Dam, Tennessee-Tombigbee Waterway, Meeting at WES,” 11–12 Sept 1975 (Bay Springs L&D Conference folder, Conferences/Review Comments subile, ile 1518–01, CWPF, WPS, OCE). 105 Charles H. Tate, Jr., Bay Springs Canal Surge Study, Tennessee-Tombigbee Waterway, Mississippi and Alabama: Hydraulic Model Investigation, Technical Report H–78–9 (Vicksburg, MS: U.S. Army Engineer Waterways Experiment Station, Hydraulics Laboratory, June 1978); and Henry A. Malec (Environmental Quality Section, Nashville District), memorandum for the record, 22 March 1974 (Bay Springs/ 1971–1976 folder, ile 1501–07, SAMPD-EI). 106 Jackson H. Ables, Jr., Filling and Emptying System for Bay Springs Lock, TennesseeTombigbee Waterway, Mississippi: Hydraulic Model Investigation, Technical Report H–78–19 (Vicksburg: U.S. Army Engineer Waterways Experiment Station, Hydraulics Laboratory, Nov 1978). 107 U.S. Army Corps of Engineers, Mobile District, Gainesville Lock and Dam Tennessee-Tombigbee Waterway, Alabama and Mississippi: Design Memorandum No. 8 Lock and Channels (Mobile: U.S. Army Corps of Engineers, Mobile District, Sept 1969); and U.S. Army Corps of Engineers, Nashville District, Tennessee-Tombigbee Waterway, Mississippi and Alabama: Design Memorandum No. N–12, Divide Section, Bay Springs Lock and Dam (Nashville: U.S. Army Corps of Engineers, Nashville District, Feb 1977), I–8. For a general discussion of lock illing and emptying systems, see Petersen, River Engineering, 357–381. 108 Nashville Engineering Div. memorandum, 22 June 1976 (Bay Springs L&D Conference folder, Conferences/Review Comments subile, ile 1518–01, CWPF, WPS, OCE). 109 “Oicial Minutes, Bay Springs Lock and Dam,” 11–12 Sept 1975. 110 Nashville Engineering Div. memorandum, 22 June 1976; Jackson H. Ables, Jr., Filling and Emptying System for Bay Springs Lock, Tennessee-Tombigbee Waterway, Mississippi: Hydraulic Model Investigation, Technical Report H–78–19 (Vicksburg: U.S. Army Engineer Waterways Experiment Station, Hydraulics Laboratory, Nov 1978); and interview, author with Herman Gray, 16 Sept 1986. he Bankhead Lock hydraulic system is discussed in N. R. Oswalt, J. H. Ables, Jr., and T. E. Murphy, Navigation Conditions and Filling and Emptying System, New Bankhead Lock, Black Warrior River, Alabama: Hydraulic Model Investigation,

250

B u i l d i n g t h e T e n n e s s e e -To m b i g b e e Wat e r way

111

112

113

114 115

116

117 118 119 120 121 122

Technical Report H–72–6 (Vicksburg: U.S. Army Engineer Waterways Experiment Station, Hydraulics Laboratory, Sept 1972). “Addendum, Bay Springs Lock and Dam, Tennessee-Tombigbee Waterway, Meeting at WES,” 11–12 Sept 1975 (Bay Springs L&D Conference folder, Conferences/Review Comments subile, ile 1518–01, CWPF, WPS, OCE). David Parsons (Assistant Chief, Nashville District Engineering Div.) and Herman Gray (Chief, Nashville District Design Branch), minutes of the TennesseeTombigbee Project Bay Springs Lock & Dam Conference, Atlanta, Georgia, 8–9 April 1976 (Bay Springs L&D Conference folder, Conferences/Review Comments subile, ile 1518–01, CWPF, WPS, OCE). U.S. Army Corps of Engineers, Nashville District, Design Memorandum No. N–17: Tennessee-Tombigbee Waterway, Mississippi and Alabama, Divide Section, Bay Springs Initial Filling Plan (Nashville: U.S. Army Corps of Engineers, Feb 1983), 2; and interview, author with George H. Atkins, 18 Nov 1986. “Oicial Minutes, Bay Springs Lock and Dam,” 11–12 Sept 1975. Parsons and Gray, minutes of the Tennessee-Tombigbee Project Bay Springs Lock & Dam Conference. For further discussion of Nashville District designers’ initial preference for gravity-type walls, see interview, author with Herman Gray (chief of Nashville District Design Branch), 16 Sept 1986. Richard C. Armstrong, memorandum to SAD Engineer, 20 May 1976 (Bay Springs L&D Conference folder, Conferences/Review Comments subile, ile 1518–01, CWPF, WPS, OCE). Nashville Engineering Division, memorandum, 22 June 1976. Interview, author with Herman Gray, 16 Sept 1986. Ibid. Powell Williams, Jr., memorandum for SAD Engineer, 29 Sept 1972 (Tenn-Tom microilm, set I, reel 14, frame 152). Gen. Raymond, memorandum for Chief of SAD Engineering Division, 13 Dec 1972 (Tenn-Tom microilm, set I, reel 66). See U.S. Army Corps of Engineers, Nashville District, Tennessee-Tombigbee Waterway, Mississippi and Alabama: Design Memorandum No. N–12, Divide Section, Bay Springs Lock and Dam (Nashville: U.S. Army Corps of Engineers, Nashville District, Feb 1977), I/5–6.

251

Chapter 7

123 U.S. Army Corps of Engineers, Mobile District, “Brieing Data for Army Audit Agency: Tennessee-Tombigbee Waterway” (Mobile: U.S. Army Corps of Engineers, Mobile District, Jan 1976), V–45. 124 Joe J. Waits to Mobile District Engineer, 17 March 1977 (Tenn-Tom microilm, set I, reel 1, frame 673). 125 Ibid. 126 Waits to Mobile District Engineering Div., 29 June 1977 (Tenn-Tom microilm, set I, reel 1, frame 466).

252

8

Benjamin Franklin homas and the Introduction of the French Needle Dam into the United States

A

case study of an individual industrial site or process is an eicient means of producing a historical record, while at the same time elucidating basic principles that are “stock in trade” for the history of technology and, more precisely, engineering. Such cherished ideas as the evolution of historical data, and the origins of technological ideas and their transmission between individuals and even nations can be illuminated. he documentation of a historic site with measured drawings, archival photographs, and a written site history is the essence of industrial archaeology, a near relative of the history of technology. hose with a tight focus on the technological and/or scientiic aspects of particular developments, say the steam engine, constitute a fraternity, and are referred to as “internalists.” he earliest proponents of the internalist point of view arose from the ranks of engineering and science. One of the most signiicant monuments to this perspective is the multi-volume history of technology published in Britain shortly ater the end of the Second World War, and later enlarged.1 Other noteworthy books and articles undergird the internalist perspective.2 While being aware of wider or social and political aspects of the subject in the limited space available, the author has elected to take an internalist approach. Following these introductory remarks, the subject of the Big Sandy needle dams is the focus of this work. In order to provide a background, improvements of other tributaries of the upper Ohio River are discussed. he wide-ranging contributions of French hydraulic engineers are recognized in order to provide a broader aspect on improvements of other tributaries of the upper Ohio. he 253

Chapter 8

main narrative focuses on the design and construction of the irst needle dam on the Big Sandy River at Louisa, Kentucky, as part of a larger scheme to improve navigation on the river with a series of locks and dams on both the Tug and Levisa forks of this river.

La Belle Rivière Following the American Revolution, water transport and hydraulic power provided the arteries and muscles for the new nation that was entering the Industrial Revolution. Rivers, supplemented by towpath canals and roads, were the primary routes serving the new nation. Albert Gallatin, in his famous 1808 report, presented a grand vision of internal improvements that would exploit the rich resources and open land for settlement in the trans-montaine area over the Appalachian Mountains, but in a larger sense this movement would be a way of binding the new nation together.3 he Ohio River became the principal route for western commerce, stretching 981 miles from Pittsburgh to its conluence with the Mississippi River. he Ohio, together with the Missouri and upper Mississippi, feeds the lower Mississippi, creating one of the world’s most signiicant watersheds. he Ohio is the dominant tributary of these three, carrying, as it does, an enormous quantity of water. Much time and efort was spent during the nineteenth century to improve navigation on the Ohio River. Work began in 1878 with the Davis Lock and Dam at Pittsburgh, which was completed a number of years later, in 1885, ater many delays experienced with federal appropriations.4 he Davis Lock and Dam provided a slack-water harbor for Pittsburgh and began one of the largest navigation enterprises undertaken anywhere. From Davis Island to its conluence with the Mississippi, the untamed Ohio River was canalized with a series of more than ity locks and movable dams, which engaged the U.S. Army Corps of Engineers for half a century until completed in 1929.5 he dams associated with each of the locks converted the free-lowing river into a series of slack-water pools rather like a set of steps, resulting in a navigation using a natural water course rather than building a towpath canal. In the latter case the small size of even a broad canal together with the severe limitation of speed to about four miles per hour limited the capacity and economic 254

Th e I n t r o d u c t i o n o f t h e F r e n c h N e e d l e Da m

advantage of towpath canals. he speed limit was the result of trying to control erosion of the canal prism, thus steam haulage was not a major advantage in replacing horse or mule power. It is only the river navigations with the much larger capacity and eiciency that have survived into the twenty-irst century. All of the towpath canals have been taken out of service or converted to recreational use. he rivermen who complained about the lack of commerce during low water in the summer and early autumn represented a vocal lobby demanding the government do something. In a capricious way, they did not want navigation restricted by locks and ixed dams during a period of adequate low. he French movable dam appeared to provide both low-water augmentation and open navigation during favorable river stages. he principle behind the movable dam consisted of a device that could be lowered on top of a low-level ixed dam, which would efectively open the river for navigation, and could be raised to create a slack-water pool behind the dam during low water in the river. Because of political wrangling resulting in delays in beginning the Ohio River navigation system, the irst canalized river was the Great Kanawha, consisting of eleven locks and movable dams. he Great Kanawha Navigation system proved itself and encouraged other projects on tributaries of the Ohio River under the aegis of the Corps of Engineers.6 Improvement schemes were developed, including the canalization of the Big Sandy. his tributary provided the only feasible means to exploit the natural resources of this secluded region before railways penetrated the area. With meticulous care the U.S. Army Corps of Engineers documented the Ohio River project during the half-century in which construction took place, including lock and dam construction and operations. In the case of the Great Kanawha Navigation, the eleven movable dams and associated locks were replaced during the Great Depression in the 1930s with non-navigable gated dams built according to the German roller-gate system. hey are in operation to this day.7 he Army Corps of Engineers district archive provides abundant material on the locks and dams on the Ohio, especially photographs. In addition, there are holdings at the National Archives and Records Administration, which is the depository of oicial correspondence of the engineers involved in federal projects. he annual reports of the chief of engineers to Congress are an 255

Chapter 8

important and rich source of oicial information on the activities of the Corps within each district. Leland Johnson and others have mined this resource, resulting in numerous publications, as have those interested in steam boating and other commercial ventures.8 If they had relied almost totally on the Ohio River as the Great Highway, the states comprising the Old Northwest Territory would have been settled from the south to the north, giving these states a much more “southern” society. But the coming of the railways and the powerful inluence of the Great Lakes–Erie Canal system dominated the later developments in this territory, ensuring that these states carved out of the Old Northwest Territory had a northern accent, as witnessed in their strong support of the Union during the Civil War.9

Ante-Bellum Upper Ohio Valley River Improvements and Early Commerce With the opening up of the Northwest Territory to settlement, lotillas of vessels plied the western waters as settlers sought new lands,, while others engaged in commercial enterprises. his rag-tag leet employed lat boats, keel boats, push boats, canoes and, more or less, any vessel aloat on the river. Gigantic log rats called “arks” joined this colorful leet not only on the nearly 1,000 miles of the Ohio, but also on its principal tributaries such as the Big Sandy. he entire river system within the Ohio watershed was characterized by riles, shoals, rock ledges, and a plethora of boulders and snags. Without the later series of dams to control loods and provide low-water augmentation, navigating these systems was hazardous at best, impeding both settlement and commerce. he characteristics of this river and its tributaries presented a stark contrast between low water and frequently occurring freshets. hus traic was restricted to selected stages of river levels. In nearly every case of the upper Ohio River above the great falls at Louisville, early users attempted to remove the worst obstacles, but the situation was nearly intolerable and with new communities springing up along the river, together with agricultural interests, a powerful lobby formed to secure outside funding for river improvements. hey turned naturally to state resources in the beginning, which proved to be not up to the task. 256

Th e I n t r o d u c t i o n o f t h e F r e n c h N e e d l e Da m

he Topographical Engineers From an organizational point of view, the establishment of the Topographical Engineers and earlier Corps of Engineers provided a professional approach to river navigation. hese groups represented the presence of the federal government through the Corps of Engineers in ante-bellum river improvements. In a real way the Corps of Topographical Engineers became the federal commitment to the internal-improvements movement. As the name implies, selected oicers of the Corps of Engineers were identiied with mapping the northern frontier and such bodies of water as the Great Lakes and Lake Champlain. Following the War of 1812, all eight topographical engineers were discharged in 1815. An act passed in 1824 restored three topographical engineers within the Engineering Department. At the time it was recognized that the military skills of topographical mapping could be usefully employed in the development of the nation, particularly with regard to waterways and canals. he 1824 Congressional act provided a direct means for the federal government to be involved in internal improvements. Also in 1824, President Monroe signed into law a bill to improve navigation on both the Mississippi and Ohio rivers. his action was justiied on the grounds of national defense, like the interstate highway system under the Eisenhower administration. hrough this act, ten topographical engineers were hired and assigned to internal-improvements projects. Although the 1824 act has been called the irst River and Harbor Act, a subsequent act in 1826 authorized work, not only surveying, to enable the topographical engineers to improve selected rivers by clearing and deepening the channels.

Enter Stephen Harriman Long, 1784–1864 In addition to ield expeditions, Long epitomized the involvement of topographic engineers with river improvements, particularly the Ohio and the Lower Mississippi. On an experimental basis, he had erected a wing dam below Henderson, Kentucky, to narrow the channel and hence increase stream velocity, 257

Chapter 8

which would scour out a local sand bar. According to Frank Schubert, this dam functioned until 1872.10 Other innovations included the use of snag boats and, where necessary, dredges. he topographic engineers represented the only professional group capable of undertaking every aspect of river improvement. On the subject of snag boats, one would be remiss if the career of Henry Shreve, for which Shreveport, Louisiana, is named, is not mentioned. He developed and applied a catamaran boat, steam powered, for employment throughout the Ohio and Mississippi system and elsewhere. Perhaps his most notable achievement was to clear a 200-mile log jam on the Red River of the south using his snag boats. While engineer oicers were involved in hydraulic works from 1824 to 1838, the work on the Mississippi/Ohio system received the highest priority. During this period annual river and harbor acts increased the funding available, making more projects feasible. Under the authority of Secretary of War Peter B. Porter in 1838, the Board of Engineers for Internal Improvements was abolished. he board was transferred to an independent Corps of Topographical Engineers. Despite its short life of just forty-ive years, this new organization under the War Department established a very worthy record. In 1863, in the midst of the Civil War, the Corps was integrated back into the larger Corps of Engineers and then ceased to exist. By any measure, the Corps of Topographical Engineers served as the father of the great civil works of the U.S. Army Corps of Engineers, notably during the Great Depression. In a perverse way, at the time of the establishment of the Corps of Topographical Engineers, which was heralded in many quarters, Congress, in its wisdom, repealed the general survey act, which severely limited the number of public works. Topographical Engineers experimented with diferent types of cribs as breakwaters and as aids in the passage of water over sandbars. hey tried to keep up with related work in Europe and were quick to borrow promising technology, such as the use of concrete in jetty construction, a practice pioneered by the French. hey made substantial contributions to engineering, ever believing that science and technology eventually would provide answers to the challenges of controlling the nation’s waterways. And they felt that where theory failed, ingenuity would succeed. heirs was a compelling faith, suited to a young, 258

Th e I n t r o d u c t i o n o f t h e F r e n c h N e e d l e Da m

sinewy nation that sensed destiny was in its hands. It molded the nation’s attitude toward water resources development and became part of the culture of the Corps of Engineers. he Topographical Engineers did their best to meet the internal navigation needs of the nation. hrough careful surveys, they helped chart the shores of the Great Lakes and the Atlantic Ocean. hey did their construction work usually with modest, and oten patently insuicient, appropriations. From Texas to Maine, from Minnesota to Florida, “topogs” constructed piers, breakwaters, and seawalls. hey built dredges, snag boats, pile drivers, and other machinery. Blasting rock, removing bars, and clearing river obstructions, the topographer remolded the land. heir rivers and harbors improvements dotted the landscape.11 Until recently the tributaries of the upper Ohio River have not been the subjects of detailed historical studies despite their acknowledged contribution to Ohio River traic. Happily, several current publications have appeared in the literature or are in progress. Topics include the Great Kanawha Navigation, the Kentucky River development, the Muskingum, and the Little Kanawha and the Big Sandy navigations.12

Improving the Ohio River Tributaries In order to provide a context for the study of the Big Sandy Navigation, a brief look at other river improvements on the upper Ohio valley is in order. As far as river improvements are concerned, the Civil War represents a watershed in American internal navigation. Before the war, attempts to open up rivers for navigation involved channel clearance by removing obstacles and enhancing the low and depth of channels by dredging and by the use of wing dams. Charles Ellet, Jr., designer of the Wheeling Suspension Bridge, 1849, proposed the construction of levees and associated reservoirs for low-water navigation and lood control. he Humphrey and Abbott report established the levees-only position at the Corps of Engineers for the lower Mississippi.13 hese were the basic documents that guided the Corps of Engineers in its general mission. Completion of many of these slack-water improvements had to wait until the middle of the twentieth century. Even with a cleared channel, these 259

Chapter 8

rivers were wild and uncontrolled, limiting navigation to a few months per year. To those involved in river improvements it became clear that locks and dams were needed to create a slack-water system, which would create a series of pools stretching from dam to dam in a stair-step fashion in order for vessels to pass. he locks were a necessary accompaniment to the dams.he slack-water system became the hallmark of the Muskingum, Little Kanawha, the Great Kanawha and its tributaries, the Big Sandy, and the Kentucky rivers. Apart from the Muskingum navigation, perceived coal resources were the primary motivation for canalizing these tributary rivers. he Little Kanawha River, which joins the Ohio at Parkersburg, sought to provide access to what was believed to be a large and rich coal basin, located in an unsettled interior position of West Virginia. hese resources included not only coal but also salt, lumber, and other natural resources. Work was begun on a private basis and later taken over by the Corps of Engineers through a federal mandate. he Corps was responsible for the construction of the ith and inal lock, which provided access to Creston but failed to reach any farther upstream, which would have required not less than ten additional locks and dams. Nevertheless, there was traic above the slack-water pools serving Creston. here is some historical information on packet boats providing passenger service as far inland as Glenville. he discovery of oil along the Little Kanawha River in the Burning Springs and Volcano area provided an early demand for riverboats, irst steam- and then gasoline-powered boats to transport oil and supplies. Never a viable project from the inancial point of view, the navigation limped along until the end of the 1920s. here are records, however, indicating that powerboats continued to operate on the river to the end of the 1930s. Moving downstream on the Ohio River to Point Pleasant one encounters the mouth of the Great Kanawha River. his was the irst totally controlled river in the nation, under construction from 1875 to the completion of Lock and Dam No. 11 in 1898. he Great Kanawha Navigation proved to be a civil-engineering and inancial success, serving as a successful example for others seeking to provide slack-water navigation on tributaries of the Ohio River. A case in point is the history of the Big Sandy Navigation. With the deterioration of the timber/stone crib dams at locks and dams No. 4 and No. 5 on the Great Kanawha, action had to be taken either to rebuild 260

Th e I n t r o d u c t i o n o f t h e F r e n c h N e e d l e Da m

these two dams or abandon the French movable dam system of eleven locks and dams and replace them with four high-lit gated dams, which precluded open navigation. Planning began in 1928, but with no federal inancial commitment to the project the high-lit dams and associated locks did not move beyond the design stage. In the depths of the Great Depression of 1931, however, the project of four high-lit dams and locks was completed in record time by 1937 as part of the New Deal public works program. he Great Kanawha Navigation represents a highly successful public work. A measure of its success is that the Winield Lock and Dam is the busiest in the entire Corps of Engineers’ jurisdiction nationwide.14 he Muskingum Navigation, lying on the Ohio side of the Ohio River, served the purpose of moving agricultural and natural resources from the upper reaches of the Muskingum to Marietta on the Ohio. It provided water transportation from Zanesville and as far up the river as Dresden. Additionally, it was linked into the Ohio canal system; thus, coal was not a compelling reason for developing this navigation. Like the New Deal development of the Great Kanawha Navigation in the 1930s, another New Deal initiative was the Muskingum watershed project, which featured the construction of fourteen dams, completed by 1937, to mitigate looding in Zanesville and all along the valley of the Muskingum. It has been an entirely successful venture and, equally important, created a magniicent recreational area not served by other natural lakes, a worthy monument to the great public works projects of the New Deal. Before the Civil War, eight timber locks and dams built from 1855 to 1859 served to exploit coal resources, particularly cannel coal, a highly volatile coal, on the Coal River, which empties into the Great Kanawha at St. Albans, then in western Virginia. It was undertaken by a private company with William Rosecrans, of later Civil War fame, acting as chief engineer at the beginning of the project. He was only partially successful. A reinery was established to produce lamp oil from the oil-saturated cannel coal. Fierce competition with petroleum producers also supplying illuminating oil reduced the tonnage shipped during the Civil War and the slack-water system sufered a disastrous lood in 1869. he fragile timber system continued to operate until its closure in 1882. Little evidence remains today of the locks and dams, the mines, or the illuminating industry. 261

Chapter 8

Before the arrival of the Chesapeake and Ohio railway, Guyandotte Village, at the mouth of the river of the same name, lourished as an Ohio River port. While it was not on the scale of the California gold rush, surveys at the end of the 1840s revealed rich timber and coal resources. A system of mixed private and public capital was used extensively by the Commonwealth of Virginia to build turnpikes and other transportation improvements in the state. he appropriations from the Virginia legislature were distributed by a formula of threeiths state funds for a match of two-iths provided by private monies. In every case the private contribution had to be raised before state matching funds were available. his system of grants applied to the Guyandotte Navigation Company, founded in 1849. Begun in 1850, ive timber locks and dams were rapidly constructed and nearly completed in the same year. hese were cheaply built, and an engineering-design error was clearly evident when it became apparent that the slack-water pools behind Lock and Dam No. 3 did not reach the base of Lock and Dam No. 4, and that the same problem existed between Lock and Dam No. 4 and the next upstream dam at Lock and Dam No. 5. Legal troubles also dogged the company regarding land ownership. Litigation resulted in a shutdown of construction, which, in turn, caused the company to seek bankruptcy protection. A fresh start in 1853 saw a new company combine the assets of the old ive locks and dams that were repaired, and two additional locks and dams. Loads began to move downriver in 1855. At low water the channel between Lock and Dam No. 1 and the Ohio River had insuficient depth even for shallow-drat vessels. At the beginning of the Civil War in 1861 Union forces breached Dam No. 1 to gain access farther upstream. he same lood that so crippled the Coal River slack-water system destroyed the timberworks on the Guyandotte. hat lood ended slack-water navigation, never to be re-established. It did not, however, end interest in river improvements as a free-lowing river without locks and dams. While in charge of work at the Big Sandy project, engineer B.F. homas, at the request of the Corps of Engineers’ Cincinnati Oice, investigated the possibility of snag removal and deepening the channel to a minimum of four feet. In reviewing nineteenth-century river improvements on the upper Ohio, mention must be made of the engineering work on the Kentucky River. With a similar record of development, the use and improvement of the Kentucky 262

Th e I n t r o d u c t i o n o f t h e F r e n c h N e e d l e Da m

River its the pattern of the other tributaries on the Ohio River. Long before improvements were made, a vigorous trade had developed, moving raw materials downstream under favorable river levels for shipment on the Ohio River. As early as 1836 the Commonwealth of Kentucky began a slack-water system on the lower reaches of the river by building a number of locks and dams. Sharing a grand vision of a completely controlled river to reach the interior of the state, the Kentucky improvement was one of the earliest on the Ohio system. Work was begun but only the lower ive locks and dams were completed at this early stage. As was the case with other slack-water projects, the original locks and dams on the Kentucky River deteriorated badly so that by 1876 navigation efectively ended. he federal government came to the rescue with a congressional appropriation in 1880, which permitted the Corps of Engineers to repair the old system and extend the locks and dams originally envisaged to fourteen. Like similar projects, the work was drawn out over thirty years and only completed in 1917. By the time of its completion the lourishing days of steamboat commerce on the tributaries of the Ohio had nearly vanished. he system fell on hard times; nevertheless, recent activities at the end of the twentieth century focused around the possible use of this very historic system for recreational and possible hydraulic power generation.15 hese brief insights into various river improvements on the tributaries of the Upper Ohio are intended to give a context for evaluation of the Big Sandy slack-water system.

A Geographical Perspective of the Big Sandy Watershed he watershed map (Figure 8.1) provides a clear picture of the dentrated riverine system of the Big Sandy and its tributaries. Even before the completion of Lock and Dam No. 3 at Louisa in 1897 and, later, additional locks and dams on the Big Sandy and one each on the Levisa and Tug forks, considerable commercial traic had developed over the entire watershed, led by timber, which was loated downstream in rats. Before the network of coal-hauling railways in the region, penetration of the Big Sandy watershed by commercial traic was in terms of log rats, push boats, and, on the lower reaches of the Big Sandy, shallow-drat steamboats. 263

Chapter 8

he limited capacity of boats on the river meant that the coal resources located upstream could not feasibly be developed unless navigation on the river was improved. his seemed to be the most rational solution since roads, which were in an appallingly bad state, could scarcely be used for commerce. And railways were still in the future. It was indeed a wild and isolated area. he Big Sandy River and its two tributaries were far from ideal arteries of transportation and posed severe challenges to using the river. he two greatest obstacles were very low low in the river at certain times of the year, notably in late summer and early autumn, and times of high water. he movement of sand was also a notable obstacle. he course of the river was at best tortuous, with numerous shoals, rocks, and other obstacles in the channel. Nevertheless, early attempts to penetrate the region were undertaken by enterprising entrepreneurs. he Big Sandy River is formed by the union of the Levisa and Tug forks at the town of Louisa, Lawrence County, Kentucky. Combined, these streams drain a 4,600-square-mile basin in the states of West Virginia, Virginia, and Kentucky. he fall of the river is distributed as follows: Piketon to Louisa via Louisa Fork Warield to Louisa via Tug Fork Louisa to Catlettsburg via Big Sandy River

129.719 feet 61.184 feet 27.479 feet16

he Big Sandy River lows in a northwesterly direction through a narrow winding valley from Louisa to Catlettsburg, Kentucky, marking the border between the states of West Virginia and Kentucky for its entire 26-mile length. he river has a fall of 1.05 feet per mile. he bordering hills are broken nearly every mile by streams that have their sources from 2 to 50 miles distant. he average width of this part of the river is 300 feet. It is to a great extent shallow, but in many places deep pools with rocky bottoms are found. he bottomlands, varying in width are about 50 feet above low-water mark and not subject to inundations except in extreme cases. he banks are in many places clear of trees, and except where the rocky hills immediately border the river they are composed of sand so ine and uniform in composition as to be easily washed away by the currents. he erosive action of this river during loods washed the sand from the roots of the trees to such an extent 264

Th e I n t r o d u c t i o n o f t h e F r e n c h N e e d l e Da m

Figure 8.1. Big Sandy watershed map.

