Future Energy Options from a Systems Perspective 3031464478, 9783031464478

​This book is an exploration of energy and its unique role to date as one of the fundamental enabling and controlling fa

110 72 2MB

English Pages 121 [118] Year 2023

Report DMCA / Copyright

DOWNLOAD PDF FILE

Table of contents :
Contents
List of Figures
1 Introduction
Energy as a Critical Phenomenon
Phases of Human Energy Use Through Time
Taming of Fire
Agriculture and Animal Metabolic Power
Natural Energy Flows, Whale Oil and Human Slavery
Fossil Fuels and Industrialisation
Nuclear and Renewables
Themes Through the Evolution of Human Energy Use
Succession and Interdependence of Energy Sources and Types
Geography and Portability
Energy Return on Investment
Complexity, Energy and the Superorganism
Current Global Energy Situation—The Global ‘Energy Bind’
References
2 The Trifurcation of Energy Futures
The Fossil-Seneca Branch
Climate Change and Depletion
Characteristics and Status of Nuclear Fission Technology
The Seneca Effect
Assessment of the Fossil-Seneca Branch
The Continued Growth Branch
Technologies with Potential to Underpin Continued Growth
Characteristics, Potential and Constraints to Nuclear Fusion Technology
Renewables
Prospects and Consequences of Continued Growth
Global Economic and Social Paradigms
Physical and Digital Footprint
Energy Accumulation
Assessment of the Continued Growth Branch
The Stabilisation Branch
Long-Term Stability of Complex Systems
Energy Technologies and Stabilisation
A Renewables-Based Civilisation
Challenges for the Renewables Transition
Systemic Challenges
Re-materialisation
Externalities of Renewables
Constraints to New Frontiers for Renewables
Enablers and Opportunities for the Renewables Transition
Degrowth and Changes to Per Capita Energy Use
Renewables Circular Economy
Interactions with the Food System
Potential Hybrid Energy Systems
Assessment of the Renewables Stabilisation Branch
References
3 Conclusions
Fossil Fuel Systemic Inertia
Longer Term Perspectives
References
Index
Recommend Papers

Future Energy Options from a Systems Perspective
 3031464478, 9783031464478

  • 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

Future Energy Options from a Systems Perspective

Nick King Aled Jones

Future Energy Options from a Systems Perspective

Nick King · Aled Jones

Future Energy Options from a Systems Perspective

Nick King Global Sustainability Institute Anglia Ruskin University Cambridge, UK

Aled Jones Global Sustainability Institute Anglia Ruskin University Cambridge, UK

ISBN 978-3-031-46447-8 ISBN 978-3-031-46448-5 (eBook) https://doi.org/10.1007/978-3-031-46448-5 © The Editor(s) (if applicable) and The Author(s), under exclusive license to Springer Nature Switzerland AG 2023 This work is subject to copyright. All rights are solely and exclusively licensed by the Publisher, whether the whole or part of the material is concerned, specifically the rights of translation, reprinting, reuse of illustrations, recitation, broadcasting, reproduction on microfilms or in any other physical way, and transmission or information storage and retrieval, electronic adaptation, computer software, or by similar or dissimilar methodology now known or hereafter developed. The use of general descriptive names, registered names, trademarks, service marks, etc. in this publication does not imply, even in the absence of a specific statement, that such names are exempt from the relevant protective laws and regulations and therefore free for general use. The publisher, the authors, and the editors are safe to assume that the advice and information in this book are believed to be true and accurate at the date of publication. Neither the publisher nor the authors or the editors give a warranty, expressed or implied, with respect to the material contained herein or for any errors or omissions that may have been made. The publisher remains neutral with regard to jurisdictional claims in published maps and institutional affiliations. Cover credit: © Melisa Hasan This Palgrave Macmillan imprint is published by the registered company Springer Nature Switzerland AG The registered company address is: Gewerbestrasse 11, 6330 Cham, Switzerland Paper in this product is recyclable.

To my dad Richard, whose inspirational interest in science, the natural world and our impact upon it, provided me with the desire to try to better understand our predicament. —Nick King

Contents

1

Introduction Energy as a Critical Phenomenon Phases of Human Energy Use Through Time References

1 2 3 25

2

The Trifurcation of Energy Futures The Fossil-Seneca Branch The Continued Growth Branch The Stabilisation Branch References

33 34 48 64 80

3

Conclusions Fossil Fuel Systemic Inertia Longer Term Perspectives References

Index

99 102 105 106 109

vii

List of Figures

Fig. 2.1

Fig. 2.2

Fig. 2.3

Fossil-Seneca Branch—illustration of system behaviour (Notes y-axis represent energy usage and the x-axis time units; all units are arbitrary and for the purpose of illustration of mode of system behaviour only. Energy usage is the metric applied to demonstrate collapse of human system function, but other measures [e.g., human population] could be used to demonstrate the effect as well) Continued growth branch—illustration of system behaviour (Notes y-axis represent energy usage and the x-axis time units; all units are arbitrary and for the purpose of illustration of mode of system behaviour only. The gradient of the line at different points in time and the energy consumption levels attained are illustrative only) Renewables Stabilisation Branch—illustration of system behaviour (Notes y-axis represent energy usage and the x-axis time units; all units are arbitrary and for the purpose of illustration of mode of system behaviour only. The gradient of the line at different points in time and the energy consumption levels attained are illustrative only. The ‘Energy Input’ level is illustrative and is not representative of any physical values)