265

Chapter 8

that they oten become so inclined as to obstruct navigation. An examination was made a few years ago by boring into the riverbed, and rock from 8 to 20 feet below the surface of low water was found throughout the entire length of the river, except at its mouth. he natural bed of this stream is covered with sand, a deposit caused no doubt, by the large quantities of this material constantly being carried by the small rills into the various tributaries; thence it is brought into the river and eventually it inds its way to the Ohio. he numerous bars found at the mouths of the branches are constantly being carried from place to place by the successive rises in the river.17 he Levisa Fork is the principal branch of the Big Sandy River. It begins on Sandy Ridge in the Cumberland Mountains of Buchanan County in the southwestern part of the state of Virginia. his area is known in geography as the Clinch range of the Appalachian Mountain system. he river then lows in a northwesterly direction through Buchanan County, Virginia, and Pike, Johnson, and Lawrence counties, Kentucky, along a narrow, sharply winding valley for a distance of 189 miles to its junction with the Tug Fork at Louisa, Kentucky. Major tributaries of the Levisa Fork are Tom’s Creek, Paint Creek, Beaver Creek, John’s Creek, Middle Creek, and Russell Fork. he fountainhead, being so far south, gives the river an advantage over northern streams in having but little ice, a feature of considerable importance to those interested in its navigation. In ascending the river the hills increase in height. In some places the banks are composed of rock; in others of sand and clay. Where they are of the latter material, the slopes are uniform. he bottomlands, like those of the main river are above ordinary high-water mark. here are many large boulders in the river that have rolled from the bordering hills, and are obstructions that ought to be removed. he peculiar feature of this fork is the great number of rock bars.18

he Levisa Fork and Tug Fork valleys are rich in both mineral and natural resources. he entire area is geographically situated in the Appalachian region of the Carboniferous age, which includes the coalields of western Pennsylvania; West Virginia; Buchanan and Wise counties, Virginia; and the eastern part of Kentucky. he hills bordering both the Levisa Fork and Tug Fork are covered 266

Th e I n t r o d u c t i o n o f t h e F r e n c h N e e d l e Da m

with forest containing a wide variety of timber including walnut, ash, poplar, oak, and yellow pine. he Tug Fork is the other major branch of the Big Sandy River. his stream rises in the mountains of McDowell County, West Virginia, and lows in a northwesterly direction through a very narrow mountainous valley for a distance of 140 miles before joining the Levisa Fork to form the Big Sandy River at Louisa, Kentucky. Approximately 18 miles of the upper Tug Fork forms the boundary between the states of West Virginia and Virginia, while the remainder of the stream forms the border between West Virginia and Kentucky. he Tug Fork valley has the same general characteristics as the Levisa Fork valley, the major basic diference being that the bordering hills are closer to the stream. Between Louisa and the falls of the Tug, a distance of 11 miles, the average width of this fork is 180 feet. his fork, as far up as the Falls of Tug, is shallow, crooked, and narrow; so shallow during low water as to render navigation impossible; but above the falls its character changes, and it becomes a succession of pools separated by rock-bars. he hills are very steep, exposing the rocky materials of which they are mostly composed. he banks are alternately of rock and of sand, but when formed of the latter material they have been but slightly cut away by the river on account of the protection aforded by trees and plants.19

Major tributaries of the Tug Fork are Mill Creek, Rock Castle Creek, and Marrow-Bone Creek. he report of 1875 identiies the resources of the entire watershed and concludes in recognizing that the only feasible solution was a slack-water system of locks and dams, owing to the limited supply of water during several months of the year. he various cross sections where sites have been selected, except at the mouth, show favorably for the erection of locks and dams, there being always rock formations on one side and on the bottom of the river, thus giving good foundations for the locks and the main portion of the dam, while the bank at one end of the dam only will have to be protected against “wash” around the end of the dam. he fall of 157.2 feet between Piketon and Catlettsburg requires the erection of

267

Chapter 8

15 locks of 10 feet lit, while the fall of 61.2 feet between Warield and Louisa, Ky., takes 6 locks of the same lit. hese locks should be 200 feet long and 45 feet wide, with 5 feet of water on lower miter-sill. he average length of the dams will be about 250 feet.

here are 14 maps accompanying this report, whose horizontal scale in 1 inch to 1,000 feet, while the vertical scale is 1 inch to 6 feet.20 Between 1883 and 1909 the federal government built a series of three locks and dams on the Big Sandy River and one lock and dam each on the Tug Fork and Levisa Fork. hese works ceased to operate during the 1950s because of lack of use. Today the Big Sandy River is navigable from its mouth upstream for a distance of nine miles in a pool created by backwater from the Greenup Locks and Dam on the Ohio River. Coal and petrochemical products are the primary commodities shipped.

French Movable Dams Before continuing with the story of navigation on the Big Sandy and its two principal tributaries, a diversion to discuss the origins of the operation of movable dams is in order. We associate the movable dam with nineteenth-century French engineering. hese engineering designs, without exception, featured an engineering solution to permit a movable framework to be mounted on a low ixed dam; the framework was capable of being lowered to rest upon the sill of the ixed dam, thus allowing open navigation under given river levels. At low water, with the framework of the movable dam raised, a series of slack-water pools formed, extending from one lock and dam site to the next one upstream. With the movable-dam framework in a raised position, all vessels had to pass through the adjoining lock. he antecedent of the French movable dam dates back as far as the Middle Ages, when the erection of mill dams on navigable waterways was a source of conlict between mill owners and rivermen. he early solution involved the installation of either stanches or lash locks in a mill dam. he English stanch resembled a castle portcullis to secure the entrance to a castle. In the stanch, an

268

Th e I n t r o d u c t i o n o f t h e F r e n c h N e e d l e Da m

overhead framework of timber allowed a gate in the dam to be raised, resulting in a rush of water downstream. A similar result could be obtained by a lash lock, small timber scantlings called needles, which are supported on a ledge or timber sill while the upper end rests against a horizontal timber that is hinged so it can be swung aside. Ater the needles have been removed, an opening of suicient width is created to allow the passage of vessels. Opening the stanch or lash lock results in a surge of water on which a vessel can be carried downstream. Any vessel working its way upstream experiences considerable diiculty passing a dam. here is evidence that, in certain situations, capstans mounted on the shore upstream could literally pull the vessel against the current. In a similar manner, a team of oxen or horses could be used to pull a crat through the dam. As we shall see later, both the needle-dam idea and lashing to create an artiicial lood were used in the nineteenth and early twentieth centuries in Europe and in North America. Although one speaks of nineteenth-century French movable dams as a general category, the irst such structure appeared on the Lehigh Navigation in the United States in 1819. his was the brainchild of Erskine Hazard and Josiah White, whose invention, called a bear-trap dam, was the irst modern movable dam. Working on a self-acting principle, the gate could be released to permit the passage of boats illed with anthracite on the way to Philadelphia. When the loaded coal boats moved downstream, lashing accelerated their movement. Later the bear trap was used in conjunction with the Chanoine wicket dams on the Ohio River. Because of its limited width, such dams were not used on the passes either on the Ohio or on the Great Kanawha navigations. However, the bear trap proved to be an eicient means of passing ice and debris lodged behind a dam. he story now moves to France with a focus on the Yonne River. Arising in the Massif Central, the river lows in a northwesterly direction, reaching a conluence with the Seine at Montereau. For more than 400 years logs were swept on a loodtide as stanch ater stanch was released, creating a great leet of logs that eventually reached Paris.

269

Chapter 8

he French engineer associated with the government Ponts et Chaussées agency experimented with the bear trap and installed such a device on a weir at Laneuville-au-Pont on the Marne River. By letting in a controlled amount of water through a conduit, the bear trap could be raised or lowered as needed. Apparently the French model lacked durability and was abandoned and not applied elsewhere by French engineers. he entry into movable-dam technology resulted from a need to enlarge the width of the chutes in the dam. In a display of ingenuity, French engineers invented four distinct types of movable dams, namely: 1. Frame or needle weirs 2. Movable shutters or wicket dams 3. Drum weirs, barrages à tambour 4. Pontoon weirs, barrages à ponton

We shall now examine examples of these types, since three of the four types were not only considered by the U.S. Corps of Engineers but were installed, including the needle dam, on the Big Sandy. As early as 1834 Charles A.F. Poirée constructed the irst needle dam at Basseville at the junction of the Yonne and the Nivernais Canal. he problem that Poirée faced was to maintain the level of the river and yet allow open navigation to accommodate the “lotage” of rats of logs when the stanches were released. Figure 8.2. Bear-trap gate, Davis Island Dam, Pennyslvania.

270

Th e I n t r o d u c t i o n o f t h e F r e n c h N e e d l e Da m

In essence, the old needle stanch relying on abutments to support the horizontal beam, which, in turn, supported the vertical needles, was replaced by a series of iron trestles parallel to the low of the river. hese frameworks supported the top of the needles that rested on the sill at the bottom. By increasing the number of trestles a passage of any width could be erected. With the needles removed, the linked trestles could be lowered sideways to rest on top of the lower dam, thus providing open navigation. Among the earliest needle dams, the 1836 dam at Decize on the Loire and the following year the Epineau Dam on the Yonne pioneered the Poirée invention. In subsequent years numerous needle dams appeared in not only France but also Belgium and as far away as Russia. he Decize Dam displayed a height of 6.2 feet with the trestles being spaced at 3.3 feet. he author has been fortunate to have inspected numerous needle dams in France, all of which displayed a rather low head of less than ten feet. In America there are two notable examples of the Poirée needle dam, the irst being the set of ive locks and dams on the Big Sandy River. Of these only the base masonry of the ixed dam and the lock survives at Louisa, Kentucky. In Figure 8.3. Chanoine shutter dam at Port à l’Anglais, River Seine, France.

271

Chapter 8

an efort to divert loodwater from inundating New Orleans the massive Bonnet Carré spillway was constructed to pass water from the Mississippi River into Lake Pontchartrain. It is without question the largest needle dam structure ever built, being 7,000 feet long, running parallel to the Mississippi River, and not acting as a conventional dam, but as a means of releasing water during lood stages on the Mississippi River. he Poirée idea was taken up with alacrity by members of the French Corps of Engineers in large part because these adaptations were developed in response to problems encountered in maneuvering needle dams in high water. In the case of the early dams the needles were handled manually and were easily broken. A sample in the author’s possession measures 3 inches by 2¼ inches, with a length of 8 feet. It was clearly hazardous duty for dam tenders to remove the needles during high water, and, in fact, if the water reached the footbridge on the trestles, operations ceased. Auguste Boulé replaced the Poirée needles with removable panels, which could be installed with a mobile, manually operated crane on the footbridge or by a maneuver boat. Boulé’s irst application was tried at Port-à-l’Anglais in 1875. Although of limited application in North America, there is a rare example at Lock and Dam No. 11 on the upper reaches of the Muskingum River, installed as part of the rebuilding of this earlier waterway. All of the other dams on this navigation were of the ixed-crib type, rock-illed dams. To contend with the possibility of looding above the new location a Boulé dam was designed to obviate looding. A notable Boulé dam was installed at the falls of the Ohio at Louisville, Kentucky. Another French engineer, Caméré, devised a roll-up panel, rather like a garage door, which was installed at Port-à-Villez during 1876–1880. his began the limited application of this type. It appeared in North America in Manitoba, Canada, on the Red River of the North. No other example is known. Movable shutter dams, also of French origin, have a parallel history to the needle dams and their derivatives. he shutter concept is simply a panel or gate imposed in a stream, which can be raised or lowered as required. Developed by hénard between 1832 and 1837, the gate pivoted at the top of the ixed dam sill. A guard gate was placed upstream, which allowed the main gate to be raised without working against the hydraulic head of the pool. A closely related 272

Th e I n t r o d u c t i o n o f t h e F r e n c h N e e d l e Da m

shutter type was the Betwa, used by British engineers for irrigation projects in India. Following hénard’s retirement the concept reached a new stage of development when Jacques Chanoine took over his work. Transformed under Chanoine, a new self-acting dam was introduced at mid-century. he irst application of the new device appeared in 1852 at Conlans-sur-Seine. he Chanoine modiication of the shutter was elegant and simple, resulting from moving the pivot point from the base of the dam to a point where the wicket or gate would rotate when the water reached the top of the wicket. With the wicket “on the swing” water was released, allowing the wicket strut supporting the wicket and iron framework supporting the footbridge to be lowered without coping with the hydraulic pressure at lood stage. Without question, this device became the preferred movable dam in the hands of the U.S. Army Corps of Engineers. On the Ohio alone there were more than ity Chanoine dams. Rare survivors include operating wicket dams on the lower reaches of the Ohio and the Illinois River Navigation. At a lock and dam on the upper Ohio River at Hannibal, an outdoor exhibition displays these wickets and a maneuver boat, which was used to raise and lower the dam. he story would remain incomplete without mention of various drum dams. Such dams were investigated by Corps of Engineers oicers who were aware of the Desfontaines drum dam in France. Since engineer Desfontaines served as chief engineer of the Marne Navigation, it is therefore not surprising that a drum dam of his design irst appeared on this canal in 1857, followed by nine more erected between 1861 and 1867. Like the Chanoine wicket dam, the drum dams could also be made to be self-acting by letting valve-controlled water behind the drum raise the dam. Hiram Chittenden, who rose to the rank of Brigadier General in the U.S. Army Corps of Engineers, modiied the Desfontaines dam and his design was installed at locks and dams No. 2 and No. 3 on the Monongahela River in Pennsylvania. American rivers, typically carrying heavy loads of sand and silt together with other debris, tended to clog up the pit in which the drum descended, preventing open navigation. hus, with high maintenance required to keep the drum dam functioning, it is little wonder that this device failed to achieve widespread popularity. 273

Chapter 8

Another French engineer invented the pontoon weir or dam. his type of self-acting dam is attributed to Krantz and irst appeared in 1868. Engineers of the Corps were aware of the pontoon dam, but they never adapted it to American inland waterways. Figure 8.4. Boulé gate at Suresnes Dam, France.

274

Th e I n t r o d u c t i o n o f t h e F r e n c h N e e d l e Da m

Engineers in America learned of French movable dams from articles that appeared in professional journals such as the Annales des Ponts et Chaussées. West Point oicers trained in reading French would have had little diiculty reading papers in these French journals. Nevertheless, many details in the construction and operation of these revolutionary dams were lacking and as a result such leading engineers as William E. Merrill, Addison Scott, and, most importantly for our story, Benjamin Franklin homas visited France to glean additional engineering insight. It is an outstanding example of the transfer of technology from one country to another.

Benjamin Franklin homas and the Needle Dam at Louisa, Kentucky he U.S. Army Corps of Engineers was responsible for the nation’s navigable waterways, in much the same way that the interstate highway system during the Eisenhower administration was justiied, in part, as a national defense system. he uninformed reader may well deduce that the work of building and operating the nation’s waterways was in fact in the hands of soldiers under the direction of engineer oicers. his is not, in fact, the case. he responsibility Figure 8.5. Chittenden drum dam.

275

Chapter 8

for the inland waterways rested on the shoulders of commissioned engineering oicers, many of whom were West Point graduates. he Corps’s responsibility was divided into a number of districts and at this level civilian civil engineers dominated the scene. he high reputation of the Army Corps of Engineers rests largely on the civilian corps of engineers. Numerous civilian engineers added luster to the reputation of this federal agency. In the case of the Great Kanawha Navigation, the team of William P. Craighill, a career army oicer, and Addison Scott, a civilian engineer, can in large measure take credit for the highly successful canalization of this coal-hauling waterway. Consisting of eleven locks and dams using French movable-dam technology, this was America’s irst river navigation controlled by locks and movable dams. William P. Craighill later became a general and chief of engineers. His support of Scott is clearly the key to the great success of the Great Kanawha Navigation, which continues to serve the nation as one of the busiest waterways in the entire Corps of Engineers’ system. As an acknowledged expert in hydraulic engineering, Benjamin Franklin homas at times was assigned to and involved in projects on the Kentucky River, improvements on the Guyandotte, Little Kanawha River, and elsewhere in the Corps of Engineers’ districts. During World War I, all of the Corps of Engineers’ oicers were transferred to regular army units, their places illed with qualiied civilian engineers. It was during this period that B.F. homas became district engineer for the Cincinnati District. Together with D.A. Watt he published a highly regarded text on river improvements that was recognized internationally. he coveted Norman Medal of the American Society of Civil Engineers was bestowed upon homas for his A.S.C.E. papers on river improvements.21 It is not diicult to see why he became responsible for the Big Sandy Navigation project.

Benjamin Franklin homas, 1853–1923 homas was an acknowledged expert on river navigations in the Ohio Valley, especially on the technology of movable dams. He introduced the irst needle dam in America while in charge of the Big Sandy River improvement. It was during this

276

Th e I n t r o d u c t i o n o f t h e F r e n c h N e e d l e Da m

work that he visited France to study French hydraulic engineering. In addition to his work on the Big Sandy, he consulted on the Kentucky River improvement, the construction of the so-called “government lock and dam” on the Little Kanawha Navigation, improvements on the Guyandotte River, and elsewhere. Born near Ironton on the Ohio River in May 1853, homas graduated from the National Normal University in Lebanon, Ohio, in 1870 as a civil engineer. Following surveying work he returned to Ohio to accept an appointment as engineer in charge of the Chattaroi Railroad with the title of assistant engineer. Leaving private practice in 1883, he joined the engineering department of the U.S. Army Corps of Engineers as a civilian engineer. His most important engineering assignment was the construction of the French-type needle dam and adjacent lock at Louisa, Kentucky, on the upper reaches of the Big Sandy River. Following completion of the needle dam in 1897, homas’s lengthy paper on movable dams, published in 1898, won the prestigious Norman Medal of the American Society of Civil Engineers. His reputation was further enhanced with his book Improvement of Rivers with D.A. Watt. It became a leading text in the ield for many years, with a second edition appearing in 1918. In addition to the Norman Medal, he was elected an honorary member of A.S.C.E. he marriage of Ada Rice to B.F. homas took place in 1880. he couple’s only daughter, Heloise, married George F. Gunnell. Although little known in the profession in the twenty-irst century, B.F. homas must be credited as one of the outstanding engineers of the late Victorian period. he River and Harbor Act of 1874 directed the Secretary of War to undertake a survey of the Big Sandy below Pikeville, Kentucky, and from the river’s mouth to Pikeville. As published in the chief of engineers’ annual report of 1875, part 1, the author James E. Bell, civil engineer, was placed in charge of the survey, It is evident that the only feasible way of procuring a suicient supply of water for navigation, and especially for a navigation by coal barges, is to canalize the river by means of locks and dams. In doing this we at once have our choice of two methods—the French method recently invented, of movable dams, and the method in use on the Monongahela, Muskingum, Kentucky, and other rivers,

277

Chapter 8

of permanent dams. he irst method is decidedly the better of the two where the natural navigation lasts for several months continuous; but its greater cost of establishment, and of maintenance, naturally causes the selection of the second method for rivers on which the total amount of navigation is not great. As the Big Sandy, at least at present, is decidedly a river of small navigation, I have chosen the method of permanent dams as the proper one under existing circumstances.

he 1870s saw a clamor for entrepreneurs to improve the selected Ohio River tributaries mentioned above. he call for an investigation of the Big Sandy river system followed the cautious approach taken by the army engineers in improvement projects. his gave them basic data for designing hydraulic structures and undertaking the work of clearance and improvements. he survey data served as the basis for the design of locks and dams and provided cartographic information not available otherwise, as well as low data and geological information. he Big Sandy watercourse posed special problems. he perceived natural wealth could be exploited only by improving the river for commercial use. In many ways the volume of water carried was greatly reduced with the season. It was obvious that improving the channel increased the capacity to move cargo. he state of Kentucky undertook channel improvements on both the Big Sandy and the Levisa Fork between 1835 and 1838. he work included removing snags, blasting large boulders, and cutting chutes through rock ledges to permit the passage of log rats and push boats. Not to be outdone, the Commonwealth of Virginia chartered the Tug Fork Navigation Company in 1853 to improve navigation on Tug Fork bordering on Virginia. Additional investment in 1872 on Levisa Fork improved navigation, resulting in 1874 in a reported 3,000 passengers transported on a variety of vessels, together with an estimated 30,000 tons of freight. None of this work, however, proved to be adequate for the expected commerce following the completion of the projected slack-water system. Colonel William Emery Merrill, 1837–1891, known afectionately as “Padre,” was recognized as the father of the Ohio slack-water system, although it was not completed from Pittsburgh to the Mississippi until 1929, long ater Merrill’s death. In addition to further channel improvements, Merrill was a irm believer in the provision of ity-two locks and dams, three on the Big Sandy to Louisa, thirteen on the Levisa Fork, and six on the Tug Fork. 278

Th e I n t r o d u c t i o n o f t h e F r e n c h N e e d l e Da m

Work began on the slack-water system in 1883, following a survey and clearing operations from 1875 to 1883. With funding from Congress sparse for lock and dam construction, the irst improvement, cited in the literature as Lock and Dam No. 3, was constructed below the conluence of the Tug and Levisa forks. his location for a lock provided a harbor for freight traic that could be held, awaiting favorable river levels, before moving farther downstream. Ater seven years not even the lock— the irst structure to be built in any of these lock–anddam locations—was inished. Sometime in 1890 construction began on a ixed timber-crib dam designed in the traditional manner of a series of timber cribs illed with large stones. he rivermen and those involved in the coal business objected vociferously because open navigation would be curtailed, greatly limiting the movement of timber rats and coal barges. Figure 8.6. Big Sandy River Lock No. 3, removing the pass needles at trial of dam in January 1897. 279

Chapter 8

As an acknowledged expert on movable dams, Benjamin Franklin homas assumed control of the Big Sandy River project with the title of assistant engineer, and committed to the slack-water system composed of locks and timbercrib dams. In addition to the vocal opposition of rivermen to a ixed dam at Louisa, a serious problem centered on the massive movement of sand along the river at high water. It appeared likely that the ixed dam would impound not only water, but also sand, and require constant dredging. hus, it appeared likely that the use of a series of ixed dams would be a costly error. Although homas was experienced in the construction and operation of the Chanoine wicket dams on the upper Ohio River, he recognized that these dams would not be suitable for the Big Sandy system. For the Chanoine wickets to function with the least amount of maintenance, a gap of approximately four inches is required between each of the units. A noticeable amount of leakage was acceptable on large waterways such as the Ohio River, or indeed the Great Kanawha River. During this period, homas visited France to view the current status of movable-dam engineering. Ater he returned he made a proposal that the dams on the entire slack-water system be Poirée needle dams. he essential features of the needle dam could be enlarged to accommodate much larger depths than those used in France. With greater depth, however, stronger needles would be required, of such size that they would have to be maneuvered by machinery because of their weight. he dam could be disassembled by irst removing the needles and then collapsing the trestles, thus providing open navigation and at the same time an eicient means of passing driting sand and other debris downstream. homas’s proposal was greeted with alacrity by the Corps of Engineers, which resulted in scrapping the ixed-crest design in favor of needle dams. Because the needles are closely spaced, the amount of leakage in a needle dam is far less than the typical Chanoine wicket dam. Approval was granted in 1892 for the needle dam at Louisa by then-General William P. Craighill and a board of engineers. he ixed concrete foundation of the lock and sill of the dam represented the state of the art of lock and dam construction. By the end of the nineteenth century concrete had pretty much replaced stone masonry, and even brickwork, and become the standard for Corps of Engineers’ engineering work. hus the 280

Th e I n t r o d u c t i o n o f t h e F r e n c h N e e d l e Da m

foundation of the dam at Louisa and the lock chamber can be considered as typical rather than revolutionary. his basic engineering work survives in remarkably good condition considering its age. All of the original metal ittings of the needle dam have long since disappeared with one exception: the steel lock gates survive in place, partially buried by driting sand. homas reported that the navigation pass adjacent to the lock was 130 feet long, with a 13-foot head behind the dam. he homas-designed trestles were spaced at 4 feet on centers. he control weir adjacent to the pass spanned 140 feet to support a 7-foot head of water. Because of the lower depth, the weir trestles were spread out to 8 feet on centers. An important innovation was that the trestles were fabricated as inverted “V”s, which allowed them to lie lat inside one another. his resulted in a minimum height above the concrete sill. Because of lows of less than 50 cubic feet per second at times in late summer, leakage was a problem. White pine was the wood species of choice for all the needles. In each case, for both the pass and the control weir, the needles were 12 inches wide and a length of 14 feet, 8½ inches. Each of the pass needles Figure 8.7. Lock and Dam No. 1 with needles raised.

281

Chapter 8

weighed approximately 263 pounds, making them too heavy to be manipulated by hand. A steam-driven derrick boat was used to maneuver these needles. A steam-hauled winch was used to raise and lower the metal trestles. A winching system could also be used to draw the needles together to further reduce any leakage to an absolute minimum. he shorter 8-foot, 3½-inch weir needles weighed only 80 pounds and could be manipulated by hand to adjust the pool level behind the dam. he total leakage with a head of 12 feet, 2 inches was reported as 9 cubic feet per second. hus the needle dam proved to be an excellent solution for the low low experienced on the Big Sandy River. One innovation called “repoussing” involved attaching a cable, secured to the derrick boat, to the top of a given needle. By pulling a given needle about two feet a considerable amount of water could be released to control the slack-water pool behind the dam. Selected needles could be withdrawn by hand on the weir, if needed. In an attempt to control debris from lodging against the dam, a boom was constructed several hundred feet upriver. he boom was so constructed that it could release any drit downstream when the dam had been lowered. Figure 8.8. Lock and Dam No. 1, Big Sandy River, supporting a total head of 21 feet; photo taken at low water stage with the dam completed, 1905. he steamer Enquirer is in the lock.

282

Th e I n t r o d u c t i o n o f t h e F r e n c h N e e d l e Da m

Although the Louisa Lock and Dam were the irst constructed, they were designated number three, and completed and tested in 1896. hey opened their gate to river traic on January 1, 1897. To proponents of the slack-water system, this marked the beginning of what was to be a slack-water system of locks and dams for both forks of the Big Sandy River improvement. In the early twentieth century, needle dams were completed at Catlettsburg and Cavanaugh on the main course of the Big Sandy in 1905. here followed one needle dam on the Tug Fork and one on the Levisa Fork, which were opened to traic in 1909. During the construction of locks and dams No. 1 and No. 2, the weir pass needles were replaced with Chanoine wickets. Each of the wickets was fabricated from steel and featured wooden battens to close the gap between wickets, thus signiicantly reducing lost water when the level of the pool reached the top of the wicket. By 1912 the Cincinnati District engineer called for an investigation of the amount of waterborne traic on the system and he proposed to open all of the dams and abandon the system. he source of the problem was obvious. he slack-water system failed to reach the rich coalields in West Virginia and Figure 8.9. A broken needle (aiguille) from the low height needle dam on the Nivernais Canal, France.

283

Chapter 8

Kentucky. With little possibility of completing the project, the district engineer felt that the system should be abandoned and the money saved put behind more productive projects. A more serious problem facing the project was railways penetrating the area, to the detriment of waterborne traic. Both the Norfolk and Western and the Chesapeake and Ohio railways constructed tracks into the coalields to provide easy access to market. hus, any further construction work ended by 1914. his represents for all inancial purposes the end of the grand Big Sandy slack-water navigation system. While the French needle dam appeared to be an ideal solution for improved navigation on the Big Sandy, this type of dam was seldom used in North America. With B.F. homas’s direction a needle dam was completed on Lock and Dam No. 10 on the Kentucky River with a proposal to extend the slack-water system with additional needle dams. his proposal, however, was never implemented. A needle dam was also erected on the Ouachita River in the Mississippi watershed basin. he greatest monument to this technology rests not with a movable dam, in the usual sense of the word. It was highly desirable to control looding in the lower Mississippi to prevent looding in New Orleans. To do this a loodway on the bank of the river was constructed. Consisting of large needles and stretching more than 7,000 feet, the needles could be removed and a signiicant portion of the loodwaters drained directly into Lake Pontchartrain, avoiding the city of New Orleans. his unusual structure has been used satisfactorily on a number of important looding episodes on the lower Mississippi. It remains in service to this day.22 Chapter 8 Notes 1 2

Charles Singer, E. J. Holmyard and A. R. Hall, A History of Technology vol. 1–5 (New York: Oxford University Press, 1954). Friedrich Klemm, A History of Western Technology (Cambridge, MA: M.I.T. Press, 1954); Richard Shelton Kirby, Sidney Withington, Arthur Burr Darling, and Frederick Gridley Kilgour, Engineering In History (New York: McGraw-Hill Book Co., 1956); T. K. Derry and Trevor I. Williams, A Short History of Technology (Oxford, England: Oxford University Press, 1960).

284

Th e I n t r o d u c t i o n o f t h e F r e n c h N e e d l e Da m

3 4 5 6 7 8 9 10

11

12

13 14 15 16 17 18 19 20 21 22

Leland R. Johnson, Men, Mountains, and Rivers (Huntington, WV: U.S. Army Corps of Engineers, 1977), 33–50, 130–133. Leland R. Johnson, he Davis Island Lock and Dam 1870–1922 (Pittsburgh: U.S. Army Corps of Engineers, 1985), 53–85. Ibid., 1–3. Emory L. Kemp, he Great Kanawha Navigation (Pittsburgh: University of Pittsburgh Press, 2000). Ibid., 95–249. Johnson, Men, Mountains, and Rivers. Ronald E. Shaw, Erie Water West (Lexington, KY: University of Kentucky Press, 1966), 303–397. Frank N. Schubert, he Nation Builders (Fort Belvoir, VA: Oice of History, U.S. Army Corps of Engineers, 1988), 37–48; Richard G. Wood, Stephen Harriman Long, 1784–1864 (Glendale, CA: Arthur H. Clark Co. 1966), 143–154, 203–210. Secretary of War, Annual Report of Major William E. Merrill, Corps of Engineers, For the Fiscal Year Ending 30 June 1875 (Appendix P, Survey of the Big Sandy River, Kentucky), H.R. Exec. Doc. 1, Part 2 (vol. 2, part 1), 44th Cong., 1st sess., 1876, 765. Kemp, he Great Kanawha Navigation; Leland Johnson and Charles E. Parrish, Kentucky River Development: he Commonwealth’s Waterway (Louisville, KY: Louisville District U.S. Army Corps of Engineers, 1999). Andrew A. Humphrey and Henry L. Abbot, Report on the Physics and Hydraulics of the Mississippi River (Washington: U.S. Army Corps of Engineers, 1861). Kemp, he Great Kanawha Navigation, 135–136. Johnson, Kentucky River Development. Secretary of War, Annual Report, 1875, 756. Ibid., 760. Secretary of War, Annual Report of Major D. M. Lockwood, Corps of Engineers (Appendix GG), Exec. Doc. 51st Cong., 1st sess., 1890. Secretary of War, Annual Report, 1875, 761. Ibid. Ibid., 759. Emory L. Kemp, “Stemming he Tide: Design and Operation of the Bonnet Carré Spillway and Floodway,” Essays in Public Works History, 17 (1990).