47

52

67

ix

CHAPTER 1

Introduction

Abstract Exosomatic (non-metabolic) energy is a phenomenon unique to humans, and in recent timeframes has become a fundamental underpinning for the continued operation of technological civilisation. A number of different energy sources have been used by humans over time (fire; agriculture and animal metabolic power; natural energy flows, whale oil and human slavery; fossil fuels; and renewables and nuclear energy), a trend which has been characterised by successions and interdependence and steered by factors such as geography, portability, return on investment and complexity. Each energy succession has brought about varied societal changes, but the flourishing of the use of fossil fuels has resulted in the most profound changes to human society. An ‘energy bind’ has emerged from the reliance of complex global society on energy to function, but growing constraints to energy (particularly fossil fuel) availability is leading to growing risks, labelled collectively as the ‘polycrisis’. Keywords Endosomatic energy · Exosomatic energy · Energy bind · Energy successions · Great Acceleration · Energy Return on Investment · Complexity · Superorganism

© The Author(s), under exclusive license to Springer Nature Switzerland AG 2023 N. King and A. Jones, Future Energy Options from a Systems Perspective, https://doi.org/10.1007/978-3-031-46448-5_1

1

2

N. KING AND A. JONES

Energy as a Critical Phenomenon Global civilisation in the early twenty-first century faces a multitude of severe, accumulating, converging, and interacting systemic challenges, labelled collectively as the ‘polycrisis’ [1–3]. One of the most significant manifestations of this confluence of natural and anthropogenic threats and hazards is the ‘Water–Energy–Food Nexus ’ (WEF nexus). This concept links three of the systems and commodities which are most critical to the functioning of human societies, describes how they are likely to be subject to constraints and challenges, and encompasses their multi-faceted interdependence, and as such reaches into the heart of the complexity of the polycrisis [4–6]. The provision of food and water are biological imperatives for the day to day survival of humans and other heterotrophic organisms, and failures in the steady and accessible supply of these commodities threatens both the stability of societies and the survival of those who inhabit them [7]. Energy differs fundamentally from food and water in that it has different forms and applications which relate to human necessity via complex mechanisms. Endosomatic energy is the energy obtained by humans from the consumption of food and which is made available via cellular metabolism and directed to use via biological systems, such as muscle power and cognition and has been the primary source at the disposal of humans for most of history. Exosomatic energy by contrast is the energy available in systems outside of direct human metabolism, which is commandeered and directed via human intelligence, and although this has been in use in some forms through most of human history, it has only started to be used by humans at scale relatively recently [8]. Endosomatic energy is a fundamental factor in survival and is directly linked to the availability and consumption of food, whereas exosomatic energy is not directly required for survival in the biological sense and is instead a technological ‘add-on’ to human energy use. However, exosomatic energy has, in recent timeframes, become a fundamental basis for the creation, expansion, maintenance and complexification of a global spanning, technological civilisation, and therefore increasingly vital to human survival. This is because the systems which underpin civilisation, such as agriculture and water supply, have become increasingly large-scale, technological, and dependent on energy inputs to function. Therefore, the provision of food and water has (particularly in recent timeframes) become highly dependent on exosomatic energy inputs, having previously

1

INTRODUCTION

3

been provided in sufficient quantities by natural systems (or with only limited exosomatic energy input) [9]. Collective humanity has created a paradigm in which exosomatic energy, which is otherwise a technological artefact unique to humans rather than a crucial factor in survival, has become critical through its progressively close interlinking with the provision of food and energy. Human civilisation is a globe-spanning enterprise comprising eight billion inhabitants, and is now reliant on the provision of continuous supplies of exosomatic energy to drive crucial technologies and systems which together permit large human populations and organised societies to function and persist. This is partly through the provision of food and water (via energetically-driven production and distribution) in quantities sufficient to support populations of these magnitudes, but also in driving the other complex anthropogenic systems (separate or only indirectly linked to water and food) which operate in parallel and maintain the functions of civilisation, including the collation of knowledge. As such, energy has become, in modern timeframes, a critical factor and ‘enabler’ for virtually all aspects of human civilisation. Due to this ‘keystone’ role, exosomatic energy will likely be the fundamental factor in the shape of human civilisation in the future; as the challenges of the ‘polycrisis’ accumulate, collective humanity will increasingly find itself in an ‘energy bind’ which will make the selection of particular energy ‘futures’ ever more of an imperative. This book explores this ‘energy bind’, and outlines the largescale, systemic energy pathways which likely lie before collective humanity, and what each of their implications are.

Phases of Human Energy Use Through Time This section outlines the different phases of energy use by humans. Over time the forms of energy and the means by which they have been obtained, harnessed and applied, has undergone a number of transitions which have led to changes in the energy types and technologies which have predominated. The following subsections provide a chronological description of each of these phases and the various factors which led to the emergence of each new paradigm.

4

N. KING AND A. JONES

Taming of Fire The emergence of exosomatic energy1 generation and use occurred prior to the evolution of anatomically modern humans, approximately one and a half million years ago. Global climatic changes (a general global cooling trend likely linked to changing atmospheric CO2 levels) starting in the late Miocene (approximately five–seven million years ago) gave rise to generally rising aridity and seasonality, resulting in global changes to terrestrial ecosystems. In tropical East Africa, ecosystem changes manifested as a transition of the predominant vegetation cover from dense moist tropical forest to drier, more open savannah [10]. This transition resulted in changes to animal biota, and for tree-dwelling ape species in the region, this became a key factor in the evolution of more preferentially ground-dwelling hominids. In parallel, this drier environment experienced a greater prevalence of natural (primarily lightning-instigated) fires, which inevitably led hominid encounters with these natural fires on a more frequent basis. This likely led to a process of increasingly direct interaction and use, and eventually the capability to start fires independently and artificially (using deliberately-collected biomass, primarily wood, and culturally-transmitted techniques). The growing prevalence of the phenomenon from approximately one and a half million years ago is evidenced through widespread archaeological evidence of burning associated with hominid sites, and from approximately 400 to 700 thousand years ago, the appearance of hearths indicating increasingly refined use of fire. Control of (biomass-fuelled) fire (‘pyrotechnology’) is likely to have had a significant influence on the evolution of modern humans through factors such as reduced mortality by predation, increased general activity through ‘daylight extension’, and greater access to nutrients from the cooking of food (which functions as a form of ‘pre-digestion’), all of which may have contributed to the evolution of larger brains and subsequent phenomena such as language and increased sociality. The enhancing feedbacks of fire’s benefits and widening use likely led to human society re-organising over time to make this type of energy use a central and permanent aspect of daily life, and which laid the foundation for later, larger scale and more sophisticated applications of fire (such as land clearing and metalworking) [11, 12]. 1 ‘Energy’ will refer hereafter to exosomatic energy unless otherwise stated.

1

INTRODUCTION

5

Agriculture and Animal Metabolic Power The Agricultural Revolution (also described as the Neolithic Revolution) describes the emergence of organised agriculture, consisting of the planned and regular cultivation of particular plant species to provide food crops, combined with animal husbandry for food and other applications, with these activities taking place within particular, fixed geographical areas. The predominant mode of obtaining food amongst different human groups shifted over time, generating a gradual transition from hunting and gathering towards agriculture (and in parallel, pastoralism), with a likely significant overlap between the two before agriculture emerged as the pre-eminent mode for the overwhelming proportion of the population (from that period onwards), due to an accumulation of self-reinforcing advantages. This phenomenon emerged simultaneously in at least seven (and possibly more) separate locations globally starting approximately twelve thousand years ago, where conditions (including, but not limited to, climate and soil conditions, the composition of the biome, and human population density) were favourable. It is likely that complex interplays of drivers motivated different groups of humans to adopt agriculture as a primary mode of living (including, but not limited to, changing global climatic conditions, over-hunting of local megafauna, and desire for more predictable and resilient food supplies), but what is clearer than these remote motivations is the profound impact on virtually every aspect of human lifestyle and society that the Agricultural Revolution had. Most fundamentally, the provision of stable and eventually surplus food supplies from agriculture allowed the overall human population to grow in a sustained manner. This in turn allowed the development of new phenomena including, but not limited to: fixed sedentary settlements in proximity to agriculturally productive land; social hierarchies and religions in response to rising urban populations and consequent new social dynamics; militaries to defend stored food surpluses/acquire new surpluses; money and trade to mediate exchange and control of surpluses; application of fire to enable new metal and ceramic technologies; and a range of other phenomena that have continued into the modern era [13–17]. This energy transition phase differs from the other revolutions, technologies and systems described in the following subsections in that it was a systemic change in the broad scale strategies applied by human groups to increase the total quantity and the stability of the food supply.