285

Figure 9.1. Sketch map of the Croton Aqueduct, from HABS/HAER.

286

9

John Jervis and the Hydraulic Design of the Old Croton Aqueduct Edward Winant and Emory L. Kemp

Having resolved on the work, they carried it forward with a degree of constancy and energy alike remarkable, so that in the space of ive years, an aqueduct was completed which, for the natural diiculties overcome, the substantial character of its structures, the very remarkable veriication, in the results, of the previous calculations of the engineers as to the low of the waters, and the quantity that could be delivered, for the extent of its course, and the abundance of its supply, may be ranked among the foremost of like undertakings throughout the world. Yet, with all this energy and perseverance, there was no rashness. he calculations of the cost, were carefully made, and it is a circumstance unparalleled probably in the history of like undertakings, and one which relects great credit on the exactness of the knowledge of the chief engineer, Mr. Jervis, and on his professional skill and idelity, that the very irst estimate he gave, ater he had made himself master of the details of the proposed work, and the experience of some few contracts, has turned out to be within, and not much difering from, the actual cost.

his statement appeared in Charles King’s A Memoir of the Construction . . . of the Croton Aqueduct . . ., published in 1843, around the time of its opening.1 Not only does King present details on construction, cost, and operation of the new aqueduct, but he places this great engineering work in the same category as the Roman aqueducts and compares it favorably with more-modern European water-supply schemes of his day.

287

Chapter 9

In many ways King’s evaluation was correct. Excepting transportation works such as the Erie Canal, the Croton Aqueduct was the greatest public work in antebellum America. It is little wonder that much has been written about this work from the time of its completion until our own day. But there has been little in the wealth of published material to explain how the aqueduct was designed and how it functioned. he engineering design of this extraordinary achievement can be considered in two parts, namely, the hydraulic design, and the structures necessary in order to ensure that the hydraulic design would supply the required amount of water to New York City. In this paper, only the hydraulic-engineering aspects of this aqueduct are considered. Before describing the salient features of the aqueduct from the Croton River to Manhattan Island, however, it is well to discuss the origins of the design methods employed for the aqueduct and the need for potable water in New York City, which is surrounded by water that is not it to drink.

New York City’s Need for Water he story of settlement on the island of Manhattan is also the story of the quest for a pure and copious supply of water for its inhabitants and businesses. hough the island is conveniently surrounded by three large rivers, the Hudson, the East, and the Harlem, its settlers have always turned to other sources, sometimes requiring great ingenuity to meet their needs. he rivers surrounding Manhattan are all subject to tides, rendering them brackish and unit for consumption. For this reason, the early settlers, both Dutch and English, resorted to small wells, natural springs, and impoundment ponds.2 As the population grew in the late eighteenth century so did the need for more adequate supplies of pure water. New Yorkers were also inding new uses for municipal water, for ire protection and disease control. Periodic ires destroyed large sections of the city. Epidemics of yellow fever and cholera also devastated New York, Philadelphia, and other cities from the end of the Revolution into the early nineteenth century. While the causes of disease were not known, beliefs of the time linked disease with ilth, and a ready supply of water was desired to wash the streets clean daily.3

288

Th e H y d r au l i c D e s i g n o f t h e O l d C r o t o n A q u e d u c t

One of the irst proposals to meet this growing need was a plan in 1774 to dig bigger wells east of Broadway and pump the water into the city with an imported Newcomen atmospheric steam engine, which had been developed expressly for pumping water and was irst successfully used to drain a mine at Dudley Castle, England, in 1712. he American Revolution, in particular the British occupation of New York City, ended this plan, but the need intensiied with the growing population.4 Following the Revolution numerous plans were proposed, and until 1798 all of them dealt with developing supplies from inside the city. Since sources on the island of Manhattan were clearly inadequate, thought turned to developing the nearby Bronx River. One plan was presented to pump 360,000 gallons of water per day from the Bronx River with a system of water wheels. Another plan proposed constructing a small dam to impound water on this river and a fourteen-mile open channel to conduct water to Manhattan, which would supply three million gallons daily.5 It should be noted that these widely difering daily amounts were based on the purest guesswork by the plans’ originators. None of these men had any special qualiications in water supply, and were merely residents interested in solving a common problem and making money from it. Money was a serious consideration, though, for while the cost of digging wells was in the thousands of dollars, conceived costs for either of these two plans ran into the hundreds of thousands of dollars. In order to secure the large amounts of capital and professional skill needed, the search for new water supplies became the province of corporations instead of interested individuals.6 he inluence of corporate interests, however, severely delayed eforts to secure new water sources. In 1799, Aaron Burr (1756–1836) incorporated the Manhattan Company and secured a legislative charter to provide water to New York City. He abandoned plans for tapping the Bronx River, using the capital to fund a bank and well water to fulill requirements of his company charter. Since Burr’s Manhattan Company held the legislative monopoly to supply water to New York City but was not doing so, the city government initiated its own plans for supplying water in 1822.7 he city turned to engineer Canvass White (1790–1834), who had

289

Chapter 9

experience with building the Erie Canal. White surveyed the routes to the Bronx River and gauged the low, reporting to the city in 1824 that it could supply over four million gallons per day (MGD). An additional four and a half MGD would be obtained from Manhattan Island ponds to provide nine MGD for the city. his igure was calculated to give the population of 450,000 twenty gallons each per day. he total cost of dams, channels, and pumping machinery was calculated at $2 million. he city, aghast at the high cost, preferred to let White complete the project with a privately funded company, but this efort failed for inancial reasons.8 Ater this, attention focused for the irst time on the Croton River, which empties into the Hudson some forty miles north of Manhattan. Although much farther than the Bronx River, it had a larger low, and geography would allow for a gravity system that never had to rely on pumping machinery. New York residents had developed a distaste for pumping machinery based on earlier water-supply schemes that were costly and inefective. he irst pumping system used in Philadelphia was not a success. hus, it is not surprising that the populace was in favor of a gravity system.9 his was how matters stood in 1833 when the city began the Croton Aqueduct project. Eforts to develop local sources were frustrated by inadequate capacity and questionable purity. An attempt to farm out the water system to private industry was misused by Aaron Burr for his own inancial gain, and so the municipality undertook to provide water for its citizens with its own resources.

heoretical Considerations he design criteria for hydraulic structures were developed by the French school of engineering, whereas traditional and practical engineering for large public works was a hallmark of British engineering in the eighteenth and nineteenth centuries. he profession of civil engineering, as contrasted to the crat of building, began in France during the reign of Louis XIV (1638–1715). Together with his famous minister Colbert, he undertook a national internal-improvements

290

Th e H y d r au l i c D e s i g n o f t h e O l d C r o t o n A q u e d u c t

program that was conceived and erected in the grand manner and intended both to serve military necessities and foster commerce in the realm. One of the most notable works of the period was the Languedoc Canal stretching from the Bay of Biscay on the Atlantic to the Mediterranean Sea. his magniicent work was under the direction of the engineer Riquet (1604–1680) and was so well known that it was included on the Grand Tour of the continent that was so popular with young English aristocrats. It was on such a tour in the middle of the eighteenth century that the Duke of Bridgewater was inspired to build the Bridgewater Canal, the irst in Britain. his started a canal mania in Britain and, at a later date, in America. he irst and by any measure the most successful early project in America was the Erie Canal, where John Jervis (1795–1885) obtained his irst experience of civil engineering.10 Sebastien P. Vauban (1633–1707), the foremost military engineer of his day, was famous for the design of fortiications. Under his initiative, the Corps des Ingénieurs du Génie Militaire was established in 1675, which provided engineers not only for military projects, but also for certain public works. With a large public-works program underway there was a shortage of engineers, which resulted in the formation of the famous Corps des Ingénieurs des Ponts et Chaussées in 1720. he Corps was enhanced by the formation of the École Ponts et Chaussées in 1747, which was reorganized in 1760 by the noted bridge engineer Jean Rudolphe Perronet (1708–1794). Not only did Génie oicers have access to a scientiic education irmly based on mathematics, but the Corps produced an outstanding lineage of teachers, researchers, and practicing engineers, especially in the ields of bridge building and hydraulic engineering. Such engineer/scientists as Antoine de Chézy (1718–1798), Pierre Louis Georges Du Buat (1734–1809), and Gaspard Clair Riche, Baron de Prony (1755–1839) from the Corps, and Daniel Bernoulli, a member of the famous Swiss family of mathematicians, derived the basic equations for low in a gravity channel (i.e., open channel low). From the equations and diagrams, it can be seen that Bernoulli concluded that the elevation of various parts of the channel, the height of the water lowing in the channel, and the velocity of the water were the parameters to be considered in open channel analysis.11

291

Chapter 9

Later in the century, the French engineer and mathematician Chézy served as an assistant to Perronet and extended the work of Bernoulli by relating the velocity of low and the shape of the open channel to the mean depth of the channel. his formula can be written in modern terms as:1

V = C √RS Where:

V is the velocity (feet per second) C is an empirical resistance √ − coeicient√ √ − √ R is the mean depth (feet) S is the slope (feet per foot)







√ −



Chézy based this formula upon the mean depth instead of the hydraulic radius, which was a concept developed by Du Buat. For large rivers such as those √ ζto the hydraulic radius. With Chézy studied, the mean depth is roughly equal λ this slight correction the formula is applicable to a wider range of channels and is still used by hydraulic engineers with the roughness coeicient determined experimentally. Apparently Chézy’s work√was−misplaced − so√that−Du Buat was − not aware of the roughness coeicient included in the Chézy formula. √ √ Pierre Louis Georges Du Buat, an eighteenth-century French nobleman and military engineer, had worked in canal and harbor development. As a result of his professional curiosity he undertook further study in determining the quantity of low in open channels. He reasoned that the driving force for low would be proportional to the efective gravity along the slope (g/S), while the resistance would vary with almost the square of the velocity. An empirical factor remained to be determined, which he did ater countless experiments on model lumes. He then published an equation based on the results, which would yield the velocity of the low. he velocity multiplied by the cross-sectional area of the √ low will then give the quantity or low rate of water.

V = (√d − 0.1)(

√ng √S −Ln√S+1.6

√ √ λ

292 ζ





√ng



S −Ln S

)

Th e H y d r au l i c D e s i g n o f t h e O l d C r o t o n A q u e d u c t

Where:

V is the velocity (inches per second) d is the hydraulic mean depth (inches) g is the acceleration of gravity (386.4 inches per second) S is the slope factor (expressed as l/S) Ln is the hyperbolic logarithm function n is an empirical factor, set at 243.713 √

he hydraulic radius, which Du Buat called the “hydraulic mean depth,” is a √ to the −radius√for a closed pipe. It − hydraulic concept for open equivalent √ channels − the √ by √ perimeter√of −wall√wetted by the is calculated by dividing the area of low low. Du Buat is credited with the concept of the hydraulic radius.14 he second term in Du Buat’s equation is a constant term for any given channel. However, Du Buat determined experimentally that it was approximately 0.3 for all cases. his term, which he labeled the “legative,” was added ζ when he noted that the equation yield low along even the slightest of √ would λ grades, which was not consistent with experience. his yields, in inal form:15



V =



307( d −0.1)





s −Ln S =1.6

− 0.3(√d − 0.1)

Because Du Buat reasoned that lowing water glided over a layer of stagnant water on the boundaries, he did not include a term for surface roughness. He made such a determination ater several experiments on glass, wood, iron, lead, and earthen materials. hese materials were all new and in good condition, that is, not frayed or unduly rough, so they yielded similar roughness conditions. For a masonry aqueduct in service, the surface roughness coeicient would be signiicantly diferent. A device used extensively in aqueduct design by the Romans was the inverted siphon; this was a means of crossing a valley without the necessity of building a major bridge to carry the aqueduct as an open channel following the hydraulic grade line. It is possible to carry an aqueduct below the hydraulic grade line, which is the surface of water in an open channel, by directing it into a pipe that descends into the valley and rises on the other side. At the low point,

293

Chapter 9

the water is under maximum pressure, which, in efect, forces it up the other side of the inverted siphon and into the channel of the aqueduct. For pipe low under pressure there were a number of equations that Jervis could use for the design of a siphon and, being systematic, he compared the results of available formulae. Most of these formulae are similar, and indeed closely resemble the open√ channel formulae since the terms are hidden in the actual equation. Additionally, pressure forces can be expressed in feet of head, thus equating them with the actual slopes and the distance below the hydraulic Two √ grade line. √ of the − − √ pipe equations that Jervis used are given in Charles √ −Storrow’s √ textbook. √ −Prony’s √ formula is:16

√D j 5

Q = 38.116 Where:

Q is the low rate (cubic feetζ per second) √ D is the pipe diameter (feet) λ j is the slope of the pipe, or diference in elevation from starting reservoir to ending discharge (z) divided by the length (l) −



− √ − √ √ Prony was director of the École Ponts et Chaussées, which was reconstituted −

under Napoleon. He was interested in research investigations and produced empiric hydraulic formulae in the tradition of the École, which stressed practical engineering. He was involved√in numerous canal and drainage schemes. Prony’s formulations follow closely the work of his teacher at the École, Chézy. Johann Albert Eytelwein (1764–1848) was the founder and irst director of √ √ the German Bauakademie. his institution of − the well-known √ − was a precursor − √ √ contemporaries, √he−too√was Berlin Technische Hochschule. Like his French involved in river hydraulics. Jervis used all these equations in his design analysis of the Old Croton Aqueduct. Eytelwein gives a similar equation:



Q = 37.614 √ λ+54D D5 ζ









− 294







Th e H y d r au l i c D e s i g n o f t h e O l d C r o t o n A q u e d u c t

Again, there is no term in these equations for surface roughness. he constant coeicient in each equation was determined empirically, and includes gravitational acceleration and shape factors for the circular pipe. he British tradition was in marked contrast to the centrally organized government Corps de Ponts et Chaussées and its associated École. he industrial revolution that began in eighteenth-century Britain is best characterized as a free-enterprise movement with the government functioning to approve the formation of corporations and the issuing of shares that could be sold to the public. here was no grand Corps of Engineers and no curriculum established in civil engineering. Under the circumstances, it is not surprising that the time-honored apprenticeship system was the means of entering the profession of engineering. he apprenticeship system also fostered the practical side of civil engineering with the emphasis on “hands-on” experience. Nevertheless, there was a transfer of engineering knowledge through journals and especially in encyclopedias. John Robison was a colleague of James Watt when the latter was an instrument maker at the University of Glasgow. Robison was a professor of mechanical philosophy (we would say engineering) at the University of Edinburgh. His Encyclopedia Britannica article on hydraulics was a standard reference in the English-speaking world.17 In 1835, Charles S. Storrow’s treatise on water works for conveying and distributing supplies of water appeared. Storrow (1809–1904) was in residence at the École Polytechnique and the École Ponts et Chaussées in 1830–1832. Upon his return to America, he gathered together his notes from his sojourn in Paris and his inspection of hydraulic works on the Continent, and published his book on hydraulics in 1835. Robison’s earlier work and Storrow’s text were the most important references available to American hydraulic engineers in the antebellum period.18

he Chief Engineers he principal engineers involved in the Croton Aqueduct represent the two engineering traditions, one French and the other British. David Douglass was associated with French engineering tradition at West Point and had additional

295

Chapter 9

opportunities for formal education at Yale University. John Jervis, from upstate New York, began his career on a surveying team during the construction of the Erie Canal and came to be one of the great American engineers of the nineteenth century. He was self-taught in all phases of civil engineering and even ventured into mechanical engineering of railways during his long and distinguished career. Because such an undertaking as the Croton Aqueduct was considered to be a major construction project, the City of New York was determined to hire the best available talent for its direction. hey turned to David Bates Douglass, a consulting engineer and professor of natural philosophy at New York University and recognized regional expert in engineering. Douglass had graduated from Yale University in 1813 ater studying natural philosophy, then was commissioned into the Army Corps of Engineers and further trained at West Point. He served in the War of 1812 as an engineer, and later taught engineering at West Point. His education and teaching career came at a time when West Point was the one school in the United States that ofered formal training in engineering, based upon the French model. he inspiration for the formation of the U.S. Army Corps of Engineers was the French Corps des Ingénieurs Ponts et Chaussées. Even its early uniforms were similar to the uniforms worn by the French. he U.S. Army Corps of Engineers, like its French counterpart, was involved in large-scale public works, which provided practical experience for engineer oicers. In an unprecedented move, U.S. Army engineers were seconded to civil projects in the private sector throughout the antebellum period, such as the Baltimore and Ohio Railroad and the National Road. hus, the French inluence was apparent in antebellum American engineering. John Bloomield Jervis (1795–1885) was one of a group of leading civil engineers trained on the Erie Canal. Jervis had been a young farmhand living on the route of the proposed Erie Canal and was irst hired to help clear trees. His interest was piqued and he remained to help the surveying parties, progressing his way up through the ranks, gradually acquiring skills and ever-increasing responsibilities. At nights and during the winter stoppage of work he pursued studies in engineering topics under the direction of the canal’s principal engineers, among them Canvass White and Nathan Roberts (1776–1852). 296

Th e H y d r au l i c D e s i g n o f t h e O l d C r o t o n A q u e d u c t

Ater completion of the Erie Canal and a year of service as a maintenance engineer, Jervis moved on to other construction projects. He undertook construction of the Delaware and Hudson Canal, where he demonstrated a keen intelligence and an innovative spirit. Jervis was also employed in constructing some of the irst rail lines in North America, and designed a leading truck assembly (i.e., bogie) for locomotives to negotiate sharp curves and uneven track. Among his many projects from 1827 until his appointment as chief engineer of the Croton Aqueduct in 1835, Jervis served as chief engineer on the Chenango Canal in upstate New York. his was a feeder canal beginning in Binghamton and joining the Erie Canal at Utica. Since this was a summit-level canal, meaning it crossed two watersheds, suicient water had to be supplied to the summit level of the canal. In order to determine if there was suicient runof for the reservoir to supply the summit level, Jervis installed a series of gauges strategically placed around the catchment area to determine the amount of rainfall available. Not content with assuming the usual 30 percent water available from runof, Jervis measured the low with conventional weirs on the watercourses supplying the proposed reservoir. He found the runof to be 40 percent and quite suicient. his percentage of rainfall became a standard in the American hydrological community. he reservoir system was so successful that it also provided water for the Erie Canal at Utica. Jervis was self-taught by assiduously reading textbooks and applying his knowledge whenever possible in surveying and hydraulic engineering. hough his training was quite diferent from that of Douglass, his career was a vivid foreshadowing of the eventual merging of the two methods of engineering education in America.19

he Old Croton Aqueduct Douglass undertook a survey of local sources of water supply. He rejected the Bronx River as being inadequate, and dismissed Manhattan sources as being too polluted. He turned to the Croton River, following earlier studies, and determined it to be easily capable of providing for Manhattan’s daily needs. he survey work was undertaken in 1833–1834 under Douglass’s direction; he was joined in 1834 by John Martineau. Upon their recommendations, the 297

Chapter 9

Figure 9.2. Cross sections of the Croton Aqueduct in various materials. (F. B. Tower, Illustrations of the Croton Aqueduct [New York: Wiley and Putnam, 1843])

298

Th e H y d r au l i c D e s i g n o f t h e O l d C r o t o n A q u e d u c t

water commission reported in February 1835 in favor of the Croton River proposal. Douglass surveyed two diferent routes from the Croton River to Manhattan. One route followed the Hudson and ran for forty-seven miles; the other route was to the east and ran cross-country for forty-three miles. Both routes lent themselves to a gravity system without the need of pumps. he cost for building a covered masonry aqueduct along either route was estimated at $4.5 million.20 he choice of a covered masonry aqueduct was based on criteria of water purity and construction cost. he three choices Douglass had for channel construction were a covered masonry channel, an open earth-lined channel, or iron pipes. he pipe design was ruled out because it was not as easily expandable should the growth of the city require more water in the future, and it was a very costly alternative. Douglass also rejected the earth-lined channel because of the distinct possibility that residents along the path of the aqueduct would use it for waste disposal.21 Based on further surveys, Douglass then selected the 47-mile river route, which followed the Hudson River (see Figure 9.1). He directed detailed surveys of this route, settling on a uniform gradient of 13.25 inches per mile for the entire aqueduct. Based on this igure, he selected a “horse-shoe” cross section with top and bottom arches for the shape of the aqueduct and calculated a capacity of thirty MGD, much greater than the nine million gallons recommended on the basis for the aqueduct design. he extra capacity was intended for the future growth of the city.22 here was increasing animosity between Douglass and members of the New York City Board of Water Commissioners, who were overseeing the aqueduct. In particular, Stephen Allen, chief of the board and one-time mayor of New York, thought that Douglass was too cautious and slow in undertaking the work. For his part, Douglass claimed that the extra surveying would shorten the route and improve the alignment. Another area of contention was that Douglass wanted more autonomy from the board in technical matters. Allen believed such a move would deprive the commissioners of practically all of their authority. Allen stated that although Douglass was a skilled engineer in mathematics and theory, he lacked the practical knowledge of construction to

299

Chapter 9

be chief engineer, so in September 1836 Douglass was released from his duties and was replaced, almost immediately, by John Bloomield Jervis.23 Jervis assumed the duties of chief engineer on the Croton at the end of the construction season of 1836, and he was thus able to spend the winter reviewing Douglass’s designs and surveys and planning work for the upcoming spring. While Jervis accepted the route and shape of the aqueduct, even praising the quality of surveying done by Douglass, he was let with the responsibility of designing every structure along the aqueduct. Remarking on Douglass’s work: He [Douglass] made the location of the line of the aqueduct from the Croton river to the north bank of the Harlem River 33 miles, and determined the grade of the aqueduct at about 13-1/4 inches to the mile. It was, in the main, well located. In regard to plans of work, he proposed a cross section of the masonry of the conduit which with some modiication was adopted. So far as I have known, he prepared no speciications for the work that were approved by the commissioners.24

Jervis noted that Douglass did not proceed with the design of any of the structures for the aqueduct. He described the cross section used and also provided an illustration (see Figure 9.2): he form and dimensions of the interior of the aqueduct are as follows: he bottom is an inverted arch; the chord or span line is 6 feet and 9 inches, and the versed sine 9 inches. he masonry of the side walls rises 4 feet above the springing line of the inverted arch, with a bevel of 1 inch to a foot rise, or 4 inches on each side, which brings the width at the top of the side walls 7 feet and 5 inches; forming the abutments of the rooing arch, which is a semicircle, having a radius of 3 feet 8-1/2 inches, or a chord line of 7 feet 5 inches. It will therefore be perceived, the greatest interior width is 7 feet 5 inches, and greatest height 8 feet 5-1/2 inches. he area of the interior is 53.34 square feet. In rock tunnels the rooing arch is generally dispensed with, but the bottom and sides are formed with masonry similar to that above described. here is an exception to this form in the irst 4.949 miles of the upper end of the aqueduct, where the side walls have an extra height, on account of the bottom being depressed, to draw the water at a lower level from the Croton Reservoir.25

300

Th e H y d r au l i c D e s i g n o f t h e O l d C r o t o n A q u e d u c t

In addition to this description and illustration, Jervis also provided a table delineating the masonry and piping that carried the water over the length of the aqueduct: The length of aqueduct from the Croton Dam to the distributing reservoir is 40.562 miles—to wit:26 Masonry conduit in Westchester County [Ditto ditto] on New York Island Total Length Receiving reservoir from end of aqueduct to south-eastern efluent gate house

32.880 4.187 37.067 0.172

Distributing reservoir

0.080

Iron pipes on bridge over Harlem valley

0.275

[Ditto ditto] across Manhattan valley

0.792

[Ditto ditto] between reservoirs

2.176

Total

40.562

he route of the aqueduct was over very rugged ground, which necessitated an irregular path to take advantage of the contour of the land. In order to maintain the constant slope of the aqueduct from the dam on the Croton River to the Harlem River, Jervis constructed a number of tunnels, culverts, embankments, and aqueduct bridges. In addition, ventilators were required to ensure that the water would low under atmospheric pressure only and also to permit access so that the aqueduct could be inspected all along its length. Armed with the scientiic formulae of Prony, Du Buat, Eytelwein, et al., Jervis designed the main channel and sized the inverted siphons. His design method was fairly basic, commonly known as “trial and adjustment.” Given the nine MGD required at the time, he allowed for a growing city and chose a design value of thirty-ive MGD. He then performed hundreds of calculations, varying the size of the channel and the depth of low until he arrived at an acceptable design. Of course, his process was somewhat simpliied as Douglass had already set the slope of the aqueduct at 13.25 inches per mile (S = 4781.9) and determined the basic horseshoe shape of the channel.27 Working within these parameters, Jervis arrived at a channel 6 feet 9 inches wide and 8 feet 5.5 inches tall. It was constructed of brick set in hydraulic 301

Chapter 9

cement, with the channel bottom laid in a bed of concrete. To obtain the steady grade a great deal of earth had to be moved. his included tunneling through mountain sections and illing in earth along low spots. Where the aqueduct was above ground, the earth was built up to cover the entire channel for protection against freezing.28 In addition to ill sections, 114 culverts were required along the length of the aqueduct to permit passage of small streams or intervening roads. he largest culvert had a 25-foot span. hey were all of arched-stone construction. In addition, 33 ventilators were constructed to allow air to circulate through the aqueduct. his was an important consideration in order to maintain open channel low and prevent pressure buildups in the masonry sections. Six weirs were also included to control the amount of water in the system.29 he Manhattan Valley proved to be a major obstacle. It was over 100 feet deep and almost a mile wide. To surmount this obstacle, Jervis used two castiron pipes, each three feet in diameter, in an inverted siphon. he pipes dropped 102 feet from a gate chamber, crossed the valley, and climbed the other side, running a distance of 0.8 miles to the second gate chamber. Provision was made Figure 9.3. Cross section of the Croton Aqueduct inill. (F. B. Tower, Illustrations of the Croton Aqueduct [New York: Wiley and Putnam, 1843])

302

Th e H y d r au l i c D e s i g n o f t h e O l d C r o t o n A q u e d u c t

to install a further two pipes should the city require additional water in the future.30 Jervis demonstrated his ingenuity in the design of this siphon. he crosssectional area of the two pipes was 14.13 square feet, considerably less than the 53.35 square feet of the main channel’s cross-sectional area. In order to pass the required daily low, he lowered the downstream end of the siphon by three feet. his gave him an additional pressure driving force, which would increase the velocity through the siphons, and thus increase the low rate. Further, he wisely included weirs in the irst gate chamber to divert any excess water from entering the siphon.31 He attempted the same sort of design for crossing the Harlem River, an even larger valley. However, in this location, the City of New York required that there be no impediment to navigation. To this Jervis replied that there was no navigation on the Harlem and never would be, but nevertheless he complied with their wishes. To allow for navigation, he had either to bury the inverted siphon below the river, or build a bridge high enough to allow ships’ masts under it.32 Figure 9.4. Croton Dam, view from below. (F. B. Tower, Illustrations of the Croton Aqueduct [New York: Wiley and Putnam, 1843])

303

Chapter 9

Upper nappe 10' radius

Lower nappe 10' radius

55' radius

Jervis’s solution combined both a high bridge and an inverted siphon running on top of the bridge. he ancient Romans had used such designs. Croton’s high bridge, a 1,450-foot-long stone viaduct of iteen arch sections, was 114 feet above the high-water line of the river. his bridge carried two three-footdiameter cast-iron pipes in an inverted siphon. Jervis lowered the elevation of the tail end by two feet to improve the low rate. He said the reason for the pipes was to protect the masonry of the bridge from water damage and to reduce the height and cost of the bridge. While the bridge was under construction, water was carried across the valley in a single pipe laid across the coferdams that protected the bridge pier construction.33 One other important structure for the aqueduct was a dam across the Croton River to ensure adequate supplies through the summer months and to raise the elevation of the head of the aqueduct. Raising the elevation was perhaps the most crucial factor, as it controlled the grade line and route of the aqueduct. Jervis noted that the elevation of the dam shortened the length of the aqueduct by ive miles; otherwise the intake would have to be located much farther upstream. Based on his careful route surveys, Douglass had stipulated a dam forty feet high. Figure 9.5. Cross section of Croton Dam showing ogival construction. (Drawing by John Hriblan, West Virginia University, 2002) 304

Th e H y d r au l i c D e s i g n o f t h e O l d C r o t o n A q u e d u c t

Jervis built a masonry dam with an overlow spillway sited on the bedrock. he remainder of the valley was blocked of with a higher earth-illed dam. he aqueduct entrance was cut into natural rock along the southern side of the new reservoir, with a gate house and tunnel leading into the aqueduct.34 Originally, the overlow spillway was to be 100 feet wide, but because the bedrock did not extend that far, it was reduced to 90 feet. However, in January 1841 an unprecedented lood during construction washed away the earth dam. In rebuilding, Jervis then designed the spillway to be 180 feet wide, in order to handle such large discharges. He readily admitted that he had not anticipated such a lood in the three years he had observed the hydrology of the area. Because he was appointed at the end of the construction season in 1836, time did not permit a thorough rainfall study of the watershed such as the one he undertook on the Chenango Canal. Because the low would be discharged onto soil and not bedrock, Jervis took innovative precautions to check the momentum of the low. First he constructed the spillway in an ogival (backwards S) shape so that the water was not Figure 9.6. Entrance ventilator. (F. B. Tower, Illustrations of the Croton Aqueduct [New York: Wiley and Putnam, 1843]) 305

Chapter 9

impacting directly on the soil. In addition, he constructed a smaller wooden weir 300 feet downstream to provide a stilling basin. his stilling basin would serve to dissipate the momentum of the low in the spillway and protect the underlying soil. Both of these methods were novel to this project and both have since been adopted into standard dam-design practice.35 When the aqueduct was open, Jervis gauged the low rate and found it to be 25 percent greater than predicted by the Du Buat equation. If a discrepancy of 25 percent seems large, it was far more accurate than predictions made before the advent of formulae such as those of Du Buat or Chézy. Robison reports that for eighteenth-century aqueduct projects, the French Academy of Science erred by a factor of ive-ninths or 56 percent, and that the renowned engineer Desaguliers predicted a supply of one-sixth the actual for the Edinburgh Aqueduct, an error of 83 percent. For the time period prior to Du Buat, Robison reports,

(Above) Figure 9.7. Croton Aqueduct at Yonkers. (F. B. Tower, Illustrations of the Croton Aqueduct [New York: Wiley and Putnam, 1843])

306

Th e H y d r au l i c D e s i g n o f t h e O l d C r o t o n A q u e d u c t

“he fact has been, that no engineer could pretend to say, with any precision, what would be the efect of his operations.”36 In fact, if the correct roughness coeicient had been used in the Chézy or later formulae, the predicted low rate would have been very close to the measured value. hus did Jervis design and build the forty-one-mile Croton Aqueduct to supply New York City with pure and wholesome water. he work, started in

Figure 9.8. Aqueduct bridge at Sing Sing. (F. B. Tower, Illustrations of the Croton Aqueduct [New York: Wiley and Putnam, 1843])

307

Chapter 9

1838, was completed by 1842, although the Harlem River bridge was not inished until 1848. he total cost was over $9 million. Certainly this aqueduct was an impressive engineering feat, the irst of its kind in North America, and the equal of contemporary projects in France, such as the Marseilles Aqueduct, and elsewhere.

Conclusion From a perfunctory reading of the published material on the Croton Aqueduct one may be led to conclude that Douglass, with his formal education and familiarity with the French contribution to engineering, was the mastermind behind the project, leaving Jervis to carry out his design. his is hardly the case when one examines Jervis’s personal papers dealing with every aspect of the design and construction of the aqueduct. Although Douglass set the general shape of the cross section of the aqueduct and determined the slope of thirteen and one-quarter inches per mile, Jervis was not content simply to accept (Above) Figure 9.9. Croton Aqueduct at Harlem River. (F. B. Tower, Illustrations of the Croton Aqueduct [New York: Wiley and Putnam, 1843]) (Opposite, top) Figure 9.10. Hydraulic design of Manhattan Valley inverted siphon. (Drawing by John Hriblan, West Virginia University, 2002) (Opposite, bottom) Figure 9.11. Distributing Reservoir. (F. B. Tower, Illustrations of the Croton Aqueduct [New York: Wiley and Putnam, 1843])

308

Th e H y d r au l i c D e s i g n o f t h e O l d C r o t o n A q u e d u c t

3' additional drop in elevation

100

50

0

these design parameters. He performed literally hundreds of calculations to determine the optimum cross sections of the aqueduct and to conirm the slope established by Douglass as having the most appropriate value. It is clear from his papers that Jervis could have surveyed and designed the aqueduct without the preliminary design work of Douglass. Jervis made an important contribution to hydraulic engineering with his original ogival dam spillway and stilling basin, devices that are still used in dam design today. His ingenious use of inverted siphons at the magniicent high-level bridge and across the Manhattan Valley show him to be a master of hydraulic and structural design with a thorough grasp of the theory available in his day. 309

Chapter 9

His claim to fame rests not only on his hydraulic design but also on the structures needed to support the aqueduct, most notably the high-level bridge. His structural designs, however, are not the topic of this paper. Jervis’s career stands apart, with few exceptions, from apprentice-trained engineers who were without the beneit of engineering education and practiced in terms of an earlier crat tradition. By sheer genius and discipline, Jervis mastered the profession of civil engineering to an amazing degree by self-study, observation, and experience. hus Jervis was one of the outstanding antebellum engineers who began the transition of American engineering from its crat origins to a science-based profession. Chapter 9 Notes 1 2

3 4 5 6 7 8 9 10

11 12 13

Charles King, A Memoir of the Construction, Cost, and Capacity of the Croton Aqueduct (New York: 1843), 220. Larry Lankton, “Manhattan Life Line: Engineering the Old Croton Aqueduct, 1833–1842,” Ph.D. Dissertation, University of Pennsylvania (Ann Arbor: Xerox University Microilms, 1977), 5. Ibid., 6. King, Memoir of the Croton Aqueduct, 85–86. Ibid., 93–94. Ibid., 95. Lankton, “Manhattan Life Line,” 10–11. King, Memoir of the Croton Aqueduct, 111. Lankton, “Manhattan Life Line,” 17. Additional information on the history of French engineering in the eighteenth and early nineteenth centuries can be found in Committee of History and Heritage, ASCE, he Civil Engineer: His Origins, American Society of Civil Engineers Historical Publication No. 1 (New York: 1970), 28–32; and Hans Straub, A History of Civil Engineering (Cambridge, MA: M.I.T. Press, English ed., 1964), 118–135. Straub, History of Civil Engineering, 189; Russell A. Dodge and Milton J. hompson, Fluid Mechanics (New York: McGraw Hill, 1937), 80–81, 233–236. Dodge and hompson, Fluid Mechanics, 202, 235–236. John B. Jervis Papers (Manuscript Collection, John B. Jervis Public Library, Rome, NY).

310

Th e H y d r au l i c D e s i g n o f t h e O l d C r o t o n A q u e d u c t

14 15 16 17 18

19

20 21 22 23 24 25 26 27 28

Ven te Chow, Open Channel Hydraulics (New York: McGraw Hill, 1988), 13. John Robison, A System of Mechanical Philosophy (Edinburgh, Scotland: J. Murray, 1822). Charles Storrow, A Treatise on Water-works for Conveying and Distributing Supplies of Water: with Tables and Examples (Boston: Hilliard Gray and Co., 1835). Robison, A System of Mechanical Philosophy. Details of Storrow’s career can be found in Peter A. Ford, “Charles S. Storrow, Civil Engineer: A Case Study of European Training and Technological Transfer in the Antebellum Period,” Technology and Culture 34, no. 2 (April 1993): 271–299. A number of publications have been devoted to the career of Jervis and his engineering works, including Fitzsimmons’s edited and annotated autobiographical publication on Jervis, Daniel Larkin’s recent biography on Jervis based on the Jervis Papers, Rome, NY, and much additional primary and secondary source material. here is a dearth of information on David Bates Douglass. In the absence of a published biography the reader is referred to the Douglass Papers at Geneva, NY. Neal Fitzsimmons, he Reminiscences of John B. Jervis: Engineer of the Old Croton (Syracuse, NY: Syracuse University Press, 1971); Daniel F. Larkin, John B. Jervis: An American Engineering Pioneer (Ames, IA: Iowa State University Press, 1990); David Bates Douglass Papers, he Archives, Warren Hunting Smith Library (Hobart and William Smith Colleges, Geneva, NY); John B. Jervis Papers, John B. Jervis Collection ( Jervis Public Library, Rome, NY). King, Memoir of the Croton Aqueduct, 117. Ibid., 117. Lankton, “Manhattan Life Line,” 48–50. Ibid., 57–59. John B. Jervis, “Memoir Presented October 18, 1876,” Transactions of the American Society of Civil Engineers vol. 6, 55. John B, Jervis, Description of the Croton Aqueduct From the Dam to the Distributing Reservoir (New York: Slamm and Guion, 1842), 10–11. Ibid., 7. John B. Jervis Papers. Jervis’s calculations for all aspects of the engineering work on the Croton Aqueduct are available at the Jervis Library in Rome, New York. Jervis, Description of the Croton Aqueduct, 10–11.

311

Chapter 9

29 30 31 32 33 34 35 36

Ibid., 12–14. Ibid., 22–23. John B. Jervis Papers. Elting Morison, From Know-How to Nowhere: he Development of American Technology (New York: Mentor, 1974). Jervis, Description of the Croton Aqueduct, 19–22. Ibid., 14–16. Ibid., 16–18. Ibid., 30–31; Robison, A System of Mechanical Philosophy, 404.

312

10

he Muskingum Navigation A Wilderness Transformed But the strongest call came from the canal, for the engineers were letting the water in for the irst time today. As far as you could see, the big ditch ran, like a hill turned down and inside out. An army of Irishmen had scooped it out, cursing at the boys who threw stones at their clay pipes when they laid them up on the ground. Now they were gone and the ditch lay new and dry. But God help the dog or cat found in it when the water came down today. hey were letting it in from the river. Oh, this would be a day to remember. Folks were coming from twenty miles to see boats loating where had been only dry earth before. Conrad Richter, he Town

A

t the end of the Ice Age, 20,000 years ago, the great ice sheet covering much of Ohio and nearly half of the Muskingum watershed receded as meltwaters draining toward the south carved out a dentrated drainage system. Erosion over the millennia produced the landscape known to the irst European explorers and generations of Native Americans—a watershed draining a fertile land, but one subject to frequent and oten devastating loods. Fur traders led the way as they penetrated the rich and heavily forested region, using the Ohio River and its many tributaries as the easiest means of transportation. Although not veriied, the famous French explorer LaSalle may have explored the upper reaches of the Ohio Valley in 1669. Fur traders were increasingly active in the late seventeenth and early eighteenth centuries. For instance, Arnout Viele loated down the Allegheny and the upper reaches of the Ohio in 1692. Earlier in the decade, from 1654 to 1664, Abraham Wood explored the tributaries of the Ohio. By the eighteenth century, the French had

313

Chapter 10

penetrated the Ohio Valley utilizing the only practical means of transportation, rivers and lakes: the St. Lawrence, the Great Lakes, and the Ohio River itself. To counter the incursions of British traders and to reclaim the entire watershed of La Belle Rivière for the French Crown, the Celoron expedition was launched in 1749. To secure the claim for the French, Celoron deposited metal plates at various locations along the Ohio River, one set being buried at the conluence of Wheeling Creek and the Ohio River and another at Marietta at the conluence of the Ohio and Muskingum rivers. his vying for territory in the Ohio Valley was a relection of a larger struggle between Britain and France, culminating in the Seven Years War (a.k.a. the French and Indian War), 1754–1763.1 With the treaty of Paris in 1763, the French ceded all of French Canada and claims on the Ohio Valley. Following the American Revolution, military engineer Colonel Rufus Putnam, former chief engineer of the Continental Army, was instrumental in organizing the Ohio Company in 1786. he purpose of the company was to settle demobilized soldiers on public lands. He is celebrated as the founder of Marietta, Ohio, at the mouth of the Muskingum River in 1788.2 For security, Marietta, the irst settlement in the entire Northwest Territory, was located across the Muskingum River from Fort Harmon, which was garrisoned by federal troops. he Ohio Company received one and a half million acres that extended to the west and southward from the mouth of the Muskingum River. At about the same time as the founding of Marietta, Colonel Putnam began construction of a fortiication called Campus Martius to provide protection for the town and also to house members of the Ohio Company and their families.3 he ragtag lotilla of vessels on the Ohio River greatly increased following the opening of the Northwest Territories to settlement (1785 and 1787 mark the dates for Northwest Ordinances). Behind this seemingly unplanned and disorganized movement of people on land, and especially by rivers, was a widely shared notion of internal improvements. he idea became a signiicant movement in the irst century of the new republic. Its origins, however, had roots in the eighteenth-century Colonial era. Notable among the promoters of canals and turnpikes was George Washington, who championed the Potowmack and James River and Kanawha canals.4 his movement sought to provide

314

T h e M u s k i n g u m N av i g at i o n

a means for exploiting the rich natural resources of the territory beyond the Appalachian mountains by a system of roads, canals, and later railways connecting the east coast ports and industrial centers with the Great Lakes and the Ohio Valley. he lack of federal support, except for the National Road, during the early part of the nineteenth century did not deter the entrepreneurial spirit. Each east-coast city sought a means to connect itself with the heartland of the Midwest. New York led the way with the highly successful Erie Canal, completed in 1825. Not to be outdone, Pennsylvania launched the Pennsylvania Mainline Canal, which featured the spectacular railway inclines at Holidaysburg. Completed in 1835 ater nine years of struggle, it linked Pittsburgh at last to Philadelphia. With no direct link to the western waters, Baltimore, in a risky venture, threw its support behind the untried railroad, which by 1852 reached the Ohio River at Wheeling, Virginia (later West Virginia). Starting at the same time, July 4, 1828, President John Quincy Adams turned the irst shovel marking the beginning of the Chesapeake and Ohio Canal, intended, as the name implies, to connect Chesapeake Bay with the Ohio River at Pittsburgh. he rivalries among these alternative transportation systems in the Potomac Valley was intense. Not to be outdone by competition from northern canals, the Commonwealth of Virginia embarked upon the James River and Kanawha Canal. Even a casual view of a map of Virginia indicates the great advantage of a water route to reach the Ohio River well below Pittsburgh. With a more southern climate, such a route would be ice-free for most of the year. Even before construction began on the great public works to link the east coast with the Great Lakes and the Ohio Valley, there were men who envisioned connecting Lake Erie with the Ohio River. Among the earliest recorded were Jeferson and Washington in 1784,5 discussing connecting Lake Erie with the Ohio River as a much larger scheme to link the Ohio-Mississippi system with the Atlantic and the St. Lawrence River. Without federal leadership and aid, the transportation system that developed was highly competitive and not cooperative, resulting in an irrational system, with many canals and local railway lines closing ater limited operation. he Erie Canal, organized in 1810 under the leadership of DeWitt Clinton,

315

Chapter 10

had a profound inluence on other canals developed in America. However, it is not well known that, having failed to receive federal aid, the Erie Canal Company sought support from Ohio to strengthen its renewed appeal by a resolution from the state of Ohio urging federal support for a project of national importance.6 It would clearly beneit Ohio as goods and passengers from the east could be brought up the Hudson River, cross the Erie Canal to Bufalo on Lake Erie, and then be shipped to the Ohio ports bordering the lake, as well provide a possible link to the Ohio River by future canals in the east. here were, however, legal implications in terms of whether the federal government could be involved in these great public works, or were they strictly the responsibility of individual states? President Madison thought federal involvement was unconstitutional and would not support such initiatives. he War of 1812 efectively ended any chance of federal aid. Following the War of 1812, DeWitt Clinton again sought aid from Ohio. In response to the initiatives of Governor Worthington, the Ohio General Assembly passed a resolution in 1816 supporting the Erie Canal but struck the pledge of inancial aid from the bill. In the same year Ethan Allen Brown, later governor of Ohio, wrote to DeWitt Clinton on the possibility of a canal linking Lake Erie with the Ohio River. In his inaugural address of December 14, 1818, Brown stated, to increase industry and develop our resources, internal communications must be improved to provide for the surplus produce of our state a cheaper way to market.7

he governor strongly urged the case for a canal and that it should be built by the state of Ohio and not by private enterprise. Various bills were presented to the General Assembly and debated over the next several years, but no action was taken. It was not until January 1822 ater reports, bills, and lobbying of the legislature, that the Assembly inally passed the bill authorizing the government to appoint a commission and employ an engineer. he total amount was a niggardly $6,000 but at least the project was underway. David S. Bates was

316

T h e M u s k i n g u m N av i g at i o n

appointed the chief engineer, together with three commissioners. he report took three years to complete, with ive routes examined, namely: 1) Mahoning and Grand River route; 2) Cuyahoga and Muskingum (Tuscarawas Branch); 3) Black and Muskingum (Killbuck Branch); 4) Scioto and Sandusky; and 5) Maumee and Great Miami. he commissioners submitted their report in January 1825 urging that a canal be built. Although the project would run on borrowed money they were conident that the investment would be repaid from revenues by not later than 1837.8 By February 4, the General Assembly of the State of Ohio passed an act that outlined the construction of not one but two canals, the Ohio and Erie Canal from Cleveland on Lake Erie to Portsmouth on the Ohio River, and the Miami and Erie running from Toledo on Lake Erie to Cincinnati on the Ohio River. Funds would be raised by borrowing, at 6 percent, $400,000 in 1825 and an amount not to exceed $600,000 per annum in subsequent years. An appropriation by the General Assembly was meant to defray interest charges ater any proits were deducted. he vision was to have a canal running diagonally across the state from Cleveland/Cincinnati. From an engineering point of view, crossing the watershed at Portage near Akron posed little diiculty in providing water at the summit. his was not the case in crossing the watershed from the Scioto River valley to that of the Miami River. In fact, the singlecanal concept had to be abandoned since the divide was notably higher than the sources of either river, thus, alternate routes were considered. Beginning in the west, the Miami and Erie linked the Great Miami River below the divide of the Ohio-Mississippi drainage basin and the Great Lakes watershed and the Maumee River above the divide. Such a route would link Cincinnati to Toledo. An amendment to the original canal act of 1822 was enacted by the General Assembly of the State of Ohio on February 11, 1828, dealing with the means and the consequences of connecting the Ohio and Erie Canal with the upper reaches of the Muskingum River.9 Although improving the Muskingum for navigation was undertaken later, the Ohio and Erie Canal would beneit from river commerce developing on the unimproved Muskingum River as well as an efective means of disposing of surplus water in the reaches of the canal below the divide and above Coshocton.

317

Chapter 10

In the Muskingum watershed, the work was advancing steadily with lock walls completed at twenty-one sites between the Licking Summit and Caldersburgh as well as between Massillon and Dover, except for two locks nearing completion. As in the previous year when the board expected the opening of two sections in the Muskingum watershed, the commissioners expected the entire division to have water let in by July 1, weather permitting.10 he great loods of January 1829 dealt a severe blow to the region, especially that portion north of the Portage Summit. hese loods caused a rise of more than two feet above any previous records on the Cuyahoga and yet the canal itself sustained only limited damage, which required less than $6,000 to repair. Nevertheless, countermeasures were undertaken to mitigate future lood damage. his work included raising general levees and creating extensive waste ways, formed by lowering the towpath in selected places while, at the same time, providing protection of these channels against erosion.11 he experience of a year of boating on selected sections of the canal called for several improvements to be made to the locks. It was found that careless or unskilled navigators on the canal caused damage to the locks and, needless to say, to the banks themselves in attempting to enter a given lock. his was the most skillful part of navigating a boat on the canal. he solution, which is still a current practice, was to build training walls extending beyond the lock chambers in line with the landside lock walls, which made maneuvering into the lock much easier. In an expansive mood, the eighth annual report of the canal board, dated January 9, 1830, declared that the northern division of the Ohio and Erie Canal from Cleveland on Lake Erie to the north end of the deep cut which took the canal into the Scioto Valley was nearly complete except for some minor work between Dover and Calderburgh. his work, the report stated, was to require but a few days to complete. hus water could be let into nearly 190 miles of canal, including all of the route in the Muskingum watershed. Ater testing of various sections of the canal with a moderate amount of water, only minimum work was required on the Tuscarawas and Walhonding feeders. he commissioners were also pleased to announce that the entire line from the Licking Summit to Portsmouth on the Ohio River, a distance of

318

T h e M u s k i n g u m N av i g at i o n

119 miles, was under contract. he contracts called for overall work to be completed by June 1, 1831. he canal board announced . . . that the locks, aqueducts, and other important structures on the canal, have so far, fully answered the purposes for which they were designed, and that no serious injury has been sustained by any of them since their completion.12

Culverts

Stone

Wood

Draws for Crossing

146 1,207.35

5

14 153

50

8

6

Aqueducts

308.14

Guard Locks No.

Lit Locks No.

Ohio Canal, Main Truck

Length–miles and chains

Names of Canals and Branches

Rise and Fall Feet and Hundredths

he canal board proudly announced in the report to the General Assembly of January 22, 1833, that their charge embodied in the Act of 1825 had successfully been carried with the exception of work at Portsmouth and locks connecting the Miami Canal with the Ohio River at Cincinnati. hese two projects were inished and open for traic in 1834. In the case of the Ohio and Erie, the following details were presented in the report:13

Tuscawaras Feeder

3.20





1









1

Walhonding Feeder

1.30





1









1

Granville Feeder

6.14

1

10

1

1

1

3



1

2.58

3

28.79



1









Columbus Feeder

11.60

2

13.90

1



1

2

1

1

Total – Ohio Canal

33.36

152 1,250.04

9

16 155

55

9

10

Muskingum Side Cut

Extract from Canal Board Report to Ohio General Assembly, 1833. See Huntington and McClelland, Ohio Canals, 31.

he 1822 survey of Ohio canals provided the General Assembly with an estimate for the entire project of $3,081,880.83, whereas the total expenses

319

Chapter 10

as of December 1, 1832, ater the canal construction was completed, was $4,244,539.64, an apparently very large percentage overrun. he scope of the work, however, changed during construction, which greatly increased the inal total expenditure. Proponents of the national internal-improvements movement sought to connect the Ohio and Erie Canal with both the Pennsylvania Mainline Canal and the Chesapeake and Ohio Canal, expecting each of these to terminate in Pittsburgh. Such a link would allow waterborne commerce from Ohio to reach the East Coast. Incorporated by a private company in 1828, the Sandy and Beaver Canal joined Bolivar on the Ohio and Erie Canal in Stark County, Ohio, with the Little Beaver River, which joined the Ohio at East Liverpool only forty miles below Pittsburgh. he construction work on the twenty-three-mile waterway was not begun until 1834. Caught in the inancial panic of 1837, the work was stopped and not resumed until 1845. With a challenging tunnel to construct across the divide between the watersheds, and a lockage of 669 feet, the work was not completed until 1850. By this time, railway competition was intense, resulting in very little traic on the Sandy and Beaver. Being in service a mere two years, the only portion in operation beyond that time was the twelvemile Nimishillen link serving Canton. he state took over control in 1856.14 A second lateral canal, the Pennsylvania and Ohio, proved to be more successful. Incorporated in 1827, construction began in 1836 by a private company, with state aid, to connect Akron at the Portage Summit of the Ohio and Erie Canal with the Pennsylvania Main Line Canal and thence into Pittsburgh on the Ohio River. From Akron, the canal passed through Kent where there is a notable surviving masonry arch dam and associated lock, thence to Ravenna Summit where it passed into the Mahoning watershed and followed that river by Warren and Youngstown for a total length of approximately ninety-three miles. Finished in 1840 to Pittsburgh, it was completed late in the canal era and at the height of the railway “mania”; both of these lateral canals had little prospect of success.15 Two other branch canals were built as part of the Ohio and Erie system. he Walhonding feeder ran twenty-ive miles up the Walhonding Valley toward the center of the state in anticipation of transporting farm produce from this rich agricultural region. To make the feeder possible for navigation, eleven locks and dams with a lockage of eighty-ive feet together with two guard locks were 320

T h e M u s k i n g u m N av i g at i o n

built on the original feeder canal. Authorized in 1836, the navigation opened for traic in 1841.16 Much farther south, in the Scioto watershed, the Hocking Canal formed a ity-six-mile branch of the Ohio and Erie Canal. Built in three sections, the entire waterway opened in 1843. he irst section was constructed by a private company incorporated in 1826 under the name of Lancaster Lateral Canal. he remainder was built by the state of Ohio. he sixteen-mile section from Lancaster to Athens was put under contract in 1836, followed the next year by a stretch to Nelsonville with the remaining section begun in late 1838. At the same time, the state also purchased the Lancaster Canal. here was talk of extending the canal, completed to Athens in 1843, to Hockingport on the Ohio River, but this idea languished and the canal was destined to serve as a local feeder to the Ohio and Erie Canal, serving agricultural interests in the Hocking Valley as well as coal and salt industries stretched along its length. Like the lateral canals to the north, neither of these branches was a inancial success.17 Number and Tonnage of Steamboats Operating on the Western Rivers, 1817–1860 Year

Number

Tonnage

1817

17

3,290

1820

69

13,890

1823

75

12,501

1825

73

9,992

1830

187

19,481

1836

381

57,090

1840

536

83,592

1845

557

98,246

1850

740

141,834

1855

727

173,068

1860

735

162,735

Table of number and tonnage of steamboats operating on the western rivers, 1817–1860. Tilton R. Peale, Transportation and the Early Nation (Indianapolis, IN: Indiana Historical Society, 1982), 72.

321

Chapter 10

he canals in the United Kingdom, which began with the Bridgewater Canal in Lancashire in the middle of the eighteenth century, dominated the transportation system of that country, forming the arteries of the Industrial Revolution, and were unchallenged until the railway age in the late 1820s. hus, the British system of canals lourished for nearly three-quarters of a century. he canal age in America, in contrast, was compressed into only several decades before the advent of the railways, which became an economic threat as early as the 1840s. In fact, the Ohio canals’ enabling legislation of 1825 was only three years in advance of the ceremonial beginning of the Baltimore and Ohio Railroad in 1828. Since the railways had built increasing lengths of track in Ohio by mid-century, it is little wonder that the political leaders, businessmen, and the public in general turned increasingly from promoters of canals to enthusiastic entrepreneurs for the new technology. It was more than just talk. he options were to limit investment in the canals and let them deteriorate, lease the system to private enterprise, or get out of the canal business altogether and sell the entire Ohio canal system. Ater much discussion, the act of May 8, 1861, provided for the state-owned public works to be leased for ten years, extended subsequently for another decade.18 Although the act required the lessees to maintain the system in good repair, it failed to provide a means of establishing the condition of the canals and associated structures at the time the lease was signed. It proved impossible then to determine what responsibility for repairs and maintenance rested with the new renters. his extensive system was rented for a mere $20,075 per annum. A clear intent for leasing the system was, however, stated by Governor Young when he said, “to keep said Public Works in good and proper condition of repair and deliver up said Public Works to the State of Ohio at the termination of said lease, in like good condition.”19 As a result of the city of Hamilton illing in part of the canal basin, the lessees refused to pay their half-yearly rent, and the Board of Public Works terminated the lease. It was also during this period that a number of unproitable canals were abandoned nationally.20 To prevent abandonment, the federal government entered the scene in a meaningful way. here was renewed interest in the government taking over existing canal projects. hese included the canalization of the Big Sandy River that forms the border between the West Virginia and Kentucky, taking over 322

T h e M u s k i n g u m N av i g at i o n

the Little Kanawha Navigation from private company, and building an additional lock, designated “the government lock.”21 A similar action resulted in the Muskingum Improvement becoming a federal waterway in 1887.22 By far the most ambitious and indeed successful project was the complete canalization of the Great Kanawha River from the falls to the river’s conluence with the Ohio River at Point Pleasant. his work was begun in 1873 and completed in 1898 with eleven locks and movable dams.23 Figure 10.1. A map of Ohio canals prepared by Captain Chittenden. (“Survey of the Ohio and Erie Canal,” H.R. Doc. 278, 54th Cong., 1st Sess.)