6

N. KING AND A. JONES

Therefore, it was largely an effort to increase the amount of endosomatic energy available to humans. However, one aspect of the new behaviours and technologies which emerged in the Agricultural Revolution, namely animal husbandry, did represent a hitherto unused form of exosomatic energy. Specifically, this was the controlled use of animal metabolism (the muscle power of particular domesticated animal species such as cattle) under human direction for tasks which couldn’t otherwise be achieved such as the pulling of ploughs (human metabolic power alone was generally insufficient and fire-derived energy could not be directed to the required applications with the technologies available at the time) [18]. Furthermore, the human population growth enabled by the Agricultural Revolution provided the basis for later (exosomatic) energy revolutions; therefore, this phase represented a complex interplay of factors which were in aggregate highly significant for overall human energy use. Natural Energy Flows, Whale Oil and Human Slavery The fixed settlements, expanding populations, technological advancements and existing energy sources (primarily pyrotechnology) enabled by the spread of agriculture allowed various human societies to conceive, build and spread mechanisms and structures that opened new energy paradigms. These new mechanisms permitted the utilisation of the regular natural kinetic energy flows contained in environmental media (namely water and air) via direct conversion to mechanical power, for a range of applications. From approximately five thousand years ago, geographically widespread groups and societies started utilising waterwheels on rivers to drive simple machines which required steady power input, for example grindstones for grain milling, bellows for metallurgical furnaces, and for the processing of cloth. Windmills were pioneered and subsequently spread across many regions from approximately one thousand years ago and their spread was more rapid than for waterwheels as they offered the key advantage in that wind energy resources were much more widely distributed and therefore available for use by a greater proportion of the population. The exploitation of wind power proliferated significantly in Europe between approximately the eleventh and nineteenth centuries, widening to applications such as pumping of water and the mass processing of commodities such as foodstuffs, textiles, paper, and timber. The total number of windmills in Europe peaked at approximately two hundred

1

INTRODUCTION

7

thousand in approximately 1900; the growth to this peak coincided with the emergence of concentrated zones of the collection and application of wind energy, exemplified by the Zaan District (in the modern day Netherlands) during the seventeenth–eighteenth centuries. Approximately one thousand windmills powered co-ordinated, large-scale industrial activity, with the intensive processing of timber being a focus of the wind-powered technologies pioneered there. The exploitation of these energy flows became a permanent feature of many societies due to their ‘free at point of use’ nature, along with the productivity gains and freeing up of human labour they provided [19, 20]. Another significant source of energy which was exploited at scale in parallel to natural energy flows was biomass from hunted cetaceans, specifically ‘whale oil’ (obtained indirectly from the rendering of blubber tissue from several different whale species) and ‘sperm oil’ (obtained directly as an oily liquid from Sperm whales). The primary energetic application for whale oil was as a lighting fuel, for which it was used all over the world (alongside non-energetic uses such as for base for detergent and a machinery lubricant) through the nineteenth and early twentieth centuries. The hunting of whales for meat and other products dates back to antiquity but whaling to supply commercial demands (primarily to supply whale oil for inexpensive urban lighting) originated in the Middle Ages and peaked in approximately the 1850s. The growth in demand for whale oil, combined with declining whale populations and technological advancement meant that there was a progression through preferred species (Right to Bowhead to Sperm), though demand eventually dropped off significantly following development of coal gasification and (‘rock’) oil extraction. Note that whaling continued following the decline in lighting oil demand (to supply food and other products), and did not peak until the 1960s, at which point changes in international law and extreme whale population depletion finally brought this practice to an (almost) end [21, 22]. Another energy paradigm which grew and peaked on approximately similar timescales to the exploitation of natural energy flows and whale oil was human slavey, which is defined as the treatment of human beings as owned commodities such that their metabolism can be controlled and directed for tasks against their will. More specifically, this involved the commandeering of human endosomatic energy as exosomatic energy by other humans, primarily for tasks such as agricultural and construction labour. Although this egregious and highly exploitative practice has likely

8

N. KING AND A. JONES

been a feature of organised human societies since the Agricultural Revolution (and potentially earlier), and trade in and use of mass numbers of human slaves is linked with societies as varied as the Roman Empire, the Vikings and the Barbary States, the period and location in which the most intense trade occurred was in the Trans-Atlantic region (between Europe, Africa and the Americas) from the sixteenth–nineteenth centuries (the three centuries preceding the Industrial Revolution). An estimated twelve million people were enslaved and traded during that timeframe and provided a large proportion of the energy which underpinned the physical and economic development of the Americas and the Caribbean [23, 24], and by extension Europe and other regions globally. Fossil Fuels and Industrialisation The next revolution in human energy appropriation and application has been the most profound and far reaching in human history to date, namely the exploitation of the energy stored within lithospheric fossil carbon deposits (fossil fuels). The extent and scale of the use of these fuels over time has been highly nonlinear, but application for mass industrial purposes (starting with the Industrial Revolution and intensifying in recent timeframes), has been the most impactful. Fossil fuels in various forms were known to and underwent limited use by societies through antiquity, but these uses remained relatively niche (primarily due to lack of technology suitable and necessary for their mass exploitation, combined with a lack of understanding of their potential) until particular contexts and events set in motion events that led to the flourishing of their use, and consequent transformations of societies at global scale. The first ‘stage’ of this new energy paradigm took place on the island of Great Britain; during the centuries preceding the Industrial Revolution biomass (primarily in the form of wood and charcoal) had been the primary fuel for ‘pyrotechnology’ applications such as simple industrial processes (e.g., lime kilns), and for space heating in large population centres. Growing constraints on the availability of this biomass over time (from the sixteenth century onwards, largely due to deforestation), combined with changes in land ownership, drove the increasing domestic and industrial use of coal. This fuel had been in use previously at small scale in more niche roles, but its relative expense and the smoke produced by its use had led to biomass being preferentially used.

1

INTRODUCTION

9

Great Britain was endowed with large reserves of high quality (i.e., high carbon and energy content) coal, with several large and relatively accessible (near-surface) coalfields distributed across the island. This endowment of resources, combined with growing demand from the innovation of coal-fired, steam-driven machinery (primarily for applications including mechanical production of commodities and for sea and land transport), set in motion enhancing feedbacks. These included widening applications (larger and improved steam engines), increasing coal mining activity (opening of new coalfields and mines), expanding infrastructure (canals and railways to transport coal), technological innovations (groundwater pumps to open up previously inaccessible reserves) and a shifting cultural setting (migration to urban areas) which drove steadily increasing penetration of coal use into society at large over the course of several decades. This process culminated with an acceleration, expansion and embedding of these technological and social trends during the approximate period 1750–1800, which marked the start of the Industrial Revolution. These industrial applications, along with others such as coal gasification which expanded rapidly for urban lighting and heating applications, rapidly spread to other parts of the world where coal resources were also abundant (e.g., Germany and the USA) or could be accessed through importation. The second ‘stage’ in the flourishing of fossil fuel use started in Pennsylvania in the northeastern US, where exploratory drilling was undertaken during the late 1850s to assess whether ‘rock’ (or ‘mineral’) oil known to be present in the subsurface could be practically extracted. These near surface reservoirs had previously been known about due to abundant natural ground seepages, but the wider context was the strong incentive to find a practical replacement to whale oil, the supply of which was increasingly subject to the effects of depletion. Successful drilling of wells led to the development of the world’s first full-scale oilfield during the 1850–1860s, and different fractions of the ‘rock’ oil obtained were successful in replacing whale oil (for lighting). However, it was several subsequent technological innovations (primarily in the form of the invention of the Otto and Diesel cycle internal combustion engines) which set in motion enhancing feedbacks (equivalent to those that drove the earlier flourishing of coal use) which led thereafter to a continuous, decadeslong global spread and scaling of oil extraction and use which accelerated at the major inflection points of the World Wars [25]. In contemporary timeframes oil (primarily diesel) fuel underpins global-spanning (road, rail

10

N. KING AND A. JONES

and sea-based) logistical and supply chains which now move a greater tonnage of material annually than natural processes (e.g., erosion and fluvial transport) [26, 27]. Natural gas emerged during the twentieth century as the third ‘stage’ of the flourishing of fossil fuel. Natural gas forms from the ‘cracking’ of long-chain hydrocarbon molecules by lithospheric heat and pressure and is therefore closely associated with crude oil. In the early stages of oil extraction natural gas was primary vented or flared, but from approximately 1900 an increasing proportion was captured for use, initially to displace gasified coal for lighting, but later for an expanding range of uses. A significant application which emerged from approximately the mid-twentieth century onwards was as the main energetic feedstock underpinning the industrial-scale use of Haber–Bosch process (for synthetic fertiliser production), which drove a sustained global agricultural intensification in the late twentieth century [28]), which in turn underpinned exponential human population growth (at which point fossil energy became a key input to the generation of human biomass). Natural gas also became a widespread source of industrial heat (e.g., for cement manufacturing and metallurgy) from the late twentieth century onwards, alongside a number of additional, primarily domestic applications (for space and water heating, cooking etc.) which emerged in densely populated regions where urban gas distribution networks were economical to build and operate. A major technological innovation which emerged in the late nineteenth century and subsequently drove a large proportion of fossil fuel consumption through the twentieth and into the early twenty-first centuries was electrification; fossil-fired thermal generation of electrical power has been the consistent and universal user of all three major forms of fossil fuel (though in contemporary timeframes coal and natural gas have emerged as the dominant fuels for power generation, and oil for most modes of transport). The range of hitherto impossible applications and technologies which conversion of fossil chemical energy to electrical power enabled was vast, and the efficiency and flexibility gains which all forms of electrical technology enabled in industrial and domestic settings provided the basis for whole new paradigms in the consumption and use of fossil fuels [29].