323

Chapter 10

Unlike towpath canals with restricted dimensions and limited speed, river navigation on the Ohio and Great Kanawha had proved to be a vital network in the nation’s transportation system. Rates for hauling bulk cargos were and are unchallenged by any other system. To overcome the limitations of towpath canals, the U.S. Congress empaneled a commission in 1894 to undertake a feasibility study for building a seven-footdeep barge-canal channel across Ohio to link Lake Erie with the Ohio River. hey reported in 1896, having determined that there were three natural routes, namely the eastern route, comprising the Muskingum Navigation, already complete; improvement of the northern division of the Ohio and Erie Canal from Coshocton to Cleveland; and the central route, with the least lockage and shortest distance, which would be the most economical to operate, but this route lacked port facilities at either Sandusky on Lake Erie or Portsmouth on the Ohio River. he western route from Cincinnati to Toledo, via Dayton, utilizing the Miami and Maumee rivers would be the most expensive to build, but it would serve an industrial corridor and enjoy developed terminal facilities at both the lake and at the Ohio River. he eastern route, including the northern division of the Ohio and Erie Canal, together with the improved Muskingum Navigation was the least favorable regarding both the estimate for future traic and water supply across the Portage summit near Akron. In its defense, the commissioners projected that such a route would have large local traic in coal, clay, limestone, and agricultural products. Because of improvements already in place on the Muskingum, the eastern route was estimated to be signiicantly cheaper than the other two routes. he commissioners reported that it was feasible to construct not only a sevenfoot-deep channel, but one of ten feet in depth. his would allow the largest boats plying the Ohio River system and the smallest ships that could safely navigate the Great Lakes to use this eastern route.24 he following estimates were presented:25 As to the cost of a ten-foot-deep barge canal, the Commission reported, Eastern route

$15,042,586

Central route

20,784,451

Western route

26,865,126

324

T h e M u s k i n g u m N av i g at i o n

he lesser cost of the eastern route is due to the fact that part of it, from Zanesville to Marietta, was already improved and operated by the U.S. government. he commission estimated that the tonnage of such products as might be shipped by such a canal would not fall short of the following: Eastern route

8,000,000 tons

Central route

9,000,000 tons

Western route

7,000,000 tons

By deinition, a barge canal would employ steam-hauled tows or self-propelled barges, eliminating either animal or mechanical tractors operating on the towpath. While supporting the feasibility from an engineering point of view, and the undoubted beneits to the state of Ohio, the report did not believe that the United States government should construct such a route. In fact the scheme was never implemented, resulting in the U.S. Army Corps of Engineers being responsible in Ohio for the Muskingum Navigation only in addition to the extensive system of locks and dams on the Ohio River itself. It appears that the reasoning in the report was that such an improvement, although of beneit to the state of Ohio, was not in general a beneit to the country as a whole. From a purely economic point of view the role of the canals in the industrialization of the state of Ohio was summarized in the following table: 26 RECORD OF THE CANALS Cost of Construction

$15,967,652.69

Cost of maintenance and operation to Nov. 15, 1903

$12,063,849.47

Total cost of canal system

$28,031,502.16

Gross receipts 1827–1903

$16,953,102.98

Deicit

$11,078,399.18

Present value entire property estimated at

$15,000,000.00

Balance in favor canals

$3,921,600.82

Gross receipts as shown above

$16,953,102.98

Cost of maintenance and operation

$12,063,849.47

Net revenue to the State not including interest on loans, cost of construction or recent value of canal property

325

$4,889,253.51

Chapter 10

his table traces the economic history of the canal from the beginning of construction to 1905. Only the Muskingum Navigation continued in operation until the 1960s. he principal canals of Ohio played a very limited role in the economic life in the state of Ohio during the twentieth century. Lying entirely within the Muskingum watershed, the improvement of the Muskingum River needs to be considered as an example of attempts to control the water resources of the entire watershed for navigation purposes.

Slack-water Navigation on the Muskingum River he river boat New Orleans inaugurated steam navigation on the Ohio River in 1811 with a memorable voyage to New Orleans from Pittsburgh. She was built by Nicholas J. Roosevelt for the Fulton/Livingston partnership, which, having already pioneered steam navigation on the Hudson River, hoped to control steamboat traic on the western waters.27 A few years later, in 1816, the steamboat Elisha arrived in Charleston, Virginia, the irst steamboat to navigate the unimproved Great Kanawha River from its conluence at Point Pleasant.28 With these exploits in mind it is little wonder that interest mounted along the Muskingum for the possibility of steam navigation on that river to open up the Muskingum Valley for development, both agricultural and industrial. Most appropriately, the steamboat named for the founder of Marietta, Rufus Putnam, cast of her moorings at Marietta and headed for Zanesville some seventy miles upstream. With the river running at or near lood level, no obstructions were encountered and the gallant little steamboat reached Zanesville that evening ater twelve hours steaming against the current. he date was June 9, 1824, two years ater the approval for a survey of possible locations for canals across Ohio.29 Despite the highly visible voyage of the Rufus Putnam and considerable interest of the inhabitants of the valley, the Muskingum River was not included in the study nor subsequently in the Act of the Ohio General Assembly passed in February1824.30 Of the two principal routes selected, the Miami and Erie and the Ohio and Erie canals connecting Lake Erie with the Ohio River, it was the Ohio and Erie that utilized the Tuscarawas branch of the Muskingum to carry the canal from the summit level near Akron to the Licking Summit at the southwestern edge of 326

T h e M u s k i n g u m N av i g at i o n

the Muskingum watershed. Rather, however, than using the Muskingum River, the shortest route to the Ohio River, the canal turned west, utilizing the valley of the Scioto and the Ohio River terminus at Portsmouth, Ohio. Without substantial hydraulic improvements in the river, steamboat navigation was not feasible except under unusual conditions of river level. hus, in 1827, the Ohio General Assembly authorized the Board of Canal Commissioners to survey the Muskingum Valley for a link between the Ohio and Erie Canal and the Ohio River at Marietta. In compliance with the resolution of the General Assembly, the Board of Canal Commissioners reported, he character of the valley and the channel of the Muskingum, render it much cheaper to make a steamboat navigation in its channel, than a canal along its margin.31

he estimate presented for this improvement from the lower bridge in Zanesville to the Ohio River was $353,443.67. In order to connect the Muskingum Navigation with the Ohio and Erie Canal, the river would need to be improved from Zanesville to Dresden. If a connection could be efected, any surplus water in the canal between the Portage divide of the Licking Summit could be discharged into the Muskingum River. With a 158.5-foot descent between Dresden and Marietta, or 129.67 feet from the low-water mark in the Ohio River at Marietta to low water at Dresden, the total distance was very close to ninety miles. In order to accommodate shallow-drat steamboats, eleven locks and associated dams, with chambers 150 feet long and 34 feet wide, would be required. hus, unlike the other towpath canals being built, the Muskingum would be a river navigation achieved by a series of slack-water pools. his improvement could be efected by improving the Muskingum to its conluence with the Tuscarawas and Walhonding branches at Coshocton, where it would join the Ohio and Erie Canal. A closer connection, however, could be made with a “Side Cut” from Dresden, a distance of some two and a half miles being served with a series of three locks. In the seventh annual report of the Board of Canal Commissioners it was reported that the “Side Cut” was approved in February 1828.32 he Board of Canal Commissioners’ report also recommended that a lock and dam be constructed above Zanesville to provide a slack-water pool 327

Chapter 10

permitting vessels to reach Dresden and efect the junction with the Ohio and Erie Canal. he commissioners recommended that extensive work be undertaken at Zanesville consisting of a bypass canal two and one-half miles long with three locks and a dam in the river. he commissioners concluded that in order to construct a slack-water navigation it would require eleven locks and dams to overcome a fall of just over 104 feet from Zanesville to Marietta. In the same year that the Ohio and Erie Canal was completed, 1832, the Ohio General Assembly authorized the dam between Dresden and Zanesville at Symmes Creek at 10.7 miles above the Zanesville Dam. It appears that the Zanesville hydraulic works recommended by the commissioners were not included in the act of the General Assembly because of proprietary rights held by the Zanesville Canal and Manufacturing Company, which would undertake such works for waterpower potential at the dams. Since Figure 10.2. he steamship Jewel in a canal landing on the Muskingum River. (Gerald Sutphin collection) 328

T h e M u s k i n g u m N av i g at i o n

the company failed to complete the dams in a timely fashion as required by the Assembly, its charter was revoked and the works completed by the state of Ohio under an act of February 19, 1835.33 he scope of work of the Board of Canal Commissioners was expanded to include state-sponsored public works, hence, the name was changed to the Board of Public Works (BPW). he Act of 1835 authorized the BPW to undertake the construction of river improvements on the Muskingum from Zanesville to its conluence with the Ohio at Marietta, according to the requirements set out in the earlier act of 1827, which recommended eleven locks and dams. he situation was complicated by the General Assembly granting private parties rights to construct two dams across the river below Zanesville to provide waterpower. hese dams were located at present Philo and Lock and Dam No. 9. A further grant was made to Robert McConner to construct a dam at the present Lock and Dam No. 7.34 Figure 10.3. A map of the slack-water system of the Muskingum River. (War Department U.S. Army Corps of Engineers, Huntington District Archive, Huntington, W. Va., September 1941) 329

Chapter 10

Before the advent of the pound lock with two sets of gates at either end of a lock chamber, an access for vessels wishing to pass through mill dams was provided by a single gate that could be raised, swung open, or dropped to permit the passage of a vessel. When opened, a great surge of water was released, hence, the term “lash” lock.35 his Medieval system reappeared at the private waterpower dams on the Muskingum. It was clear that these private dams were not suitable for the proposed navigation system. To extricate itself from this situation, the State of Ohio provided compensation to the owners of the dams. he essence of the arrangement involved the owners’ forfeiting their right to operate a dam for a guaranteed perpetual right to use waterpower at the new lock and dam. Contracted in 1836 for a total of $1,627,018.20, the waterway was completed in 1841 and opened for commerce on October 1.36 By extending the navigation along the river above Dresden, the Muskingum Waterway could join the Ohio and Erie Canal, providing a connecting link between canal towpath barge traic with steam navigation on the Muskingum. It was never intended for steamboats to use the Ohio and Erie Canal but rather that the canal traic could be extended to the Muskingum and provide service to Zanesville and even Marietta. In the other direction, canal traic could move Muskingum Valley goods to Lake Erie via the Ohio and Erie Canal. Figure 10.4. Skirting canal at Zanesville showing typical early timber barges and a diminutive sternwheel towboat. (Huntington District, September 1947)

330

T h e M u s k i n g u m N av i g at i o n

At the time it seemed highly desirable to connect these two systems. he initial idea was to join the Ohio and Erie at Coshocton. With this in mind, the General Assembly issued an act entitled “to amend the Act for the protection of the Ohio Canals,” which said, in part, that “the Canal Commissions be, and they are hereby authorized and empowered to cut a navigable side cut or branch canal from the main canal, to enter the Muskingum River at or near the town of Dresden.”37 A cursory review would indicate that the link should be made at Coshocton. he report, however, concluded that “a side cut or branch canal of about 2½ miles in length from the main canal to the Muskingum at Dresden, with three boat locks, overcoming a descent of about 28 feet from the canal into the river, will also be necessary to perfect the plan.”38 hus, the side cut near the town of Dresden was located and put under contract to be inished a year hence, in 1830. he “cut” or branch canal was only two and a half miles in length. In this short distance, it crossed Tomika Creek on an aqueduct and required three locks to descend 28.79 feet from the canal to the river at Dresden. According to the January 11, 1831, report of the Board of Canal Commissioners the side cut was completed during the previous year. Slack-water navigations suitable for steamboats posed special problems in the design of the locks and the dams. he solutions evolved from current practice of towpath canals in which the standard American canal featured lock chambers 90 feet long by 15 feet in width, with a lit of about 8 feet.

Crib Dams he least-improved hydraulic structure was the widely used crib dam, which consisted of rock-illed timber cribs, built rather like an open box reminiscent of log house construction. Ater being illed with stones, the dam was faced with timber planks. In service, these crib dams resisted water pressure as gravity dams. Quick and easy to construct, a crib dam lacked the durability of more permanent masonry dams. Built by the score, timber cribs illed with stones became the American standard for stationary dams for both navigation and waterpower. here were two standard types, both used on the Muskingum, namely, the slope dam and the step dam. In the former, the cribs are decked over with a solid 331

Chapter 10

timber sheathing installed on a slope, while the step dam created a series of cribs diminishing in height as they marched downstream. All of the original crib dams on the Muskingum Navigation were of the slope-dam type, but ater a lood in 1837 the dams were subsequently repaired as step dams. he slope dam passed ice and debris more readily than the step dam, whereas under normal river levels, the step dam dissipates the discharge and causes less turbulence at the foot of the dam and farther downstream. In designing crib dams, engineers have to consider overturning of the dam, the possibility of sliding on the bottom of the river, and the excessive penetration of water between the bottom of the base of the cribs and the foundation. As a result, the bottom timbers were usually secured to a rock foundation with bolts or, in the case of foundations of soter material, the bottom members were secured to timber piles driven into the river bed. Although crib dams were occasionally built inside temporary cofer dams, the usual method was to lay the timbers during periods of low water. he preferred wood was white oak or yellow pine but any of the more solid hardwood species were used. Some of the early dams used logs or hewn timbers, but later, sawn timbers eight or ten inches square were standard. he sheathing was generally of sawn timber but occasionally hewn planks were used. At low water, at Lock and Dam No. 4 on the abandoned Little Kanawha Navigation, it is possible to see the bottom timbers of the crib dam and the bolts securing them to the rock foundation. Despite design knowledge that would provide engineers with methods for proportioning crib dams, many were built on an empirical basis. Even in the best-built crib dams, leakage was always a problem, as were timbers that were not continuously submerged rotting and requiring replacement.39 Later, concrete caps were added by the U.S. Army Corps of Engineers at locks and dams nos. 4, 5, 7, and 10. he irst ten dams ranged in length from 472 feet at Lock and Dam No. 7 to 840.2 feet at Lock and Dam No. 3, but the length of the concrete Lock and Dam No. 11 measured only 340 feet. In order to preclude looding upstream at Lock and Dam No. 11, a movable dam of the French Boulé type was installed. With the wicket panels removed, the support frames (i.e., fermettes) could be folded lat on the top of the concrete crest of the dam, allowing freshets to pass over rather than impounding water above the crest level behind the dam. Also, under looding conditions, a movable dam 332

T h e M u s k i n g u m N av i g at i o n

would permit open navigation of vessels and rats.40 Lock and Dam No. 11 was built later ater acquisition of the Muskingum Navigation by the U.S. Army Corps of Engineers. It was located above Zanesville at Ellis but below the earlier lock and dam at Symmes Creek, completed in 1838. his new location would ensure an adequate depth to the pool behind the Zanesville Lock and Dam. he Army engineers had considerable experience with French movable dams on the Great Kanawha and Ohio Rivers, but in those cases, the Chanoine wicket dam was employed and by 1929, ity-two such structures were built on the Ohio River and a further ten on the Great Kanawha River. he irst movable dams were erected and in operation before the Corps of Engineers acquired the Muskingum system in 1886, and it was not until 1910 that the Ellis Lock and Dam was opened for traic. he Boulé movable dam was a variant on the Chanoine wicket system and featured panels that could be removed panel by panel. hese wickets were supported on trestles erected on ive-foot centers with a depth over the ixed crest of four feet.

Figure 10.5. A stepped crib dam of a type used on the Muskingum slack-water system. (Edward Wegmann, he Design and Construction of Dams [New York: John Wiley & Sons,1900],150) (Overleaf ) Figure 10.6. An early view, 1887, showing repair work on Dam No. 8 ater the slack-water system was acquired by the U.S. Army Corps of Engineers. (Huntington District Archive)

333

334

335

Chapter 10

Locks In order to provide for the passage of steamboats, the locks above Marietta were built approximately 228 feet in length and 35.5 feet in width to accommodate vessels 35 feet wide and 162.72 feet in length inside the lock chamber. his size was selected to represent the current and expected steamboat dimensions for use on the Muskingum. hey were not large enough, though, to accommodate typical vessels plying the Ohio River. An Ohio River port above Lock and Dam No. 1 was of suicient dimensions (360 feet by 56 feet) to accommodate Ohio River boats and thus provide a so-called ice harbor for steamboats during the winter as well as ensuring that Marietta would serve as a river port. Ice harbors were also established at Parkersburg at the mouth of the Little Kanawha River and at Point Pleasant at the conluence of the Great Kanawha and Ohio rivers. All of the original sets of locks, including the irst at Harmon across the river from Marietta, were constructed in the technology of the day, namely ashlar masonry walls backed by rubble stonework. he bottom of the lock chamber utilized a timber deck. Since the sleepers and the deck planking were always under water, rotting of the wood was obviated. In excavating the tide lock of the Alexandria Canal in 1983, all of the deck timbers in this 1844 lock were found to be sound.41 In a sense, lock walls are a form of dam or retaining wall but, unlike these structures, they must be designed to withstand earth and water pressure imposed behind the walls as well as pressures developed when the lock chamber is full. In photographs of early locks, ashlar masonry can be clearly seen on the faces of the lock walls but, hidden behind the smooth masonry, rubble stonework was utilized to achieve the masonry wall thickness needed to withstand these pressures. he backill behind the wall was not usually compacted but allowed to settle over time.42 Many of the early canal locks in America were crib walls with the cribs illed with stone and the inside face of the chambers lined with planks. hey are a near relative of the crib dam. None of these crib-wall locks were used on either the Ohio and Erie Canal or the Muskingum Navigation.43 Lock loors were usually of timber planks spiked to a series of wooden sleepers which, in turn, were bolted to timber piles to hold the loor down 336

T h e M u s k i n g u m N av i g at i o n

when the chamber was empty. In this condition with the high water table, there was an uplit pressure that would loat the loors if they were not anchored down. Concrete was used in later lock designs like Lock and Dam No. 11 on the Muskingum.44

Lock Gates With thirty-six-foot-wide chambers, the timber gates were unusually large and heavy compared to towpath canal gates with iteen-foot-wide chambers. In the beginning, the gates were operated with geared drum winches and chains. (Balance beams traditionally used for miter gates could not be used with such large gates, necessitating a mechanical means of opening and closing the gates.) According to Raber, Malone, and Parrott, two of these winches survive on guard gates at Beverly and Zanesville.45 he most obvious means of attaching the chain would be at the top of the lock gate. Such a simple arrangement would tend to warp the gates and, in time, lead to ill-itting closure of the miters. he answer was to attach the chain at the bottom of the gate and through a series Figure 10.7. General view of Lock and Dam No. 9, Muskingum River, Huntington District, July, 1932.

337

Chapter 10

of sheaves run up the side wall where the winch was mounted. While sound in conception, this arrangement tended to collect debris and thus required heavy maintenance. hese early devices are quite rare, with extant examples in Canada on the Rideau Canal and also featured in the waterpower canal network at Lowell, Massachusetts.46 A more eicient device for opening the lock gates is the rack and pinion, which consists of a long tooth rail, or rack, pinned to the lock gate and engaged in a circular horizontal gear, and the pinion, which was actuated by a crank mounted on the lock wall. In addition to the Muskingum slack-water system, the nearest extant example is a rack and pinion on a lock in the Little Kanawha River.47 he rack-and-pinion system for opening lock gates was extensively used on the Ohio River as well as the Muskingum Navigation.

Gates While a variety of designs have been used for lock gates, such as the roller gate used on the Ohio River or the drop gate employed on the early Lehigh Canal, miter gates were the most popular and were used exclusively on the Muskingum and the Ohio and Erie Canal. In the early days, gates were framed with vertical timbers but, for most ordinary locks, horizontal timbers extending from the heel post to the miter toe were used. he toe ends were mitered so that the two gates closed in a “V” shape and it together to reduce leakage. he heel post was shaped to it the quoin in the lock-gate wall recess. White oak was the preferred wood for such gates, but yellow pine was also used. he heavy timber frames of the gate were sheathed with timber planks creating a solid wall to resist water pressure.

Lock Chamber Valves he common type of water valve used for medium- and low-head locks is the wicket, balanced, or butterly valve: all of these terms are used. On towpath canals constructed before the Civil War, the illing and emptying of canal-lock chambers was accomplished by pivoting wickets mounted in

338

T h e M u s k i n g u m N av i g at i o n

the lock gates and operated through gears and cranks at the top of the gates. While efective, this system was slow in moving water into and out of the lock chambers for large locks designed to accommodate steamboats. heir wickets are constructed with a vertical shat operated by a crank mounted on top of the gate. Traditionally, the wicket was cast iron with wrought-iron drive shats and cranks.48 A variation of the traditional lock-gate wicket was a horizontal cylinder mounted at the bottom of the gate or in the masonry below which could be rotated to let water in or out of the dam depending on the position of the upper and lower gates. he French Fontaine valve found favor in America, where it was named a cylinder or drum valve. his type of valve was installed ater the Corps of Engineers assumed control of the Muskingum Waterway in 1886. A pair of valves was mounted on either side of the miter sill and connected across the lock chamber with a culvert pierced by rectangular oriices below the miter sill at the head of the lock chamber. he valve, mounted in a conduit at the side of the lock, ills the chamber from a port or ports in the lock-chamber wall. It provides a much quicker means of illing the lock than the installation of wickets in the lock gates. he device consists of a ixed outside cylinder with a movable vertical cylinder within the outer shell. he cylinder is raised and lowered with an iron stem attached to a rigid cone mounted on top of the cylinder. Raising the cylinder allows water to pass from the conduit to lock chamber.49

Corps of Engineers to the Rescue During the irst decade following the opening of the waterway in 1841, considerable traic developed, providing revenue for the waterway, supplemented by receipts from the sale of waterpower at various locations. he improvement was a catalyst in the industrial and domestic expansion of communities dotted along the course of the Muskingum River. Even though the railway did not arrive along the river until 1888, various railway companies built lines criss-crossing the watershed. his caused a steady decline of passenger and goods traic on the waterway. As we have seen earlier, the decline also gripped the entire Ohio Canal system, prompting the General Assembly

339

Chapter 10

of the State of Ohio to lease the system at the beginning of the Civil War in 1861, an arrangement that lasted until 1878. In an extraordinary arrangement, the lessees made no payment to the state but rather were allowed to retain all of the tolls collected and apparently received $100,000 from the state for operating costs. Without speciic provision for maintenance, the lessees were not compelled to keep the waterway in good working order. Not surprisingly, it was allowed to deteriorate. he situation became so bad that the lock and dam at Symmes Creek, formerly called Lock and Dam No. 1, was not even operable by 1877.50 Rescue was, however, on the way. he federal River and Harbor Act of July 5, 1884, included funds for a survey of the Muskingum Waterway. he report was submitted on January 9, 1886.51 Its salient point was a recommendation that the waterway be transferred to the United States. he General Assembly of the State of Ohio acted quickly and on May 14, 1886, agreed to transfer all of the assets of the Muskingum River Improvement to the Unites States, efective July 1. he United States Congress in turn accepted the ofer in the River and Figure 10.8. Upper gate of Lock No. 8, showing original gate opening mechanism on the right. (Huntington District photo archive)

340

T h e M u s k i n g u m N av i g at i o n

Harbor Act of August 5, 1886. he federal government felt that to be viable, it would be necessary for the State of Ohio to turn over all assets: he provisions of this act, as far as they relate to the Miskingum River, shall not take efect, nor shall the money hereby appropriated be available, until the State of Ohio, noting by its duly authorized agent, turns over to the United States all property asked by the act of the general assembly aforesaid, and all personal property belonging to the improvement aforesaid, and used in its care and improvement, and any balance of money appropriated by said state for the improvement of said river, and which is not expended on the iteenth day of July, eighteen hundred and eighty-six.52

he consummation of the transfer was delayed until the General Assembly acted on March 21, 1887. he United States took possession on April 7 of that year. hus, the Muskingum, together with other tributaries of the Ohio, now formed part of a network of river navigations stretching nearly a thousand miles Figure 10.9. Details of a Fontaines drum valve. (Benjamin F. homas and D. A. Watt, he Improvement of Rivers [New York: John Wiley & Sons, 1903], 14)

341

Chapter 10

from the conluence of the Monongahela and Allegheny rivers in Pittsburgh to Cairo where it joined the Mississippi River. In the case of the Muskingum River, all of the dams and a number of the locks, except Lock and Dam No. 1 at Symmes Creek, were rebuilt. he skirting canal at Philo was abandoned and a new Lock No. 9 built at the end of the dam. To provide greater slack-water-pool depths, the crests of ive of the dams were increased as part of the rebuilding program. Even before the United States’ acquisition of the Muskingum River Improvement, the Corps of Engineers had been busy building an “ice harbor” at Marietta. Similar refuges were provided at Parkersburg and Point Pleasant to allow Ohio River steamboats to shelter from the ravages of ice loes on the Ohio River during the winter. In the case of Marietta, the ice harbor involved building a new lock at Marietta at the other end of the dam and abandoning the original Harmon Lock. he new lock, which had an unusual hour-glass shape with three pairs of gates, was begun in 1880 but was still under construction when the United States acquired the Muskingum River Improvement. Ohio River steamboats required more lock chamber space than the locks and dams above Marietta, which were only 36 by 186 feet clear in the chamber and thus unable to accommodate the typical Ohio River steamboat. As a result, the number of steamboats that could shelter in the winter was greatly limited. Lock and Dam No. 1 (No. 11 in the old system) featured the so-called Monongahela-size lock chamber of 55 by 360 feet in the clear. Despite the need, the work dragged on until 1891.53 Symmes Lock and Dam had not functioned for many years, but by the end of the century there was renewed interest in re-establishing the connection between the Muskingum River Navigation and the Ohio and Erie Canal with a view of providing an expanded link between Lake Erie and the Ohio River. he replacement of the Symmes Creek Lock and Dam was a key to re-opening navigation from Dresden to the canal. hus, in a typical approach, the Corps of Engineers undertook a survey of the upper reaches of the river authorized by the River and Harbor Act of August 17, 1894. On the basis of the survey, the estimated cost for a new lock and dam at a location downstream from the original Lock and Dam No. 1 located at Ellis was $110,000. his new lock and dam would increase the navigable depth to six feet. he subsequent River and Harbor Act of 1902 provided the resources for a survey of the river from Zanesville 342

T h e M u s k i n g u m N av i g at i o n

to its headwaters at Coshocton. he upshot of this report was not to recommend improvement up to the conluence of the Walhonding and Tuscarawas rivers at Coshocton, but favored an improvement only as far as Dresden and then through the side cut to join the Ohio and Erie Canal. he next river and harbor act, March 3, 1905, included among other subjects approval of the $110,000 estimate for the lock and dam at Ellis, provided that not less than $200,000 be provided from non-federal sources for improvement of the Ohio and Erie Canal from its junction with the side cut, as far as Cleveland on Lake Erie. he conditions were met and the lock and dam constructed two and a half miles below the old Lock and Dam No. 1. Being located farther downstream and with a higher crest on the dam, there was thought to be danger of looding upstream. As mentioned earlier, the solution was to employ a French-style Boulé movable dam, the only such structure on the waterway.54 Begun in 1907, the work was completed in 1910. While the navigation depth above Zanesville to the link with the Ohio and Erie Canal had provided a four-foot drat, compared with the same depth on the Ohio and Erie Canal, the later Ellis Lock and Dam increased the efective depth to six feet. Earlier, the lood of 1847 caused considerable damage to the crib dams, and in repairing the damage, the slope dams were rebuilt as step dams. Subsequently, all of the dams were converted to this step coniguration. Ater responsibility for the locks and dams was transferred to the Corps of Engineers in 1887, the step dams were reconstructed as slope dams and thus rebuilt in their original coniguration.55 Further, from 1902 to 1922 concrete caps were added to each dam, except for No. 7 and No. 10, where only partial caps were installed. hus, with the completion of repair work in 1922, the waterway was in better condition for navigation than it had been for decades. hese improvements did not, however, attract additional traic. In fact, the gigantic lood of 1913 became a deining point in the history of the entire watershed, with a deep concern by engineers and the public for lood-control measures on the Muskingum watershed as well as elsewhere in the state of Ohio. From 1824 to 1913 the steam packet boat ruled supreme on the Muskingum River, conveying passengers and a wide variety of goods and raw materials. Literally hundreds of steamboats operated on the river and ventured out on the 343

Chapter 10

wider reaches of the Ohio River. Details of these steamboats form an appendix of Gamble’s Steamboats on the Muskingum.56 Following the great lood of 1913, barge and gasoline tows replaced the steam packet boats carrying coal, sand, gravel, petroleum products, and bulk cement. he canal traic was conined to the last three slack-water pools, primarily to supply fuel to the coal-ired electric generating station at Philo. Although trade had diminished over the years, the decade of 1922–1931 shows a considerable tonnage of raw materials transported. he commerce was largely internal, with fewer than 10,000 tons leaving or entering the waterway at Marietta. hus, future projections at the depth of the Great Depression were based largely on the amount of coal, sand, and gravel that would be moved in the future.57 To improve the potential of the slack-water system, a nine-foot-deep channel with high-lit locks and movable crest dams was proposed. his is precisely the solution used on the Great Kanawha River. Beginning in 1933, four highlit dams with German movable roller gates were constructed as part of the public works program of the New Deal. Such a system for the Muskingum attracted little attention from federal New Deal agencies, rather, a comprehensive lood-control system became a model of interagency cooperation with a focus on conservation. he Ohio River nine-foot channel resulted among other locks and dams with the completion of the Bellesville Lock and Dam downstream from Marietta. he slack-water pool on the Ohio River was raised to approximately the level behind the Marietta Lock No. 1. hus, neither the lock nor the low dam were required. In 1955, the W.P. Snyder passed through Lock No. 1 to serve as a permanent exhibition of the steamboat museum. She was the last steamboat to use the lock, which was removed in 1969.58 With diminishing commercial traic, the United States transferred the Muskingum slack-water system back to the state of Ohio on October 16, 1958.59 With this legal action, the waterway once again became part of the state of Ohio, as it had been in the beginning. Under the aegis of the Ohio Department of Natural Resources, a program was initiated to repair the locks and dams for recreation purposes. Completed in 1967, this work permitted

344

T h e M u s k i n g u m N av i g at i o n

the slack-water system from Marietta to Zanesville to be opened for pleasure crat.