1

INTRODUCTION

11

From these initially slow and semi-accidental beginnings over several centuries, use of all types of fossil fuels’ energy content2 grew exponentially over the twentieth century and in contemporary timeframes it dominates global energy usage (contributing 82% of global primary energy use in 2021) [30, 31]. The flourishing of fossil fuel use and the Industrial Revolution which these fuels underpinned marked a true systemic change in human energy usage, as exploitation of this huge energy reserve enabled the start of the large-scale and organised application of energy for purposes in addition to food production. Natural energy flows that had started to be captured prior to the Industrial Revolution were largely (though not only) applied to support food production, but with the emergence of fossil-fuelled industrial production new phenomena such as mass-produced commodities, consumerism, sustained economic growth and mass mobility emerged as major phenomena and drivers in the dynamics and growth of human civilisation. Nuclear and Renewables Fossil fuels continue to dominate the total energy consumption of global civilisation, but in recent decades two additional, broad classes of energy technology have been developed and put into use in widespread locations around the world and make significant overall contributions to total energy use. The first of these, renewable energy, involves technologies and systems which extract energy from natural flows in a similar manner to those historical technologies described above. Two of the types of modern renewable energy technology (wind and hydroelectric power) are direct ‘descendants’ of the technologies employed in previous centuries in terms of their general principle of operation but employ modern engineering materials and convert the natural energy flows into electricity instead of mechanical power. In addition to these historically-exploited energy sources, modern technologies and materials have allowed additional natural gradients to be exploited at scale (for electricity and also

2 Fossil fuels also provide the material substrate for commodities vital to the functioning of modern civilisation such as plastics, pharmaceuticals and petrochemicals, but are omitted from consideration due to these comprising non-energy uses.

12

N. KING AND A. JONES

applications such as space heating); namely ocean waves, tidal flows,3 geothermal heat, and solar radiation. Geothermal energy harnesses the thermal gradient between the Earth’s interior and the surface; the temperature varies from approximately 5000 °C in the core regions to several hundred °C in the upper mantle, primarily as a result of persisting primordial and ongoing radiogenic heat. The heat flux at the Earth’s surface is generally very low but in tectonically/volcanically active regions (and certain other geological settings) the localised flux can be much higher. Human societies have been aware of geothermal energy since antiquity (primarily in the form of hot springs) but it is only with modern technologies that this has been exploitable for controlled power production (also for space heating to a lesser extent). Solar power, which harnesses the energy of solar radiation directly, was not possible historically due to the absence of certain materials and capabilities not available until the modern era (though it was previously applied for niche applications such as directed seawater evaporation and drying of materials including food) but can now be exploited at industrial scale wherever incident solar radiation is available (though latitudinal and climatic constraints apply to the total amount of energy which can be collected at different locations globally). Renewables of all forms have grown significantly in term of installed capacity, global distribution and penetration into national and regional power grids in recent decades. This is a phenomenon which has been driven by factors such as growing awareness and concerns amongst governments, private companies and the public over greenhouse gas (GHG) emissions and their accumulation in the atmosphere as a result of fossil fuel use, and the reducing costs of renewables generation enabled by technological advancements, economies of scale and government policies and subsidies. As with the development and spread of fossil fuel technologies, modern renewables technologies have developed through successive phases in accordance with complex societal and economic factors, and certain technologies (hydroelectric, wind and solar, and to lesser extent geothermal) have emerged as dominant in terms of overall output. Modern renewables, especially onshore wind technologies, are now the cheapest form of electricity generation in the world in most regions. 3 Tidal and wave energy are not considered further due to their slow growth in capacity and current minor contribution to global energy supplies.

1

INTRODUCTION

13

The first of the modern renewables technologies to undergo large scale development was hydroelectric power; the inception of this technology occurred in the mid-nineteenth century when the Francis turbine was developed. Following periods of strong growth in capacity globally during the 1940–1970s and 2000–2020s, hydroelectric plants now have the largest installed renewable energy capacity by a large margin (including the largest single power plant in the world, the Three Gorges Dam, at 22.5 GW capacity) [32]. Hydroelectricity dominated renewables capacity for multiple decades prior to the development of modern wind energy during the 1980s, and which has since undergone very strong growth from the start of the twenty-first century onwards. Global onand off-shore wind capacity has undergone an approximate two order of magnitude increase during that period, with offshore capacity and increasingly large individual turbines making growing proportional contributions with time [33]. Onshore renewable capacity growth has often been curtailed by policy considerations (where there has been a perception that communities do not want large turbines located near their homes) rather than technological ones. Geothermal energy is produced through three main technologies/ approaches (dry steam, flash steam and binary cycle power), and have deployed at increasing scale since the start of the twenty-first century, though deployment is concentrated in regions and countries with conducive geological conditions (notably the USA, Indonesia, Iceland and New Zealand) [34]. Solar energy is harnessed through two different approaches: photovoltaic (PV) systems, which use (primarily) silicon-based semiconductor cells and panels to directly generate electricity via the photoelectric effect; and concentrated solar power (CSP), in which reflectors or lenses are used to concentrate solar radiation onto a receiver to raise steam and generate electricity via turbines. PV has largely overtaken CSP as the preferred solar technology and has experienced very large capacity increases since the start of the twenty-first century (solar has experienced growth rates second only to wind) [35]. This growth has often been driven by government policy, such as feed-in-tariffs, which offer incentives for domestic installations and large scale rural deployment especially in countries including Germany, coupled with large investments in manufacturing in countries including China bringing down the costs of solar exponentially.

14

N. KING AND A. JONES

The second of these energy sources, nuclear power, utilises the energy stored within heavy nuclei (actinides), which is released via artificially induced and controlled nuclear reactions. Due to the advanced scientific understanding that underpins its exploitation, and the complexity and high-technology systems and materials required to enable it, it is a purely modern technology. Nuclear science developed through a succession of theoretical and experimental advances in the late nineteenth and early twentieth centuries; the latter of these coincided with the start of World War II (WWII) which provided an incentive to pursue the potential of nuclear energy (primarily for weapons, initially), and drove the development, testing and deployment of many key early technologies. Following the use of nuclear weapons and the conclusion of WWII, the (peaceful) potential for this technology to provide a new source of energy was explored, though the military aspect remained with many of the wide range of reactor designs explored in these early stages having a dual energy and weapons material production function, whilst others were tailored to naval propulsion. Key technologies were later released from national government control and commercialised, which led to a globally distributed fleet of reactors being constructed in the following decades. This dual use (military and civilian) meant the initially high costs of nuclear technology could be cross subsidised between defence and energy government budgets. The Light Water Reactor (LWR) type emerged as the numerically dominant technology with the largest total installed capacity globally; the approximate period 1950–1980 saw the most intense phase of growth in global nuclear capacity, which was followed by a period of stagnation (due to a range of factors including major accidents and changing economic conditions), followed by a limited (relative to the initial growth phase) ‘nuclear renaissance’ after approximately 2000 [36]. Themes Through the Evolution of Human Energy Use This section identifies common themes apparent through the shifts in energy paradigms identified in the previous subsection, the systemic factors which influenced them, and the nature of the energy ‘bind’ in which contemporary human civilisation finds itself in as a result of the long evolution of its modes of energy use.

1

INTRODUCTION

15

Succession and Interdependence of Energy Sources and Types The preceding sections describe the ‘grand arc’ of human energy appropriation and use from the earliest days of the human species (including related but now extinct hominids) through to the modern era. In considering this long and complex journey it is clear that a fundamental theme is that rather than a series of successions of energy types completely replacing and superseding ‘incumbents’, it has instead been a case of new energy types and technologies being pioneered and spread whilst the preceding types remain in use to some degree (and in many cases were applied to leverage and work synergistically with the new sources). Additionally, several of the established energy technologies underwent evolution to new forms (whilst retaining their fundamental character), and/or ‘reappeared’ later as a result of technological innovations. Overall, this has resulted in total primary energy use by collective humanity increasing by approximately five orders of magnitude from the Palaeolithic through to the modern era [37, 38]. Fire (and pyrotechnology) was the first form of energy appropriation and use by humans, and represented a major event in the evolution of life on Earth in that it was the first ‘leap’ by an organism to deliberately obtain and control exosomatic energy. Fire use by humans was initially only in the ‘open’ form i.e., burned freely in the open atmosphere, even when spatially contained in hearths, and was based on the combustion of biomass. Although this was revolutionary in many ways, the constrained energy output of such fires limited their applications to heating, lighting and cooking. With the Agricultural Revolution animal husbandry opened up a new energy source in the form of the control of animal metabolic output, but biomass-based fire continued to be used in parallel to this, but in new forms and for increasingly controlled and sophisticated applications. A key example is the emergence of charcoal fires and metallurgy, which also allowed increasingly effective harnessing of animal power i.e., through obtaining and forging metals (copper and later iron) into ploughs, which were much more effective than wooden ploughs and therefore allowed greater agricultural success [39]. The exploitation of natural energy flows to obtain mechanical power, along with human slavery, increased in scope over many centuries as complex societies arose following the Agricultural Revolution. Fire however continued to be applied for basic (space heating and cooking) and increasingly advanced applications (such as increasingly sophisticated metallurgy, and glassmaking) whilst animal-based power remained very