Other Slack-water Navigations In an efort to reach the rich coal resources in Pennsylvania and West Virginia, a company was formed to improve the Monongahela River. In 1841 locks and dams No. 1 and No. 2 on the lower reaches of the river, a mile above Pittsburgh’s Smithield Street Bridge, were completed. he new navigation company was authorized by an act of the Pennsylvania Assembly in 1836 to build a slack-water navigation from Pittsburgh to the Virginia line. By 1856, the system was open for navigation to the Virginia line above Point Marion, Pennsylvania.60 As in the case of the Muskingum Navigation, the United States acquired the slack-water system and continued to upgrade it as late as the decades following World War II, the most recent being the Grays Landing Lock and Dam completed in the 1990s. Because of the high sulfur content of the coal, the mining industry in the Monongahela Valley has decreased rapidly following implementation of the Clean Air Act in 1990. hus, the new locks and dams in West Virginia and Pennsylvania have not been fully utilized since they were opened in recent decades. he other river that forms the Ohio River at Pittsburgh, the Allegheny, was skirted by a towpath canal with locks as part of the Pennsylvania Mainline Canal, which opened in 1834. It was not, however, until later that locks and dams were erected on the river to provide a slack-water navigation in the lower reaches of the Allegheny. Sensing rich coal and oil deposits upriver on the Little Kanawha River, a company was formed in 1874 and proceeded to construct four locks and dams that served the Volcano oilield and timber interest upstream from Parkersburg, (Overleaf ) Figure 10.10. he handwritten note says “Where Putnam St. Bridge landed. Flood Mar 1913.” his is the “hourglass” Lock No. 1 at Marietta, Ohio, on the Muskingum River. It is the second lock at this site. (Huntington District Archive)

345

346

347

Chapter 10

West Virginia. With inancial support from the United States government, a ith lock was completed in 1891 and called the Government Lock. his provided slack-water navigation to Creston, West Virginia. Later the Corps of Engineers in 1905 acquired the entire system. Surveys were made with a view to extending the system to Bulltown and the salt deposits associated with that area. his would have required ten additional locks and dams. he extension was never built.61 Moving from the Little to the Great Kanawha, a series of ten locks and dams were begun in 1873 and completed in 1898 with French Chanoine movable dams, which provided open navigation during periods of favorable river stages and, in other times of the year, were raised to produce a series of slack-water pools and locks.62 Without question the ten movable dams that created a slackwater system from the falls of the Kanawha to its conluence with the Ohio at Point Pleasant were one of the most successful projects ever completed by the Corps of Engineers. Opened for traic in 1898, the slack-water system proved an economic success. By 1929, however, it became necessary to undertake major repairs on the irst crib dams that were built, but it was decided to abandon the entire system and replace it with four high-lit dams and twin 55-by-360-foot locks, the so-called Monongahela size. With increased coal traic, a new 800-by-110-foot lock was opened to navigation at Winield in late 1997, with a similar-size lock being constructed at Marmet. By any evaluation, the canalization of the Great Kanawha River is an engineering and economic triumph.63 A much less successful navigation, but nevertheless of considerable interest, is the Big Sandy slack-water system utilizing the French Poirée and Chanoine movable dams. Completed in 1897, the irst needle dam and associated lock at Louisa marked the beginning of the slack-water system constructed by the Corps of Engineers. he needle dam at Louisa became the only such movable dam constructed by the Corps; the other dams utilized the well-tried Chanoine wicket dams. Locks and dams No. 1 and No. 2 on the main channel on the Big Sandy were opened in 1905 followed by Lock and Dam No. 1 on the Tug Fork and Lock and Dam No. 1 on the Louisa Fork in 1909. Although intended as a complete slack-water system, construction ended in 1914 and was never renewed. In fact, the system was abandoned and the movable dams dropped

348

T h e M u s k i n g u m N av i g at i o n

on the crest level to provide open navigation should it be required.64 All traces of the Big Sandy locks and dams have been removed, except for the concrete structure at Louisa, which has been let intact. he tributaries of the upper Ohio River presented both opportunities and challenges for the Corps of Engineers in their attempt to establish a comprehensive river navigation system. Conceived with great enthusiasm, many of these slack-water systems on tributaries of the Ohio River did not stand careful beneit-cost studies, but were nevertheless built. When unable to meet inancial obligations, political solutions were oten obtained to acquire private companies, then transferring ownership to the United States with the responsibility for operation by the Corps of Engineers.

Conclusions While not an extreme case, the Muskingum River Improvement its the pattern with its assets being transferred to the United States. he transfer of the system back to the state of Ohio is an example for the region and one hopes a harbinger of other waterways being transferred for recreational purposes only. Taken together, the Ohio and Erie Canal and the Muskingum River slack-water system represent the irst attempt to tame the Muskingum watershed, not this time for lood control, but to provide a much-needed transportation network as part of the national internal-improvements movement, which was the hallmark of public works in the nineteenth century.

Notes his account of the Muskingum Navigation is part of a larger work on the Muskingum watershed. he study begins with the Slack-water Navigation System and moves to a consideration of the large New Deal Flood Control Project, which resulted in the construction of fourteen lood-control dams. It was a model of inter-governmental cooperation at both the state and federal levels. Recreation and conservation, coupled with important engineering innovations resulted in a project followed by other such lood-control schemes across the nation. he study

349

Chapter 10

of the early slack-water system, and later lood-control dams and reservoirs, provides a record of attempts to control the great forces of nature for the beneit of the people, agriculture, and industry in the Muskingum watershed and beyond. Chapter 10 Notes 1

2

3 4 5 6

7 8 9 10 11 12 13 14 15 16 17 18

U.S. Army Corps of Engineers, Ohio River Division, Ohio River Navigation: Past–Present–Future (Cincinnati, OH: he Division, U.S. Army Corps of Engineers, 1979), 3–4. Michael C. Robinson, History of Navigation in the Ohio River Basin ([Fort Belvoir, VA]: National Waterways Study (NWS-83-5), U.S. Army Resources Support Center, Institute for Water Resources, 1983), 2–3; Leland R. Johnson, Men, Mountains and Rivers: An Illustrated History of the Huntington District, U.S. Army Corps of Engineers (Washington, DC: GPO, 1977), 2–3. Johnson, Men, Mountains and Rivers, 3. Ronald E. Shaw, Canals for a Nation: he Canal Era in the United States 1790– 1860 (Lexington, KY: University of Kentucky Press, 1990), 7–9. Ibid., endnotes 9 and 15. C. C. Huntington and C. P. McClelland, Ohio Canals, heir Construction, Cost, Use and Partial Abandonment (Columbus, OH: Ohio State Archaeological and Historical Society, 1905), 9. Ibid., 10. Shaw, Canals for A Nation, 127–130; Huntington and McClelland, Ohio Canals, 10–15. John Kilbourne, Public Documents Concerning the Ohio Canals (Columbus, OH: I.N. Whiting, 1832), 317–319. Ibid., 353. Ibid., 325–326. Ibid., 356. Huntington and McClelland, Ohio Canals, 31. Ibid., 38–41, 48–49. Ibid., 40. Ibid., 40. Shaw, Canals for a Nation, 131, 134. Huntington and McClelland, Ohio Canals, 47–48.

350

T h e M u s k i n g u m N av i g at i o n

19 20 21

22 23 24 25 26 27 28 29 30 31 32 33 34 35 37 38 39 40 41

42

Ibid., 47–48. Ibid., 48–51. Larry N. Sypolt and Emory Kemp, “he Little Kanawha Navigation,” Canal History and Technology Proceedings, vol. X, 1991 (Easton, PA: Canal History and Technology Press, 1991), 49–93. Johnson, Men, Mountains and Rivers, 100. Emory L. Kemp, he Great Kanawha Navigation (Pittsburgh, PA: University of Pittsburgh Press, 2000), 42–81. Huntington and McClelland, Ohio Canals, 139–141. Ibid., 140. Ibid., 110. Jay Mack Gamble, Steamboats on the Muskingum (Staten Island, NY: Steamboat Historical Society, 1971), 1. Kemp, he Great Kanawha Navigation, 16. Kilbourne, Public Documents, 44. Huntington and McClelland, Ohio Canals, 158–160, 163–165. Kilbourne, Public Documents, 298. Ibid., 317–319. Ohio Board of Canal Commissioners, Fourteenth Annual Report of the Board of Canal Commissioners (Columbus, OH: James Gardiner Printers, 1836), 10. Kilbourne, Public Documents, 447–448. Charles Singer, et al., A History of Technology (London: Oxford University Press, 1957), vol. 3, 440–444. Huntington and McClelland, Ohio Canals, 40. Kilbourne, Public Documents, 317, 359. Ibid., 404. Edward Wegmann, he Design and Construction of Dams (New York: John Wiley & Sons, 1900), 140–145. B. F. homas and D. A. Watt, he Improvement of Rivers (New York: John Wiley & Sons, 1913), part 2, 530, 608–616; Wegmann, he Design and Construction of Dams, 162–164. homas S. Hahn and Emory L. Kemp, he Alexandria Canal: Its History and Preservation (Morgantown, WV: West Virginia University Press, 1993), Monograph Series No. 1, Institute for the History of Technology and Industrial Archaeology,

351

Chapter 10

43

44 45 46 47 48 49 50 51

52 53 54 55 56 57 58 59 60 61 62 63 64 65

45–54. homas and Watt, he Improvement of Rivers, plate 53; Michael S. Raber, Patrick M. Malone and Charles Parrott, Muskingum River Lock and Dam Study (Columbus, OH: Wolpert Consultants, 1991), unpublished revised drat report, 18. homas and Watt, he Improvement of Rivers, 392–408. Ibid., 410. Raber, Malone and Parrott, Muskingum River Lock and Dam Study, 12, 14. Ibid., 19–20. Sypolt and Kemp, “he Little Kanawha Navigation,” 66–68. homas and Watt, he Improvement of Rivers, part II, 493–496. Ibid., 497–498. War Department, Corps of Engineers’ Report on the Muskingum River, Ohio, Covering Navigation, Flood Control, Power Development, and Irrigation (Huntington, WV: United States Engineering Oice, Huntington), 10 Dec 1932, 23; RG 77, Entry 111, File 7249 Bulkies (Preliminary Examinations–Muskingum River), 23, NARA–Philadelphia. U.S. Army Corps of Engineers, Annual Report (Washington: GPO, 1886), Part 2, 1552. Report on the Muskingum River, 23A. Ibid., 24. Ibid., 26. Ibid., 29–31. Gamble, Steamboats on the Muskingum, 130–141. Report on the Muskingum River, 30–31. Gamble, Steamboats on the Muskingum, 114. Ibid., 115–116. Jay Mack Gamble, Monongahela Navigation Co. (Pittsburgh, PA: Bakewell and Marthens, 1873), 617. Sypolt and Kemp, “Little Kanawha Navigation.” Kemp, he Great Kanawha Navigation, 42–82. Ibid., 249–252. Johnson, Men, Mountains and Rivers, 108–113; homas and Watt, Improvement of Rivers, 563–592, 688.

352

11

French Movable Dams on the Great Kanawha River

T

he history of inland navigation can be divided into traditional towpath canals and various attempts to control natural water courses for the beneit of navigation. Although there were earlier successes, towpath canals lourished in the irst half of the nineteenth century in America and elsewhere, but these great engineering works have been relegated to tourist attractions and in some cases even re-watered. he Ohio-Missouri-Mississippi system constitutes one of the world’s great transportation arteries for the transport of bulk cargos. Unlike towpath canals, rivers are used in various and oten conlicting ways. Mill dams, for example, preclude open navigation. Water supply and irrigation also make claims on water resources. he natural ebb and low of rivers results in periods of low lows on the one hand and freshets on the other. As early as medieval times, attempts were made to control rivers with staunches in mill dams as well as the use of wing walls and chutes accompanied by the removal of snags, boulders, and other debris. Staunches operated in a number of conigurations, such as vertical portcullis gates raised on ropes or chains, and drop gates that could be released to allow the passage of vessels. Especially important were needle staunches in which a series of closely spaced boards, called needles, rested against a sill at the bottom of an opening, supported at the top by horizontal beams spanning the opening. To open the gate the needles were removed and the beams swung out of the way. In all of these systems, lashing was usually employed so that an artiicial lood was created, sweeping boats and arks in its wake. Flashing was an annual ritual on the Yonne River, a tributary of the Seine in France. For more than 400 years, logs were rolled into the river at various points, illing the stream from 353

Chapter 11

Figure 11.1. Details of a Poirée needle dam showing the needles resting on a fermette. (B. F. homas and D. A. Watt, he Improvement of Rivers [New York: John Wiley & Sons, 1913], 560)

354

F r e n c h M o va b l e D a m s o n t h e G r e at K a n aw h a R i v e r

bank to bank. At an appropriate time when the low in the river was adequate, the staunches were opened sequentially and the great leet moved down river from Clamecy into the Seine and thence to Paris.

he Modern Era in France he modern era of movable dams begins not in France but in America, and not with logs but with anthracite coal. he Lehigh River presented a similar challenge to that found by the French in moving bulk cargo downstream. he solution devised by Josiah White was the self-acting “bear-trap” dam.1 he design was simplicity itself, with two leaves supporting each other and operated by water, which was let in between the leaves to raise the dam or drained to create a void under the leaves so that the dam would lie lat on the bottom, permitting vessels to pass over the bear-trap dam on a lood tide. In this manner anthracite coal could reach Philadelphia from mines near Mauch Chunk. As early as 1834, C.A. François Poirée (1785–1873),2 an “ingénieur des Ponts et Chaussées,” elaborated upon the needle staunch to provide a large open pass and weir connected at the end with a traditional miter-gate lock. he dam consisted of a series of iron frames pivoted at the bottom so that they could be raised perpendicular to the low of the river. Once in place, these frames, called fermettes, were locked together to form a three-dimensional framework. he needles, which are usually placed by hand, rested on a sill at the bottom of the dam and on a horizontal member spanning the fermettes at the top. Unlike earlier needle staunches, a series of fermettes could provide a wide opening for vessels and arks during periods of adequate river low. A section of the dam, at a higher elevation, was provided with a movable dam on top of the weir to control the level when the dam was raised and the lock in operation.3 In the beginning, the Poirée dams were of limited height, about two meters. With increased height and the resulting increase in the needle cross section, maneuvering became diicult. With larger and heavier needles, it was necessary to use a rail-mounted crane for both raising and lowering the dam. he largest use of the needle principle is the nearly mile-long Bonnet Carré spillway just above New Orleans controlling Mississippi loods by diverting substantial

355

Chapter 11

quantities of water into Lake Pontchartrain. he spillway is operated by a derrick running along the top of the dam above the needles.4 Not only are the needles diicult to maneuver with increasing hydraulic depths, leakage becomes a problem. Several modiications were introduced to cover the gaps between the needles, including planks and straw and other debris.5 he popularity of the Poirée dams spread to Belgium and elsewhere in Europe. Numerous needle dams continue on the Yonne River alone. Apart from the Bonnet Carré, a notable series of needle dams was erected by Benjamin homas of the U.S. Army Corps of Engineers over the Big Sandy River, which forms the border between West Virginia and Kentucky.6 His work was completed in the last decade of the nineteenth century. he only remaining evidence of these needle dams is at Louisa where the ixed portion of the dam, including the pass and the weir together with the lock, chamber, and gates, survives. All of the movable dam iron work has long since been removed. he other three locks and dams of this system have disappeared. Later in the century, in 1874, Auguste Boulé, chef ingénieur (chief engineer) des Ponts et Chaussées, modiied the Poirée dam. Instead of needles, he substituted planks or panels, which were of such a size that they could not be manually maneuvered. hey could, however, be proportioned to resist increasing water pressure with depth.7 he greatest departure from the Poirée movable dams using fermettes was introduced by Caméré at Port Villez, 1876–1880. he basic concept was not unlike a roll-up door used in industrial buildings. Its application was limited to a number of locks and dams on the Seine below Paris.8 As mentioned above, shutters (or wickets) had been used from early times; however, the braced wicket hinged at the bottom and held by a prop can be dated as early as 1778.9 hénard developed the idea in 1831 for a dam at SaintSerrin, featuring a tripping bar for lowering the wicket and a counter-shutter upstream to facilitate raising the main shutter.10 Several dams were erected according to his speciications. In 1843, hénard was authorized to erect a dam across the Seine near Monteneau with shutters seven feet high and ive feet wide. Before he could undertake his greatest work he was retired and his place taken by Jacques Maurice Chanoine (1805–1876).11 It was Chanoine who developed the wicket dam that was so widely applied in America by the U.S. Army Corps 356

F r e n c h M o va b l e D a m s o n t h e G r e at K a n aw h a R i v e r

of Engineers. By moving the pivot point of wickets such as those developed by hénard to a point one-third of the depth of the wicket from the bottom at the center of pressure, the wicket could be raised in a horizontal position with only minimal water pressure acting on it, thus there was no need for the counter shutter featured by hénard. he prop was housed in a quick-release cast-iron shoe called an “heurtoir” (anglicized to heurter) ixed on the apron of the dam. A notable feature of the Chanoine wicket was the self-acting movement of the shutter; when the water exceeded the height of the wicket it turned the wicket, releasing a lood of water. Chanoine’s irst success was in 1857 with the completion of the Conlans Dam across the Seine.12 he Chanoine wicket dam could be operated from a fermette framework like that used on needle dams, or from a maneuver boat. By moving the prop sideways in the heurtoir, the shutter folds lat on the apron of the dam. he Figure 11.2. Needle dam at Louisa, Kentucky, by B. homas. A modiication of the Poirée dam with an 18-foot height. (Edward Wegmann, he Design and Construction of Dams [New York: John Wiley & Sons, 1900], plate 59)

357

Chapter 11

heurtoir, later modiied by another French engineer, Pasqueau, found widespread application in wicket dams in France and America.13 In order to overcome the diiculty in raising a hénard shutter dam, Dominique Girard (1815–1871) introduced the idea of a hydraulic jack of suicient force to overcome the water pressure. He was authorized to erect hydraulic shutters at Île Brûlée Dam at Auxerre on the Yonne River in France in 1869.14 Most unfortunately, he was killed in the Franco-Prussian War and the work was completed by Callon. A turbine waterwheel furnished the power for these jacks. Although the shutters proved successful at Auxerre, the use of hydraulic jacks was apparently limited to one dam until much later, when it found other applications. Objections to the Girard hydraulic wicket were that it featured expensive equipment and called in question the reliability of such a system. hese concerns probably limited its widespread use. In 1846, Desfontaines introduced a totally new system designated a drum dam.15 By means of sluices, a spoke wicket could be rotated with water pressure. Figure 11.3. View of Boulé gate and Caméré roll-up curtains mounted on movable fermettes in the itting-out shop of the dam at Suresnes. (Auguste Boulé, Le Barrage de Suresnes [Paris: Librairie Polytechnique, 1889], ig. 8) 358

F r e n c h M o va b l e D a m s o n t h e G r e at K a n aw h a R i v e r

Desfontaines was chief engineer of the Department of Marne and was responsible for seven drum dams between 1861 and 1867 on the Marne River. Cuvinet made improvements to the drum dam, but it is not known that any of these modiied drum dams were constructed in France.16 On the other hand, Captain H.M. Chittenden, U.S. Army Corps of Engineers, developed his own drum dam and built one at Lock and Dam No. 1 on the Monongahela River above Pittsburgh.17 his drum dam was found to be diicult to operate because of debris clogging the drum space, and thus found limited application. Even a cursory reading of the record and examining extant structures leads to an appreciation of the widespread use of movable dams in France and the many ingenious devices available to permit open navigation on a river in favorable rises in the water level, while at the same time efectively canalizing a river by raising movable dams to create slack-water pools, using traditional gated locks in conjunction with the movable dams.

Movable Dams on the Great Kanawha River Transfer of technology from person to person and between nations is a persistent theme in the history of technology. he development of movable dams in America provides a particularly illuminating case of searching for the most Figure 11.4. hénard shutter dam, a predecessor to the Chanoine wicket dam. (Edward Wegmann, he Design and Construction of Dams [New York: John Wiley & Sons, 1900], 174) 359

Chapter 11

suitable design to be applied to the Ohio River. his wild and capricious river was and remains today a major transportation artery in mid-America. Early attempts to improve navigation were made by removal of snags, boulders, and other debris, combined with blasting of chutes in shoals and constructing wing walls to narrow and deepen the channel. Still the river levels luctuated wildly. he upper limit of summer navigation was a shoal just upstream from the Wheeling Suspension Bridge, efectively cutting of Pittsburgh for weeks at a time in late summer. Without adequate control of river levels, shippers resorted to the ancient lotage method for moving vast leets of coal barges and other cargos when a suitable rise in the river occurred.18 It is little wonder that rivermen wanted both open river navigation and slack-water pools to augment low-water conditions. Some type of movable dam appeared to ofer the only solution. his was pursued under the capable direction of Colonel William E. Merrill (1837–1891). Following service in the Civil War, Merrill spent his career on river improvements from Pittsburgh to the mouth of the Ohio. By all accounts he deserved the sobriquet “Father of the Ohio River Locks and Dams.”19 Merrill promoted the irst lock and dam on the Ohio at Davis Island, a few miles below Pittsburgh. Committed to the concept of the movable dam, Merrill and his colleagues investigated bear-trap dams on the Susquehanna River and considered Josiah White’s devices as modiied by Felix R. Brunot.20 Merrill was the one who pointed out that a bear-trap dam had been built in France during the 1820s and was the forbear of the great development of movable dams that later occurred in France.21 During their investigations, the Army engineers discovered the work of the French. Of all the types of movable dams developed, the Chanoine wicket dam appeared to be the most suitable for use on the Ohio River and its tributaries. Merrill struggled to secure federal funding for the Davis Lock and Dam project as a prototype of a proposed system stretching nearly 1,000 miles to the mouth of the Ohio. Beginning in 1879, the Davis Lock and Dam was begun but not completed until 1885. By 1929, the entire system of ity locks and dams was completed on the Ohio River. At the same time as Merrill was involved in building the Davis Island Lock and Dam, the Corps of Engineers was busy canalizing

360

F r e n c h M o va b l e D a m s o n t h e G r e at K a n aw h a R i v e r

the Great Kanawha River from the Falls of the Kanawha to its conluence with the Ohio at Point Pleasant. he work consisted of ten Chanoine wicket dams and associated locks. Beginning with the River and Harbor Act of 1873, the project was opened for commerce in 1898. As the irst canalized river in the Corps of Engineers’ system, it provided not only a clear example of technology transfer but also new insights into engineering practice in the Army engineers in the latter half of the nineteenth century.22 With the irst appropriation, the engineering design for what was later called Lock and Dam nos. 4 and 5 was begun by the Corps of Engineers under the overall direction of Colonel William P. Craighill and his civilian assistant, Addison Scott. It was this team, with Craighill in Baltimore and Scott serving as resident engineer, that was credited with the great success of the project. Craighill (1833–1909), a West Point graduate, later became Chief of Engineers in 1895. Scott (1843–1927) retired in 1901, shortly ater the project was completed. He was highly respected in the Kanawha Valley not only as a leader in the navigation project but also as a prosperous business man.23 By the 1870s it was clear that major improvements had to be made if the rich resources of the Kanawha Valley, especially coal, were to be transported Figure 11.5. Chanoine wicket on navigation pass of a typical Kanawha dam. (B. F. homas and D. A. Watt, he Improvement of Rivers [New York: John Wiley & Sons, 1913], plate 65)

361

Chapter 11

to markets. At the time, it was equally important that the Great Kanawha be canalized as the missing link in the Central Water Line project. his visionary scheme resuscitated the early James River and Kanawha Canal project, so cherished by the Virginians in the antebellum period but moribund following the Civil War. he grand Central Water Line project envisaged an inland transportation navigation system of canals and rivers, stretching all the way from Norfolk, Virginia, to the foothills of the Rocky Mountains via the James, Kanawha, Ohio, Mississippi, Missouri, and Kansas rivers. he missing link was connecting the James with the Ohio via the Great Kanawha River.24 hus, the proposed complete canalization of the Great Kanawha River, as part of the Central Water Line, included locks and dams at the Falls of the Figure 11.6. Map of proposed Central Water Line. (Edward Lorraine, he Central Water-Line rom the Ohio River to the Virginia Capes [Richmond, VA: Gary, Clemmitt, and Jones, 1869],ig. 9). he legend reads: MAP showing the Great American Central Water Line, which, with the uninished section of Virginia COMPLETED, will form a perfect WATER CONNECTION between the Ocean Harbors of Virginia and the Rivers of the West and Interior.

362

F r e n c h M o va b l e D a m s o n t h e G r e at K a n aw h a R i v e r

Kanawha and a inal lock, later designated No. 12, at the mouth of the river. With the abandonment of the Central Water Line as a federal project, Lock and Dam No. 1 at the falls was not built and, with changes in pool level in the Ohio River, Lock and Dam No. 12 was not needed. he need for river improvement was dramatically documented by Fisk in his report of 1855: he following extract from Mr. Lorraine’s report gives in a few words an idea of the condition of the navigation of the Kanawha at the time of its examination by Mr. Fisk in 1855: Two months in the year are lost rom dry weather; one half month rom the presence of ice; one month, boats can run of 3-foot draught; three and a half months, boats can run of 4 to 6 foot draught; ive months, boats can run of 6-foot draught and upward. An efort was made, subsequent to Mr. Fisk’s report, to improve the navigation by means of “chutes” excavated through the shoals, with occasional slight wing-dams to throw the water into the chutes. his attempt was too limited in its extent, and failed in its object. Mr. Lorraine states that, in October last, when the water was at its lowest stage, there was a depth of only 18 to 20 inches in the chutes below Charleston. It is desired to obtain a useful depth of not less than 6 feet at all seasons.25

he Work Begins he situation had changed very little from 1855 to the 1870s. With the possibility of the Central Water Line and the need to tap the rich resources of the Kanawha Valley, the improvements were begun following passage of the River and Harbor Act of 1873. he irst appropriation was limited and allowed only smallscale river improvements to be undertaken, including such details as rip-rap walls at Cabin Creek and Elk Shoals and rip-rap wing dams at Elk Shoals and Ten Mile Island, as well as removing obstacles in the channel. he work only maintained an obsolete open-river navigational system begun earlier with state resources. 363

Chapter 11

Figure 11.7. Cross section of dams used on the Ohio and Kanawha rivers. he upper view shows a concrete dam supported on piles; the lower view depicts a rock-illed crib dam. (B. F. homas and D. A. Watt, he Improvement of Rivers [New York: John Wiley & Sons, 1913], plate 67)

364

F r e n c h M o va b l e D a m s o n t h e G r e at K a n aw h a R i v e r

With prior investigations of the river dating before the Civil War and continuing as late as 1872 by Craighill, N.H. Hutton, and C.R. Boyd, coupled with knowledge of French movable dams, the Corps of Engineers was ready to prepare contract drawings and speciications in anticipation of an appropriation from Congress.26 Just because a project is authorized is no assurance that funding will follow at suicient levels to sustain the work at an eicient pace. Canalization of the Great Kanawha River was no exception. Government funds, in small amounts, over the next quarter of a century provided resources to complete the entire system of ten locks and movable dams. On August 20, 1875, D.M. and C.P. Dull received a contract to construct a lock at Brownstown (later named Marmet), while Charles McGaferty and Co. were low bidders to construct Lock 4 at Cabin Creek Shore. In every case, the locks at each site were constructed irst so that navigation could be maintained while the navigation pass and weir sections were built across the river channel inside large coferdams. Each lock was constructed on the side of the river with land and river walls deining the channel. Work was begun on locks and dams nos. 4 and 5 rather than 2 or 3 to better serve coal interests, but also to evaluate the new movable-dam technology, since the upper two dams were to be ixed crest because the fall of the river necessitated closely spaced dams. In this case, in order to avoid lowering a pool below a navigational level when one dam was released, it would be necessary to operate locks and dams nos. 2 and 3 in concert. A more reliable method was thought to be to build ixed-crest dams and associated locks, thus open navigation was not provided above Lock No. 3 except in unusually high water. Dams at both sites 4 and 5 were stone-illed cribs.27 By the 1920s, decay of the timber cribs necessitated extensive repairs. Rather than overhaul the dams, it was decided to replace the entire system with four roller-gated locks and dams. he lowest of the set was constructed at Gallipolis on the Ohio River. Because of delays in funding, the irst phase was not begun until 1875, namely the weir-and-pass dam at site 4 as well as the lock and dam at site 3. he contract for Lock No. 4 was delayed, not for lack of funding, but rather (Overleaf ) Figure 11.8. Typical Kanawha lock under construction. (Steamboat photo collection of G.W. Sutphin)

365

366

367

Chapter 11

because the Corps of Engineers had failed to secure title to the land at the lock location. Actual construction was not begun until October 1875. Awarded in October, Lock No. 4 work continued at the quarry until January, when cold Figure 11.9. Kanawha River, West Virginia. (B. F. homas and D. A. Watt, he Improvement of Rivers [New York: John Wiley & Sons, 1913], 691)

368

F r e n c h M o va b l e D a m s o n t h e G r e at K a n aw h a R i v e r

weather forced a shutdown. At the close of the construction season in 1878, both locks and dams nos. 4 and 5 were essentially complete and in operation, while work was well underway at Lock and Dam No. 3. hus, locks and dams nos. 4 and 5 were the irst French-type movable dams in America, since the Davis Island Dam was delayed and not completed until 1885. he two locks and dams were opened for traic in 1880. With ive years of construction completed, the estimated cost per site was $350,000. At that time, June 4, 1880, $1,142,000 had been spent. he engineers, armed with this iscal information, urged that suicient funds be allocated so that one new lock and dam could be undertaken each year. Congress did not fully support this schedule; work was authorized for Lock and Dam No. 2, but with an appropriation in the amount of only $200,000, out of which operating funds for locks and dams nos. 4 and 5 were to be taken. (Overleaf, pp. 370–79) Figure 11.10. Lock and Dam No. 9 during construction, 21 October 1896 showing some of the wickets and part of the service bridge. he wicket in foreground is on the swing (en bascule). (Steamboat photo collection of G. W. Sutphin) (pp. 370–71) Figure 11.11. Lock and Dam No. 9 during construction, 21 October 1896. Work inside coferdam nearing completion. (Steamboat photo collection of G. W. Sutphin) (pp. 372–73) Figure 11.12. Lock and Dam No. 11 during construction, 1 July, 1898. he weir wickets are in place. Maneuvering winch shown on service bridge. Coferdam appears on the let, with the cableway tower serving the stone yard, and concrete batch plant is in the background. (Steamboat photo collection of G. W. Sutphin) (pp. 374–75) Figure 11.13. Raising 76,000-pound gate at Lock and Dam No. 11, 16 July, 1898. Ater being raised to the vertical position the gate was supported on a jack mounted on top and lowered onto the pintle at the bottom. (Steamboat photo collection of G. W. Sutphin) (pp. 376–77) Figure 11.14. Lock and Dam No. 11, 1 July, 1898. Erecting the fermettes with a railmounted winch. he wickets are in the lowered position. (Steamboat photo collection of G. W. Sutphin) (pp. 378–79)

369

370

371

372

373

374

375

376

377

378

379

Chapter 11

he construction of Lock and Dam No. 6 fell behind schedule and the contractor pleaded for a time extension. he work was not inished until 1883; its original completion date had been 4 November 1880. In March 1882, with locks and dams nos. 4 and 5 in operation, the Corps empaneled a board of engineers to examine movable-dam design. In addition to a favorable report on the Chanoine wicket system, the board recommended that the Chanoine heurtoirs be replaced in future construction with Pasqueau heurtoirs, half on Dam No. 5, which was under construction, and throughout Dam No. 6. he heurtoir is a cast-iron device which receives the end of the wicket prop. By moving the prop laterally, it was released and the wicket dropped to a horizontal position on the top of the dam. Chanoine heurtoirs tended to plug up with debris and occasionally not function. he Pasqueau heurtoirs proved to be a superior choice and were later adopted on subsequent wicket dams. hey proved quite successful in service. Locks and dams nos. 4 and 5 were built with 50- by 271-foot chambers to accommodate four 24-foot-wide by 130-foot-long barges. Lock No. 6 and all the remaining locks downstream used the so-called Monongahela lock dimension of 55 by 313 feet to accommodate a new line of barges 26 feet wide. hus, in a most curious action, the engineer board recommended that Lock No. 2 be built to the larger dimensions despite the fact that Locks 4 and 5 were downstream and of smaller dimensions. In a conidential letter to Scott, Craighill explored the idea of making Lock No. 2 two feet wider.28 he idea was dropped at Lock No. 2, thus Lock No. 2 to Lock No. 5 featured smaller locks. Construction of Lock No. 2 began in 1883, while Lock and Dam No. 6 was still under construction. he missing link in the Central Water Line envisaged a canal-and-river navigation system stretching from Richmond, Virginia, to Charleston, West Virginia, with open navigation from Charleston to its conluence with the Ohio River at Point Pleasant. he river, however, sufered exceptionally low water in 1886, providing strong motivation for completing the lower ive locks and dams, since river traic was brought to standstill below Lock No. 6. he lower reaches of the Great Kanawha were the scene of construction of locks and dams nos. 7 through 11, beginning in 1889 for the work on Lock

380

F r e n c h M o va b l e D a m s o n t h e G r e at K a n aw h a R i v e r

and Dam No. 8. he contract for Dam No. 7 was not let until 1891. Despite the Christmas holidays in 1890, Scott brought the designs for the last lock and dam, number 11, with him to a meeting in Baltimore on December 28. Locks and dams 7 and 8 were under construction for nearly four years from 1889 to 1893, when both were opened for traic in October of that year. Delays dragged the project contract for Lock and Dam No. 11, which was not let until 1893, and the whole project was inally completed in 1898. he October 22, 1898, issue of the Charleston Daily Gazette stated: he great value of the improvement, by the government, of the Kanawha river, is splendidly illustrated at this time. Today, 150 barges of coal, containing two and one-fourth millions of bushels, are in the Point Pleasant basin, and about to leave for the Cincinnati, Louisville and other lower river markets. his coal all comes from the Kanawha collieries, and will ind a ready market below. Our chief competitor for this trade is Pittsburg—and this is where one valuable advantage comes in. Our coal can go out when the Ohio river has six feet of water from Point Pleasant to Cincinnati, while Pittsburg must await a good sized rise in the river, a regular coal boat freshet, all the way up to that city. As an illustration, we have, according to the government reports, 251 days in the year, on the average, when we have a stage of six feet of water below Point Pleasant, while Pittsburg has but 155 days in which her coal can go forward to market, giving Kanawha an immense advantage over our up-the-river neighbors. Our last lock was only quite recently completed, and this will be the irst shipment from the lower pool. Ater looking anxiously forward to this event for more than twenty-ive years, and patiently laboring for it in season and out of season during all that time, our coal men naturally feel good over this excellent piece of fortune now realized. And the whole business community are interested in, and rejoice over, this event. Our business men, professional men, manufacturers, and all, have loyally pulled together in behalf of this enterprise.29

hus, this work represented the irst fully canalized river in America. It was highly successful on an economic basis and served until the 1930s, when the

381

Chapter 11

entire system was replaced with four German-designed roller-gated locks and dams. he present Winield Lock and Dam is the busiest in the entire Corps of Engineers system. As a result, a new 800-by-110-foot lock has been completed and opened to river traic at that site. Canalization of the Great Kanawha River represents one of the great successes of the many projects undertaken by the U.S. Army Corps of Engineers and represents an outstanding example of a transfer of technology from French sources and later German-inspired rollergated dams. Chapter 11 Notes French practice is to use surnames only, with no irst names or even initials. he author has tried without success to ind the full name of several of the engineers featured in this paper. 1

2 3

4 5

Norris Hansell, Josiah White Quaker Entrepreneur (Easton, PA: Canal History and Technology Press, 1992), 50–53 and 140; Benjamin F. homas, “Movable Dams,” Transactions of the American Society of Civil Engineers 39 ( June 1898): 437; Edward Wegmann, he Design and Construction of Dams (New York: John Wiley & Sons, 1900) 191; B. F. homas and D. A. Watt, he Improvement of Rivers (New York: John Wiley & Sons, 1913), Part II, 537–538, 651–665, 697. Charie-Marsaines, “Notice Nécrologique sur M. Poirée,” Mémoires et documents No. 20, Annales des Ponts et Chaussées (1876), 263–276. homas and Watt, he Improvement of Rivers, 559–563; Wegmann, Design and Construction of Dams, 151–158; Leveson F. Vernon-Harcourt, “Fixed and Movable Weirs,” Minutes of Proceedings of the Institution of Civil Engineers, Paper No. 1655, 20 January 1880, 30–33; Cambuzat, “Sur les Barrages Mobiles du Système Poirée et du Système Chanoine Qui Fonctionnent Simultanément Pour les Éclusées de l’Yonne,” Mémoires et documents No. 146, Annales des Ponts et Chaussées (1866), 135–138. Emory L. Kemp, “Stemming the Tide: Design and Operation of the Bonnet Carré Spillway,” Essays in Public Works History 17 (December 1970):1–25. Wegmann, Design and Construction of Dams, 154–155; homas and Watt, he Improvement of Rivers, 561–563; homas, “Movable Dams,” Trans. ASCE 39, 487–488.

382

F r e n c h M o va b l e D a m s o n t h e G r e at K a n aw h a R i v e r

6 7 8 9 10 11

12

13 14 15

16 17 18 19 20 21 22 23

homas and Watt, he Improvement of Rivers, 159–162; homas, “Movable Dams” Trans. ASCE 39, 560–587. Auguste Boulé, Le Barrage de Suresnes (Paris: Librairie Polytechnique, 1889), 1–33; Wegmann, Design and Construction of Dams, 162–164. Wegmann, Design and Construction of Dams, 616–622; Vernon-Harcourt, “Fixed and Movable Weirs,” 30–33. homas, “Movable Dams” Trans. ASCE 39, 492. Vernon-Harcourt, “Fixed and Movable Weirs,” 30–33; Wegmann, Design and Construction of Dams, 172–174. Vernon-Harcourt, “Fixed and Movable Weirs,” 34–36; Wegmann, Design and Construction of Dams, 174–181; homas and Watt, he Improvement of Rivers, 640–645. Vernon-Harcourt, “Fixed and Movable Weirs,” 34–36; Wegmann, Design and Construction of Dams, 174–181; homas and Watt, he Improvement of Rivers, 640–645. Wegmann, Design and Construction of Dams, 177; homas and Watt, he Improvement of Rivers, 574, 586, 589, 595. Remise, “Barrage Mobile Automoteur,” Mémoires et documents No. 52 Annales des Ponts et Chaussées (1873), 360–377. Vernon-Harcourt, “Fixed and Movable Weirs,” 36–38; homas and Watt, he Improvement of Rivers, 646–648; Wegmann, Design and Construction of Dams, 185–187. Wegmann, Design and Construction of Dams, 187–188. Wegmann, Design and Construction of Dams, 188–190; homas, “Movable Dams,” Trans ASCE 39, 553–559. Leland R. Johnson, he Davis Island Lock and Dam 1870–1922 (Pittsburgh: U.S. Army Engineer District, 1985). Leland R. Johnson, Men, Mountains and Rivers (Washington: U.S. Government Printing Oice, 1977), 74–77. Johnson, Davis Island, 22–27; Wegmann, Design and Construction of Dams, 195. Johnson, Davis Island, 35. Emory L. Kemp, he Great Kanawha Navigation (Morgantown, WV: Institute for the History of Technology and Industrial Archaeology, 1997), 1–81. Ibid., 33–38.

383

24

25 26 27 28

29

Secretary of War, “Examination and Survey of the Great Kanawha River, From the Great Falls to the Mouth” (Appendix T 28), 836. Note this survey was based on earlier reports by Edward Lorraine. An Appeal for the Speedy Completion of the Water Line of Virginia, and hrough hat of the Great Central Water Line of the Union (Norfolk, VA: C. H. Wynne, 1857). Kemp, he Great Kanawha Navigation, 4–7, 9–22. Ibid., 836–842. Wegmann, Design and Construction of Dams, 137–150 and Plate 82. Craighill to Scott, 4 Aug 1884; RG 77, Entry 1331, Box 1, Letter 77 (Misc. Oficers’ Letters, Kanawha River), NARA, Philadelphia, PA; Craighill to Scott, 4 Nov 1884; RG 77, Entry 1331, Box 1, Letter 123 (Misc. Oicers’ Letters, Kanawha River), NARA, Philadelphia, PA. “River Improvement,” he Charleston Daily Gazette (Charleston, WV: Daily Gazette, 1891), 1–9.

12

he Little Kanawha Navigation Larry Sypolt and Emory Kemp

I

n October of 1770, George Washington began a trip to the Ohio Valley to secure land claims along the Ohio and its tributaries as a result of land grants authorized to veterans of the French and Indian War. His party let Mount Vernon on the Potomac on October 6 and by the 13th had reached Captain Crawford’s house, beyond Great Meadows along the Youghiogheny where the village of New Haven now stands. Crawford was a boyhood friend of Washington and was his partner in land acquisitions in western Pennsylvania and along the Ohio River. In his book George Washington and the West,1 Charles Ambler stated, “Temporarily lost in eforts to become successful on lands poorly favored by nature, Washington could not have been wholly indiferent to the results of the war between the French and the English. Permeated with traditional Washington land hunger, he still saw in the west, a land of opportunity.” As an example, Crawford received a letter from Washington in September, 1767, directing him to “look me out a Tract of about 1,500, to 2,000 or more acres.”2 On October 16, Washington reached John Stephenson’s house and there learned about a report by a Mister Ennis who had traveled down the Little Kenhawa (Kanawha) from near the headwaters to the mouth of the river where later Parkersburg was to be founded. He reported that the lands were broken with the river bottoms being neither very wide nor very rich, but covered with large stands of “Beach trees.” he following day, the party reached Fort Pitt, and on the 18th dined with Colonel George Croghan, a famous Indian trader, land speculator, and British agent for Indian afairs. Although not mentioned in his diary, Washington had quarreled with Croghan in 1754 over horses Washington had commandeered. Not only did Croghan invite them to dinner at his home, but on the 20th when they embarked in a large canoe for their 385

Chapter 12

party and a separate canoe for Indians, Croghan accompanied them during their irst day’s journey. On the 27th, the two canoes reached the mouth of the Little Kanawha. In his diary, Washington states: his is about as wide as at the mouth of the Muskingum. But the water is much deeper. It runs upwards toward the inhabitants of Monongalia, and according to the Indian’s accounts, forks about 40 or 50 miles up it and the ridge between the two prongs leads directly to the settlement.3

he party continued down the Ohio, eventually reaching the mouth of the great Kanawha River on November 3. Ater their brief excursion up this river they returned back up the Ohio and past the Little Kanawha between November 7 and 8. George Washington was not the irst to explore the tributaries of the Ohio, but his role in the internal-improvements movement makes this visit symbolic of what was to be a national trend. Earlier, in the winter of 1751, Christopher Gist, as a representative of the Ohio Company, explored the lower reaches of the Little Kanawha River. In fact, in February, 1751, “1752” by the new calendar, he had reached the vicinity of Elizabeth where on a great standing stone he carved in large letters THE OHIO COMPANY FEBY 1751 BY CHRISTOPHER GIST. Later, in 1765, George Croghan traveled down the Ohio and let a brief comment about the mouth of the Little Kanawha River. He was later stationed at Fort Pitt. Following the American Revolution, the valley of the Ohio was rapidly settled along with many of its tributaries. he Little Kanawha Valley, unlike most of the western lands, was settled from west to east. here was suicient settlement along the Little Kanawha in 1790 for the local inhabitants to forward the following petition to the Harrison County Court: PETITION FOR A ROAD By the Inhabitants of Neal’s Station of the Little Kanawha in the year 1790 To the Worshipful Court of Harrison County: he petition of the inhabitants of Neal’s Station, on the Little Kenaway, humbly showeth, that your petitioners, as well as the settlers on the west of the Ohio,

386

T h e L i t t l e K a n aw h a N av i g at i o n

and travelers from Caintucky, labour under great diiculty for want of a road from said station into the State road, as also southward to the Kenaway County line, as many of the travelers from Caintucky leave their canoes at Belveal, and come across by land to Clarksburg, and oten bewildered in the woods, or obliged to hire a pilot to bring them through. Your petitioners, therefore, humbly pray your worships, to grant them an order for laying of and opening a road into the above mentioned State road, that leads to Clarksburg, and your petitioners, as in duty bound, shall ever pray, etc. West Virginia, At a County Court held for the County of Harrison, on the 20th day of September, 1790, ORDERED, hat Michael homas, Jeremiah Sergeant, James Neal and Moses Hewitt, or any three of them, they being irst sworn, do view and mark a way for a road from the State road, by Neal’s Station, on the Little Kenaway, and from there to the Harrison and Kenaway County lines, and report the conveniences and inconveniences to Court.4

Before the advent of suitable roads and later railways, the little Kanawha provided the principal artery of commerce through the central part of western Virginia, but its development as a navigation can be viewed in a larger context as part of a national movement for internal improvements.

American Internal-Improvements Movement he internal-improvements movement in America predates the founding of the country, but gained considerable momentum ater the American Revolution as a means of developing the vast resources of the new republic by a network of roads and canals and later railways. As Gallatin stated in his report on internal improvements published in 1808,5 the internal improvement movement would also serve a nobler purpose: to foster a more perfect union of the former colonies into a new nation. From colonial times, leaders in Virginia were interested in developing a transportation network. George Washington and other prominent igures proposed the improvement of the James and Potomac rivers to penetrate beyond the mountains to the Ohio Valley. In fact, Washington can be credited with being the founder of both the Potomac Company, which later resulted in the 387

Chapter 12

Chesapeake and Ohio Canal and the James River and Kanawha Canal and turnpike. Washington discussed a full range of internal improvements concerned with rivers and canals, in which he said: to do this, the waters which empty into the Ohio on the east side and which communicate nearest and best with those which run into the Atlantic must also be delineated—hese are Monongahela and its branches viz Yohiogany and Cheat and the little and great Kanawhas and Greenbrier which enter into the latter. 6

In a letter to homas Jeferson dated January 1, 1788, Washington speaks of opening a communication between the James River and the Great Kenawha or between the Little Kenawha and the west branch of the Monongahela which is said to be very practical by a short portage. As proof of this a road is now opened or opening under the authority and at the expense of the states of Virginia and Maryland from the north branch of the Potomac communicating at the mouth of Savage River to Cheat River and continued from thence to the navigable waters of the Little Kenawha at the cost of the former. 7

Following the reorganization of the Potomac Company and the founding of the James River and Kanawha Canal, work was started on the two projects with the intention to reach the Ohio River. But, in fact, neither of them was completed to form the great arteries of transportation envisaged by their proponents, especially Washington. Nevertheless, both served local and regional purposes, particularly the C & O Canal which became a major hauler of bulk cargo, especially coal. In 1816, Virginia established a Board of Public Works under the leadership of Claudius Crozet, who had served as an engineer oicer in Napoleon’s army during the French wars in Europe. he Board of Public Works fostered internal improvements in the Old Dominion through a series of companies established to build turnpikes and canals on a system of mixed capital in which funds would be raised locally and matched by funds from the Board of Public Works. hus began the establishment of a number of transportation companies throughout 388

T h e L i t t l e K a n aw h a N av i g at i o n

Virginia. Although many were ill conceived and ill funded, a number were constructed that served the state well in the nineteenth century. Despite Gallatin’s initiative as Secretary of the Treasury, the federal government was to play a minor role in the development of a national road and canal network. It was let to the states, municipalities, and private enterprise to build the system. herefore, it is not surprising that the roads, canals, and railways were built not in a cooperative manner, but in a iercely competitive spirit serving local and regional needs with little concern for the development of a rational, national network. If there was a single igure to epitomize the internal improvements movement in Virginia it would be Claudius Crozet. Under his leadership, many turnpikes were built under the aegis of the Board of Public Works. In addition to the Little Kanawha, two of the foremost east-west turnpikes ended in Parkersburg. hey were the Northwestern Turnpike beginning in Winchester, and the Staunton to Parkersburg Turnpike. his was later followed, just before the Civil War, with the completion of the Northwestern Virginia Railroad, a wholly owned subsidiary of the Baltimore and Ohio Railroad connecting with the main stem of the B & O at Graton and following roughly the Northwestern Turnpike to Parkersburg, and thence to Cincinnati and St. Louis. Parkersburg, being located on the navigable Ohio and the terminus of these internal improvements, was second only to Wheeling as a transportation center in western Virginia.

Early Surveys and Legislation Regarding the Little Kanawha River he General Assembly of Virginia took note of the Little Kanawha River very early in the nineteenth century. An act passed on January 12, 1803, authorized the construction of mill dams across the Little Kanawha, provided they did not obstruct navigation, as the state of Virginia considered the Little Kanawha River a public highway from its conluence with the Ohio River at Parkersburg to Bulltown. However, by the 1830s those interested in navigation were at loggerheads with mill owners who had erected dams and various other obstructions to navigation in the Little Kanawha. 389

Chapter 12

Chapter 153, Acts of the Virginia Assembly, passed on March 11, 1834, required mill dams to have locks the full length of the river for navigation.8 he locks were to be a minimum of 20 feet wide and 100 feet long. In lieu of locks, a slope of not less than 20 feet wide and 40 feet long had to be constructed at one end of the dam for the passage of ish and boats. Many of these slopes apparently were controlled by gates at the top, which could be opened or dropped to permit the passage of a boat. hese constituted a primitive type of lash lock. his legislative act requiring the provision of slopes and control gates for mill dams was further modiied in the 1834–1835 acts of Figure 12.1. Map of the Little Kanawha Valley showing the relationship of the oil ields and the Little Kanawha River to Parkersburg, West Virginia. (he Sunday Morning News, Parkersburg, West Virginia, 20 June, 1897). 390

T h e L i t t l e K a n aw h a N av i g at i o n

the General Assembly “with regard to the dam on the river where the state road crosses it,” stipulating that a slope not exceeding three degrees and not less than thirty feet wide be erected. Five disinterested commissioners were appointed by the county courts to provide a fair tarif for use of the locks at each dam. his compensated the owners for expenses incurred in the construction of the locks and their maintenance. On February 20, 1838, the General Assembly passed Chapter 119, directing a survey of the Little Kanawha River from its mouth to the Bulltown salt works. his act requested the survey to include a cost for erecting locks the full length of the river and other improvements deemed more advisable. Crozet undertook this survey and recommended that ten locks and dams be constructed at a cost of $480,000. his would allow the river to be navigable its full length and permit river traic to the Bulltown salt works. hus begins the irst attempt to establish a navigation on the Little Kanawha River. he river that Crozet surveyed rises on the western slope of Laurel Ridge, which is one of the principal ridges of the Appalachian mountain system. It lows across Braxton, Gilmer, Calhoun, Wirt, and Wood counties. From its source at Kanawha Head in Upshur County, the river lows 158 miles in a rather tortuous path to reach the Ohio at Parkersburg. Its course is depicted in an early map, which also shows great natural resources that an improvement in river navigation was expected to exploit (Fig. 12.1). hese include the Great Coking Coal Basin on the upper reaches of the river and the abundant sources of oil and natural gas throughout the valley. he original survey by Crozet, which was quite detailed with regard to distances, elevations, and characteristics of the river and its valley, was supplemented in 1874 by engineer oicer E.J. Carpenter under the authority of W.E. Merrill, a wellknown civil and military engineer. he General Assembly again took up the matter of the Little Kanawha River in 1845. It allowed for the courts of each county along the river “to appoint three discrete and disinterested persons acquainted with the currents of the river to select the most suitable place in each dam for a lock or slope.” Further emphasis was put on this act with regard to erecting dams, their maintenance, and compliance with laws by adding:

391

Chapter 12

it shall be the duty of judges of superior courts of counties traversed by the Little Kanawha to give this act, and all others relating to the navigation of the Little Kanawha, in special charge to the grand juries. And, it shall be ex oicio the duty of the prosecuting attorneys of such courts to proceed by information against any person ofending against their provision.

he act further provided that when a lock instead of a slope is constructed in any dam, it shall be lawful for the proprietor . . . to demand a toll of one dollar per ton on boats and lading passing through such lock; and so in proportion for a greater or less quantity than a ton.

Books of subscription for the sale of stock in the Little Kanawha Navigation Company were opened in Parkersburg, Virginia, in 1847. One thousand shares of ity dollars each were authorized under the supervision of J. Dickerson, Alfred Beauchamp, Hiram Pribble, Peyton Butcher, Willis Leach, Abraham Enoch, Peter G. Van Winkle, and Jeferson Gibbons. he company had a goal of opening, for navigation, the Little Kanawha from its mouth to Bulltown. he company was organized according to an act entitled “an act prescribing certain general regulations for the incorporation of turnpike companies.” Purchasing powers were authorized to the oicers for materials necessary for the construction of dams and toll houses and the company was authorized to collect tolls for each ten miles of completed work. In 1848, the Board of Public Works was instructed “to purchase, on behalf of the Commonwealth of Virginia, three-iths or thirty thousand dollars worth of stock of the Little Kanawha Navigation Company, ater the board is satisied that three-fourths of the remaining two-iths of the capital stock shall have been subscribed to by solvent persons.” his is the same procedure used for inancing scores of turnpike companies. he work was to be divided into three sections. he irst was from Parkersburg to Elizabeth; the second from Elizabeth to Gilmer Court House; and the third from Gilmer Court House to Bulltown (see Fig. 12.1). Money was to be appropriated in thirds, with no more than one third going to each section. Subscriptions for the second and third sections would not be made until the irst was completed. 392

T h e L i t t l e K a n aw h a N av i g at i o n

Little had been accomplished since 1848, so in 1853 the charter was revived with new supervisors of subscription. hey included John R. Murdock, D. R. Neal, James Cook, J. M. Jackson, and R. W. Dickinson of Parkersburg; Alfred Beauchamp, W. P. Rathbone, and J. A. Williamson at Elizabeth; and Robert Erwin, P. Hays, and Minter Jackson at Glenville. he total capital would consist of 1,000 shares of stock at 50 dollars each. “he Little Kanawha would be improved from its mouth to the falls of said river at Haymond’s mill in Braxton County.” Again in 1860, ater very little progress, the charter was revived, amended and re-enacted under the superintendence of Daniel R. Neal, James Cook, and A. I. Boreman at Parkersburg; Daniel Wilkeson, Abraham Enochs, S. A. Ruble, and Fidellus Ott at Elizabeth; Collins Bells, Hiram Firrell, and George W. Hardman at Big Bend in Calhoun County; and Robert Linn, C. B. Conrad, Levi Johnson, and S. G. Stalnaker at Glenville. he capital was decreased to $25,000, consisting of 1,000 shares at $25.00 each. he Little Kanawha River was to be opened for navigation from its mouth at Parkersburg to a point where the Gilmer and Braxton county line crosses the river. he Virginia Assembly amended the act of 1860 in 1861. It extended the subscription area to Bulltown under the superintendence of C. S. Harlay, Addison McLaughlin, William P. Haymond, and Moses Cunningham. he Moses Cunningham farm has recently been restored to its antebellum condition and is part of the Burnsville Historic Park. It is located on the Weston & Gauley Bridge Turnpike at the site of the Bulltown covered bridge across the Little Kanawha. his turnpike connected the Staunton to Parkersburg Turnpike with the James River and Kanawha Turnpike and was the only north-south turnpike in western Virginia. A Civil War skirmish was fought on the turnpike near the farm on October 13, 1863.

Natural Resources Coal and timber were thought to be in great enough quantity to justify the improvement of the Little Kanawha River for navigation. he Great Coking Coal Basin was in the area around Glenville and Burnsville. Timber had been cut and loated down the river using the slopes in mill dams for passage since 393

Chapter 12

the beginning of the century. hese two natural resources, although exploited in notable quantities, were not suicient justiication for making the river navigable its whole length. he discovery of natural gas and, most important, oil was to spark an excitement in the business community that demanded the river be improved for navigation. he Volcano Anticline, the most pronounced anticline in the whole Appalachian region, collected oil under an impervious stratum. Oil and gas exuded from the ground at Burning Springs in Wirt County. All of the oil-producing strata in the Volcano Anticline occur at a fairly shallow depth. he Ritchie and Wood counties oil ield also came into existence within months of Drake’s well in Pennsylvania, and just a year before the Civil War. A heavy, dark green, 29-degree-gravity lubricating oil was discovered on the headwaters one mile south of Volcano. he town was given the name Oil Springs, and the run on which it was discovered named Oil Springs Run. hree miles south of this was the town of Petroleum. William P. Rathbone of New York and New Jersey began buying land at Burning Springs in 1840 during a visit to his daughter, who lived in Parkersburg. He was speculating in the salt business. He and his family moved there the following year. Later, two of his sons, John Valleau and John Castello (Cass) were to go into the oil business with him. hey bought up land along the Little Kanawha River and Burning Springs Run. By 1848, the Rathbones and Peter G. Van Winkle, who married W. P. Rathbone’s daughter, owned more than 21,000 acres in Wirt County. he Rathbones began a medicine business soon ater their arrival. A small pottery in Parkersburg produced a quart-size jug which the Rathbones illed with oil from the river and marketed as Rathbones’ Rock Oil–Natures Wonder Cure. It had a wide distribution and sold for ity cents locally and for one dollar elsewhere. he Rathbones operated other businesses in conjunction with their medicine enterprise. hese included a general store, a steam-powered sawmill, a steam-powered grist mill, and logging operations. he Titusville, Pennsylvania, oil strike in 1859 caused speculation for oil at Burning Springs. General Sherman P. Karns leased an abandoned salt-brine well from Cass Rathbone. It had been abandoned because the salt was too saturated with oil to be of any commercial value. Using a steam-powered pump, he pumped 394

T h e L i t t l e K a n aw h a N av i g at i o n

for a week before the well began producing oil at a rate of seven barrels a day. Johnson Newlon Camden, the Rathbones, and a few other associates from Parkersburg now began drilling for oil on lands owned by William Rathbone. Ater drilling to a depth of 303 feet, General Karns struck oil in 1860. he oil sold for ity cents a gallon. During that same year, other wells were drilled with great excitement at Burning Springs. J. N. Camden was president of the bank in Parkersburg; he had been raised in Lewis, Harrison, and Braxton counties and held various jobs in law, county courts, and investments. Parkersburg was one of the fastest-growing cities in (West) Virginia; it was second only to Wheeling as a commercial and industrial center. he stockholders in the Little Kanawha Navigation Company, largely from Parkersburg, were to become political leaders in the new state. When West Virginia became a state in 1863 her irst governor, A. I. Boreman, and one of her irst senators, Peter G. Van Winkle, resided in Parkersburg and invested in many enterprises including the Little Kanawha Navigation Company. By the time the Civil War started, several hundred wells had been drilled. Many wells were now abandoned as their owners went to ight for their beliefs. However, men such as J. N. Camden, the Rathbones, William Paden, the McConougheys, the McFarlands, D. A. Roberts, and Lewis Wetzel remained. hey agreed to continue oil operations with existing wells, but not to open new ones until the end of hostilities. On May 9, 1863, Confederate General William E. Jones arrived at Burning Springs with 1,500 troops. hey set ire to every oil house, derrick, and tank. At that time, over 120,000 barrels of oil were stored there or in barges on the Little Kanawha River, waiting for the water to rise for the trip to Parkersburg. Each barge had a hole cut into its deck and the oil ignited. he ire could be seen glowing all the way to Parkersburg. Trees along the river up to Palestine were killed by the ire and heat. he soldiers were obviously sent to destroy the oil to keep the federal government from collecting a one-dollar-a-barrel tax to help inance the war. It was the irst time a major strategic industrial target was destroyed for military causes. Total destruction was the result of the Confederate raid. For two years it was feared that the oil industry would die. Derricks were ruined. Storage tanks were of no use. Houses were torn down and their materials used for irewood. he land was bought up cheap. 395

Chapter 12

During the Civil War, West Virginia became a separate state. he Restored Government of Virginia and later the legislature of West Virginia gave priority to the charter of the Little Kanawha Navigation Company. he Restored Government of Virginia, sitting in Wheeling, passed a special act incorporating the Little Kanawha Navigation Company on February 4, 1863. he company was authorized to issue $100,000 of capital stock for the purpose of improving the navigation. he incorporators were J. N. Camden, John V. Rathbone, Peter G. Van Winkle, James Cook, Moses Kinchelo, Daniel Wilkinson, E. G. Hopkins, Jonathan Weaver, Charles Chadock, John Weir, and James Wilkinson. J. J. Jackson, a cousin of homas J. “Stonewall” Jackson, served as the irst president of the reorganized company. An oicial conirmation of stock held is shown in Fig. 12.2. Gideon D. Camden was a judge in Harrison County and a cousin of John Newlon Camden. he West Virginia legislature amended the charter of the Little Kanawha Navigation Company on March 1, 1864. he same subscriptions’ commissioners were named. he company now had the power to improve the tributaries of the Little Kanawha River, as well as the main channel. he act also stated allowable rates to be charged on transporting oil and other products on the improved sections of the river. he end of the Civil War brought renewed interest in Burning Springs and the oil business. he Rathbones set up two new saw mills and soon hundreds of new homes and many businesses existed. A new cooperage was constructed and a dozen coopers were hired to make storage tanks and barrels for oil. By the end of 1865, the town had almost reached its prewar population. Renewed interest in the Little Kanawha Navigation Company was taking place. his time the incorporators were determined to make the river navigable so oil and other natural materials could be transported to Parkersburg for transshipment to other markets. An area of contention that hampered navigation was the Kanawha Bridge in Parkersburg. his was a covered bridge that crossed the Little Kanawha River on the south side of Parkersburg. he bridge did not permit boats with high smokestacks to pass under it. he public demanded that it be removed and a draw bridge put in its place to permit steamboats to travel upriver. Others contended that the bridge existed in violation of law by preventing navigation to the great mineral wealth along the river. Eventually the 396

T h e L i t t l e K a n aw h a N av i g at i o n

problem was solved in a dramatic way, when the Ohio River looded and the bridge was washed away. he Parkersburg Daily Times ran many articles promoting the improvement of the river because it was navigable for only a few months a year, when the water was high. he people of Parkersburg saw that they could beneit greatly from the increased river traic that would have to pass through their city. Oil, lumber, coal, and iron would have to be transferred to larger vessels before being shipped to other markets. hey urged the citizens of Parkersburg to purchase stock in the navigation company for improvement of the river. June 1866 saw the commencement of removing mill dams from the mouth of the Little Kanawha River to Burning Springs. Captain Coles had the contract Figure 12.2. Advertisement for the steamer Oneida, part of the Parkersburg and Creston Packet company leet, showing departure times and round-trip fares for all points along the Little Kanawha River. (Daily Morning News, Parkersburg, West Virginia, 25 May, 1900)

397

Chapter 12

for removing them from the mouth at Parkersburg to Slabtown, and Captain J. Marshall from Slabtown to Burning Springs. By July 1866, most of the mill dams had been removed. his meant that boats could come downriver on a ive-foot rise that would have required a nine-foot rise while the mill dams were in place. here were an estimated 70,000 barrels of oil sitting at Burning Springs and a potential production of 2,300 barrels per day. At a ten-cents-per-barrel freight rate, the Little Kanawha Navigation Company stood to make a great proit. he major problem was that the oil could be brought downriver only during high water. A slack-water navigation created by locks and dams was still needed for the eicient movement of goods and people. An Oil Convention of West Virginia was announced for August 8, 1866, at the American Hotel in Philadelphia. he major portion of business was concerned with the construction of locks and dams for slack-water navigation. Oil had to be taken to market, and transportation by team and wagon was costly and possible only in good weather. A solution had to be found and found soon, thus each company was requested to send a representative to the convention. he oil interests at Burning Springs were at stake. he next several months produced much discussion on plans for locks and dams. Types of material for the locks and dams and chamber sizes for locks were discussed. Meanwhile, in 1867, the West Virginia Transportation Company came into existence for the express purpose of transporting oil through a pipeline.9 he company had already constructed three miles of piping connecting Oil Rock with the Little Kanawha River at the mouth of Standing Stone Creek. he Rathbone Oil Tract Company refused to let the West Virginia Transportation Company cross its property to the river, even though they were willing to link up with the Little Kanawha Navigation Company no matter how far upriver they provided locks and dams. On February 14, 1867, Mr. McArthur, civil engineer of the Little Kanawha Navigation Company, announced that the irst lock and dam would be at Shacktown and the contractors were Messrs. Dawson and Maurice. He also announced that due to the high cost of timber, the dams would be constructed of stone. He further stated that bids would be accepted for the second lock and dam below Newark at Leeche’s Mill. 398

T h e L i t t l e K a n aw h a N av i g at i o n

he Elizabeth Gazette announced in March 1867 that the town’s citizens had subscribed $5,000 in stock to the navigation company. Mr. S.P. Wells, superintendent of the Rathbone Oil Tract Company, had promised the people of Elizabeth that the irst lock and dam would be built there if they subscribed $5,000. he money was raised within a half-hour of the adjournment of the meeting. Work started on the lock at Shacktown in May 1867. Mr. Maurice found a quarry of suitable sandstone near the site of the irst lock and dam, and promised to push the workers hard and make good progress. Meanwhile, river navigation continued. Timber was loated down the river and caught in booms. he counties and communities grew restless and wanted more say in the policies of the navigation company. he Little Kanawha Navigation Company announced on March 20, 1869, that General T.J. Power of Rochester, Pennsylvania, was awarded the contract on the two upper locks. hey were to be located at Elizabeth and Palestine. He was given one year to complete both. he four locks and dams were to be built on the same general pattern. he lock was to be 23 feet wide and 125 feet long, and built of stone masonry with hand-operated mitre gates. It is not clear why the unusual lock size was selected; it may have been because of the dimensions of lat boats or rats using the river. he dams were of a ixed type of stone-illed timber-crib construction. his was the cheapest kind of construction. he dams ranged from 274 to 289 feet in length. In May 1869 S.W. Jordan was contracted to inish the work on Lock and Dam No. 1, which had been started in 1867. He had to complete the work by the time the remaining locks and dams were inished. he oicers of the company expected all work to be completed in 1869. he locks and dams were not completed until 1874. Lock No. 1 was located at Shacktown, 3.5 miles from Parkersburg; it had a 15.7-foot lit. Lock No. 2 was located at Leachtown, 14 miles upriver, and had a lit of 10.1 feet. Lock No. 3 was located 22 miles upriver, just below Elizabeth, and had a lit of 11.8 feet. Lock No. 4 was located 32 miles upriver from Parkersburg, at Palestine, and had a lit of 12 feet. Seven years had passed since the work had been started. Much of the need for oil transportation had passed. he navigation company had missed its opportunity to promote the Little Kanawha Valley and make a good return 399

Chapter 12

on its investment. he lower portions of the river, however, now provided slackwater navigation and a new era was about to open. he locks were carefully located to provide the correct elevations for the slack-water pools and ensure that the structure could be founded on a secure rock bottom. he lower dams were rock-illed cribs extending from the lock across the river. he faces were spackled and grouted to cribs to provide a smooth low of water over the dam. he “Government Lock,” constructed in 1891, was a crib dam built with a reinforced concrete apron, apparently using natural cement. Only the sill timbers of the crib dam remain at Lock and Dam No. 4. he dams on the lower portion of the navigation are not visible for examination. he irst locks were constructed of stone resting directly on the rock bottom of the river, whereas Lock No. 5 rests on wood piles. All except Lock No. 1 have survived intact. he river wall of the irst lock was removed when the pool level on the Ohio River was raised as a result of the construction of new large locks and improved dams on the Ohio River in 1969. Ater the entire navigation was purchased by the government in 1905 and operated by the U.S. Army Corps of Engineers, the locks and dams were repaired. In the case of the locks, concrete was used to repair the masonry walls and raise the height of the locks. Apparently, new lock gates were installed in all ive locks.

River Commerce Before the development of the Little Kanawha Navigation, which would permit the use of steam boats, the river was frequently used for loating logs and, later, barrels of oil, and occasional latboats with agricultural produce destined for Parkersburg. here was very little commercial traic upstream. Later, logs were rolled or hauled to the river, assembled into large rats and loated to Parkersburg to be sold. Still later, logs were loated downriver and caught in booms. 10 Ater the locks and dams were completed, rats were assembled in lengths and widths that would permit them to go through the completed locks. hey were manned by two or three men and the trip took two or three days. his type of traic ceased ater 1915, when gasoline-powered boats began taking lumber to Glenville for building purposes. 400

T h e L i t t l e K a n aw h a N av i g at i o n

Flatboats were used to haul goods to people who lived along the river. A latboat was known by several names such as pushboat, because it was pushed by poles; tow boat, because it was sometimes towed by horses on the bank; bateau, because it was light and lat-bottomed; and barge, because of its style. In addition to household goods and supplies, these boats carried coal and other marketable items to Parkersburg. he steam and gasoline boats replaced this type of boat by the mid-1800s. he gasoline boat lasted from 1898 until commercial traic stopped in the 1940s. (One informant states that traic lasted until 1950.)11 hey represented the most numerous type of vessel plying the river. he steamboats were the irst to make major use of the river ater completion of the locks and dams in 1874. Until then, only shallow-draught boats made trips up the Little Kanawha, and then only when the water was high. he Scioto Belle was reported to be the irst steamboat to go up the Little Kanawha to Creston in 1842; some say she was the irst steamboat on the river. he Lodi was the second to go up in 1847. Another source states that the Paul Pry was the irst steamer to ascend above one mile on the Little Kanawha River in 1837. It ascended to Leeche’s Mill (now called Leachtown) at high water. It is reported that the Zanesville was the irst boat to reach Elizabeth by the Little Kanawha in April 1842. Regardless of the date, it is clear that steamboats navigated the river during high water over two decades before the slack-water navigation was opened. Soon ater the locks were opened, many boats appeared on the river. hey included the George hompson, Zebra, Naomi, Silver Heel, Do Ra Me, and others. he early boats were no wider than 22 feet nor longer than 125 feet to it into the locks. hey had a shallow draught of four feet or less. Sidewheelers and sternwheelers were the irst types of steamboat to travel the Little Kanawha. hey carried supplies to the towns upstream and provided packet transportation to all points along the river. Everything imaginable was carried on these boats. People, animals, food, farm supplies, oil drilling supplies, household goods, and anything else one can think of was probably on one of these boats at one time or another. he packets operated regular runs weekly or daily to towns along the river. he Little Kanawha Transportation Company, the Parkersburg and Creston 401

Chapter 12

(Top)Figure 12.3. he Edith H. and her captain. Note the smaller size packet. (Cam DePue Collection, West Virginia and Regional History Collection, West Virginia University Library) (Above)Figure 12.4. he Leone traveling the Little Kanawha River ater the turn of the century. (Cam DePue Collection, West Virginia and Regional History Collection, West Virginia University Library)

402

T h e L i t t l e K a n aw h a N av i g at i o n

(Top)Figure 12.5. he Dove, built in 1916, loaded with supplies and passengers. She was owned by the Richter family of Creston, West Virginia. (Cam DePue Collection, West Virginia and Regional History Collection, West Virginia University Library) (Above)Figure 12.6. he Port of Creston, West Virginia, on the Little Kanawha River. Note the various sizes of packets and the general store in background. (Cam DePue Collection, West Virginia and Regional History Collection, West Virginia University Library)

403

Chapter 12

Packet Company, and others were formed to provide passenger transportation, as well as carry goods to the people above Parkersburg. Other companies were formed for similar purposes over the seventy years that the river was used. he type of boat changed with the times. hey went from steam driven to oil driven and inally to gasoline engines. Photographs of typical boat types are shown in Figs. 12.3–12.6. J. N. Camden and Val Rathbone exchanged their Parkersburg reinery and certain properties for stock in John D. Rockefeller’s Standard Oil Trust on May 12, 1875, but the Little Kanawha locks and dams were not included in the deal. Camden now had free time on his hands and ran for a United States Senate seat, which he won in 1881. He served one full term of six years but was not reelected. He was later appointed to ill an unexpired term of two years ater the death of John E. Kenna, in 1893. hese men, and others with political inluence, advocated extending the slack-water navigation on the Little Kanawha as well as improvements to the four locks and dams. As a result, the Corps of Engineers was instructed to undertake a survey in the 1870s to see if it would be feasible to put locks and dams throughout the rest of the river from Burning Springs to Bulltown. In his report E. J. Carpenter, assistant engineer at the United States Engineers Oice in Cincinnati, Ohio, wrote that twelve additional dams and locks would be needed to complete the system to Bulltown.12 He mentioned that the irst four locks and dams cost the Little Kanawha Navigation Company $240,000, and it would cost the federal government $758,400 to provide the remaining twelve. He also stated that the natural resources and industry in the valley did not warrant the investment for the value of goods that could be obtained. Between 1876 and 1899, $213,418 was appropriated by the federal government for improving the river for navigation. hese improvements were primarily the removal of obstacles from the river to keep it open for river traic. In 1891, a ith lock and dam were constructed by the federal government just below Burning Springs. his provided transportation as far up the river as Creston. his dam remained in use until the late 1940s. Traic continued to move up and down the river ater the navigation company opened the locks in the mid-1870s. Revenues increased steadily, but now the main revenue source was lumber instead of oil. he Little Kanawha 404

T h e L i t t l e K a n aw h a N av i g at i o n

Navigation Company showed annual receipts of $12,075.91, $11,121.97, and $14,439.68 for the years 1878, 1879, and 1880 respectively. Oil shipments were now producing only 10 percent of the revenue. With the opening of Lock No. 5, slack-water navigation was now possible to Creston and, during high water, to Glenville. he federal government supervised Lock No.5 and the Little Kanawha Navigation Company continued to control the irst four. A great deal of river traic moved between towns along the river. Receipts from the diferent locks, according to Corps reports, show that the lower locks took in less revenue than did the upper ones, from goods going downriver to Parkersburg. his indicated that, although goods were coming down the river to go to outside markets, many goods and people traveled between towns along the river. In 1894, the Corps of Engineers began discussions with representatives of the Little Kanawha Navigation Company regarding purchase of their irst four locks and dams. In a letter dated October 27, 1894, from Mr. L.B. Dellicker, the secretary of the Little Kanawha Navigation Company, to Major D.W. Lockwood, Dellicker stated that the construction cost of the four locks and dams was $250,365.78. he letter further states that the “board of directors . . . will recommend to its stockholders to sell this property to the United States for $250,000 . . .” and that the company was paying a dividend of 6 percent on its capital stock.13 Negotiations were carried on for ten years among the Corps of Engineers, the directors of the navigation company, and oicials of the city of Parkersburg and Wood and Wirt counties. he city and the two counties owned $80,000 of the $158,625 of capital stock in the navigation company. he remainder of the capitalization consisted of $20,000 in 5 percent bonds. he government proposed to pay 50 cents on the dollar for the stock and give full value to the bonds, for a total purchase price of $99,312.50. All four of the locks and dams were in great need of repair. Some of the lock chambers leaked excessively, and the company was going to incur heavy costs to repair them. On April 29, 1898, the Little Kanawha Railroad made its irst run from Elizabeth to Newark, a distance of eight miles. he stockholders of the Little 405

Chapter 12

Kanawha Navigation Company now had competition for both freight and passenger tolls. he railroad was going to be completed to Parkersburg and Burnsville. With the state of disrepair in the locks and dams, and now railroad competition, the time was right to sell. he Board of Engineers for Rivers and Harbors conducted a public meeting in Parkersburg on September 25, 1903. Representatives of the city of Parkersburg and Wood and Wirt counties agreed to donate their stock (51 percent) to the federal government without cost. hey also agreed to purchase the private stock of the company for $75,000. On November 1, 1905, ater approval of Congress, the federal government took control of locks and dams nos. 1 to 4 on the Little Kanawha River. he thirty-year operation by the Little Kanawha Navigation Company had ended. River traic, however, was to go on for another thirty years. Coal and timber were still brought out of the Little Kanawha Valley in abundance. Plans were still being made by the federal government to construct locks and dams all the way to Burnsville. West Virginia Senator Stephen B. Elkins, in 1910, proposed this work and is quoted as saying “the improvement of the Little Kanawha River would have the efect of creating an enormous traic, besides keeping down rates on the railroads which is most desirable. I believe the traic which would be moved on the Little Kanawha River, if improved, would exceed that of any other river in the States.”14 Major Frederick W. Alstatter prepared plans for iteen locks between Burnsville and Parkersburg in 1910, to handle coal barges. An alternate plan proposed using the Ellet reservoir system of maintaining navigable stages, by building dams and reservoirs on Steer Creek, Sand Fork, and at Burnsville. 15 However, this plan was never implemented as improved highways and the Little Kanawha Railroad provided transportation for people and freight along the head of the slack-water navigation at Creston. Annual reports of the Corps of Engineers pertaining to the tonnage of goods and numbers of passengers exist for the years ater 1891 for Lock No. 5, but are not complete ater 1905, when the government purchased the original four locks. Judging by the number of boats (gasoline) plying the river, the heyday of river traic would probably have been between 1905 and 1915. An average of 10,000 passengers is estimated to have traveled the river during these years. 406

T h e L i t t l e K a n aw h a N av i g at i o n

Conclusions he River and Harbor Act of 1878 opened a loodgate of projects for canals and navigation such as the Little Kanawha, despite the fact that studies of a number of Ohio River tributaries by the Corps of Engineers indicated that the beneits did not justify the cost of construction. Earlier, in 1876, Congress appropriated funds to clear the upper reaches of the river for navigation during high water rather than construct the series of locks and dams recommended in Carpenter’s report in 1874. Not only were numerous snags and boulders removed and wing dams constructed, but bear-trap locks were installed at four mill dams above Glenville. Nevertheless, in 1880 Congress authorized the construction of Lock and Dam No. 5 to provide slack-water navigation nearly ity miles from the mouth of the Little Kanawha to Creston. Even this project was delayed; the new structure was not completed and put into operation until 1891. Although river traic had diminished to a trickle, there were still those who championed river improvement, this time in connection with development of hydroelectric power and provision for lood control. General Lytle Brown, chief of the U.S. Army Corps of Engineers, reported to Congress in 1931 on the condition of the Little Kanawha Navigation and prospects for its improvements. He wrote: he district engineer concludes that further improvement of the Little Kanawha River for navigation is not economically feasible at the present time, that the production of hydroelectric power at low cost is not feasible unless in connection with the combined Ohio River project, the study of which is not yet complete; that the loods of the Ohio and Mississippi Rivers will not be appreciably afected by reservoirs on the Little Kanawha. He recommends that no further improvement of the stream be undertaken at the present time. 16

he proposed hydroelectric dams were not constructed on the Ohio and its tributaries. However, as a result of severe looding in the Ohio and Mississippi 407

Chapter 12

valleys in the 1930s, the Ellet reservoir scheme was reconsidered. he Flood Control Act of 1938 provided for the construction of numerous lood-control dams, four of them on the Little Kanawha. Little was done to implement this act as far as the Little Kanawha River was concerned until the Corps of Engineers began construction of the 90-foot-high and 1,400-foot-long Burnsville Dam in 1972. he dam was placed in service December 1978, and represents the irst major improvement on the Little Kanawha River since the construction of Lock and Dam No. 5 in 1891. New methods of transportation and improved transportation systems caused changes in life along the river. As a result, when one travels the Little Kanawha today, it is hard to imagine that boats of over 100 feet in length traveled that river, let alone were built at such towns as Grantsville, Creston, and elsewhere. Chapter 12 Notes 1 2 3 4 5 6 7 8

9

Charles Henry Ambler, George Washington and the West (Chapel Hill: University of North Carolina Press, 1936), 134. Ibid., 137. Worthington Chauncey Ford, he Writings of George Washington (New York: G. P. Putnam’s Sons, 1889), 298. “Petition for a Road,” he Pioneer Daughter 2, no.5 (Parkersburg, WV, Oct 1902). Albert Gallatin, Report of the Secretary of the Treasury on the Subject of Public Roads and Canals, 1808 (republished, New York: Augustus M. Kelley, 1968). Archer Bubler Hulbert, he Great American Canals, Volume I (Cleveland: Arthur Clark Company, 1904), 40. John Pickell, A New Chapter in the Early Life of Washington (New York: D. Appleton and Company, 1856), 148. Virginia, Acts of the General Assembly, “An Act Authorizing the Erection of Mill Dams Across the Little Kanawha River” (March 11, 1934), Chapter 153, 181– 183. West Virginia, Acts of the Legislature of West Virginia at its Fith Session, “he West Virginia Transportation Company,” 252. It should be noted that the irst cross-country pipeline was not completed until 1880. It pumped oil from upstate

408

T h e L i t t l e K a n aw h a N av i g at i o n

10 11 12 13 14

15

16

New York at Olean to Bayonne, New Jersey. Edward J. Lenik, “he Olean-Bayonne Pipeline,” IA, he Journal of the Society for Industrial Archeology, 2, no. 1 (1976). James Marion Murphy, “Transportation on the Little Kanawha River in Gilmer County,” unpublished Masters thesis, 1950. Informal talk with Mr. Lockhart, owner of one of the lockmaster’s houses at Lock No. 5, Burning Springs, spring 1989. United States Congress, House of Representatives, 44th Congress, First Session, Executive Document #1, Part 2, Volume II, 1875,740–745. United States Congress, House of Representatives, 53rd Congress, hird Session, Executive Document #98, 1894, 5–6. Leland R. Johnson, Men, Mountains, and Rivers, An Illustrated History of the Huntington District, United States Army Corps of Engineers, 1754–1974 (Washington D.C.: GPO, 1977), 106 Charles Ellet, Jr. (1810–1862) was famous for his construction of long-span suspension bridges and for the role he played with his ram leet in the Battle of Memphis. Although not as well known, he became involved in river hydraulics on the lower Mississippi River. He proposed the building of lood control reservoirs on tributaries of the Ohio and Mississippi as well as the traditional levees built on the Western waters. Johnson, 68–69. United States Congress, House of Representatives, 71st Congress, hird Session, Executive Document N1.732, 1931, 3.

409

Acknowledgments

Instead of the usual form of acknowledgements which lists those who have made substantial contributions to the essays in this book, I will name the institutions in this country and abroad where the research has been conducted. First, I owe thanks to the Institution of Civil Engineers Library in London, England, and the Institute for Advanced Studies in the Humanities in Edinburgh, Scotland, where I served as a visiting Fellow, followed by brief trips to the University of Bristol. Second, in the United States, I used the research collections of the Smithsonian Institution, the Historic American Engineering Record, and the Library of Congress, all in Washington, DC. Elsewhere, I consulted the Charles Ellet Jr. collection of papers at the University of Michigan. I made short term visits to the University of Oklahoma’s Center for the History of Science and Rensselaer Polytechnic Institute’s Roebling Collection with supplementary material on Roebling at Rutgers University. I also used collections in the library at Oglebay Institute in Wheeling, WV, and the Wheeling Public Library, when I was involved in the restoration of the Wheeling Suspension Bridge and the 1859 Wheeling Custom House. I would like to ofer heartfelt thanks to the many professionals at these different libraries and colleagues, now good friends, in research and restoration projects who have aided my various endeavors in the history of technology. To the Director and staf of the West Virginia University Press I owe a very special debt of gratitude. Emory L. Kemp

410

About the Authors

Emory L. Kemp is Professor Emeritus of History and of Civil Engineering at West Virginia University. He is an international expert in the history of technology and the restoration of structures. He worked for leading engineering consulting irms in England before receving his Ph.D. in heoretical and Applied Mechanics from the University of Illinois. He joined West Virginia University in 1962 to establish a graduate program in structural engineering. He founded WVU’s program in the history of science and technology. Fostering the use of a material-culture approach for the study of the industrial past, he has researched and preserved historic industrial sites around the country and overseas and has advocated their public interpretation. Kemp is a founding member and past president of the society for Industrial Archeology, and past president of the Public Works Historical Society. He is a distinguished member of the American Society of Civil Engineers and has received numerous other prestigious awards. Robert J. Kapsch, PhD, Hon. AIA, ASCE, spent iteen years as the Chief of Historic American Buildings Survey/Historic American Engineering Record, the U.S. government’s premier documentation program and has served as project engineer for many historic restoration and rehabilitation projects along the Chesapeake and Ohio Canal and other parks. He is the author of several books and articles on historic architecture and engineering, including he Potomac Canal: George Washington and the Waterway West (West Virginia University Press); Historic Canals and Waterways of South Carolina (University of South Carolina Press); CANALS, the Norton/Library of Congress Visual Sourcebook in Architecture, Design, and Engineering; and Over the Alleghenies: Early Canals and Railroads of Pennsylvania (West Virginia University Press). 411

A b o u t t h e Au t h o r s

Lance E. Metz received his B.A. in history from Moravian College and his M.A. in maritime history from the University of Maine. He was the historian for the National Canal Museum, Easton, Pa, from 1980 to 2010, and is currently the emeritus historian of the Museum. He is a past president of the Roebling Chapter of the Society for Industrial Archeology and was a national board member of the Society, as well as of the American and Pennsylvania Canal Societies.  He is co-author of Anthracite Iron Industry of the Lehigh Valley (with Craig Bartholomew, Canal History and Technology Press, 1988) and numerous other works. He was the historian for the HAER/NPS recording project at the Bethlehem Plant of Bethlehem Steel, and the lead researcher and writer for the Historic Resource Study of the Delaware and Lehigh National Heritage Corridor, and has served as a consultant for numerous canal preservation and restoration projects as well as various media productions.

412

Emory L. Kemp is the founder and director of the Institute for the History of Technology and Industrial Archaeology at West Virginia University, where he also served as a chair and professor of civil engineering and a professor of history. his collection of essays encompasses over ity years of his research in the ield of the history of technology. Within these twelve essays, Kemp describes and analyzes nineteenth-century improvements in building materials such as iron, steel, and cement; roads and bridges, especially the evolution of the suspension bridge; canals and navigable rivers, including the Ohio River and its tributaries; and water supply systems. As one of the few practicing American engineers who also research and write about civil engineering, Kemp adds an important historical context to his work by focusing not only on the construction of a structure but also on the analytical science that heralds a structure’s design and development.

west virginia university press www.wvupress.com