16

N. KING AND A. JONES

widespread (e.g., for ploughing, and moving heavy loads over long distances, including the materials required to construct energy-harvesting infrastructure such as waterwheels). The pioneering of fossil fuel use and the subsequent Industrial Revolution is described as revolutionary because energy use started to dramatically increase in terms of total amount, along with the range of applications. The energy sources which had been harnessed subsequent to the pioneering of fire and pyrotechnology were not combustion-based (i.e., flowing water), but the use of fossil fuels reversed this trend in that it represented a return to, and then dramatic spread in the prevalence and use of direct combustion. The nature of the fire used in this context was different to that used previously in that it was based on fossil carbon (which has a higher energy density than biomass, and is found in concentrated reservoirs such as coal seams) and was also utilised in much more controlled manner (i.e., with burning taking place within the combustion chambers of machinery rather than primarily in open hearths) but was nonetheless a return to a combustion-dominated energy paradigm. In the early stages of the Industrial Revolution previously key energy sources such as animal-based power remained in widespread use, but were increasingly (though never totally, even in the twenty-first century) displaced by fossil-fuelled processes and machinery. The stocks (primarily in accessible biomass) and environmental flows (which were accessible through early technologies) available to premodern societies had an overall magnitude which constrained their overall societal size and growth rate. The large stock of energy provided by accessing the global reservoirs of fossil fuel energy changed that paradigm, and consequently has been the primary driver underpinning a recent phenomenon labelled as the ‘Great Acceleration’ (a ‘sub-phenomenon’ of the wider ‘Anthropocene’ Epoch) [40]. This is defined by energy use, along with virtually every other measure of human activity (including, but not limited to, overall human population; use of water, minerals and materials; and economic growth as measured by gross domestic production (GDP); and financial activity) having increased dramatically in the mid-twentieth century (after WWII), taking on near-exponential characteristics in recent decades. This required the dramatic increase in the extraction and consumption (total quantity and rate of use) of fossil fuels, and also the addition of energy from new sources (namely renewables and nuclear) [38].

1

INTRODUCTION

17

These new energy sources were also enabled by fossil fuels; the societal complexity which underpinned the population base and education systems which led to scientific innovations, along with the financing necessary for the development of renewable and nuclear energy, are phenomena resulting from fossil fuel use, as well as of course the direct use of fossil energy to mine key resources, transport them and build renewable and nuclear infrastructure. The industrial processes required to manufacture high-tech and complex materials and components required for these systems, were also largely constructed with and operated by fossil fuel energy. Geography and Portability Two key factors which were instrumental in the innovation, spread and relative success of different energy paradigms over time were geography (position and access of energy sources relative to users according to landscape, topographical and other factors), and portability (amenability to and ease of transport and distribution of energy sources to desired points of use). Early human use of fire was not significantly constrained by either of these factors as biomass was generally readily available in biomes inhabited by humans, and therefore could be collected from the local environment (in the small quantities involved), as required. The Agricultural Revolution was partially driven by constraints to the portability of food for mobile bands of humans; the shift to sedentary modes of living largely eliminated the need to transport food (for survival, though this occurred later as food was traded) but was the first instance where geography started to shape energy paradigms, as agriculture could only become established and be maintained through time where climatic, soil and biome conditions permitted. The harnessing of natural energy flows was also highly geographydependent; waterwheels were inherently linked to river access and could only produce useful energy on rivers with sufficient and reliable flow. Therefore, the total number of suitable sites in a given region was constrained, which contrasted with windmills which were not as spatially fixed (i.e., they could be located to exploit wind wherever conditions and land availability permitted and were therefore not restricted only to the smaller number of suitable sites as waterwheels were) and were therefore available for use by a greater proportion of the population over larger areas (and consequently spread more rapidly).

18

N. KING AND A. JONES

Whale oil represented one of the first instances of a fuel which had a sufficiently high energy density to be conducive to transport over large distances to specific end users. The oil itself was a highly specialised product which could only be obtained, extracted and transported via a highly complex system (consisting of highly specialised whaling ships serviced by large-scale infrastructure), but this was emergent as a function of the high portability of the stable liquid product with a specialised application (lighting) for which there were few practical or economic substitutes. Fossil fuels represented a new energy paradigm not only because of their superlative energy content and (at the start of the Industrial Revolution) abundance, they were also stable stores of energy which could be used instantaneously as desired, and also offered a level of portability which no other energy source (save whale oil; but this was much more constrained in availability and application) had previously been able to match (though the portability varied between the different forms of fossil fuels). The combination of these factors combined to make the impact of fossil fuels more potent than any previous paradigm. However, the highly heterogenous global distribution of the deposits of these fuels became a constraint which generated a range of highly complex dynamics in the socio-political development and interaction of human societies. As consumption of coal started to increase rapidly in the early stages of the Industrial Revolution infrastructure (e.g., canals, railways, ports for coastal shipping) suited to transporting large quantities of highmass material emerged in different countries globally (initially in Britain), however the effectiveness of the modes of transport were constrained by the solid and bulky nature of the material. Coal gasification was a partial solution to making coal energy more amenable to distribution at local scale (through distribution in urban pipe networks), but it was the fluid nature of oil (combined with a greater energy density) which made its use inherently more flexible. Oil (and later natural gas) could be transported and transferred at a range of scales using pumps, pipelines and tanks, which made it amenable to integration into a wide range of industrial and domestic uses, and therefore underpinned the emergence of oil and gas as dominant (excepting some applications such as power generation for which bulk transport infrastructure for coal could be operated) [41]. Importantly oil and coal also had sufficient energy density such that they could be transported and used to power that transport which saw a significant increase in the range and speed of industrial vehicles, initially

1

INTRODUCTION

19

through the steam engine and subsequently with the internal combustion engine. The distribution of fossil fuel reserves in the modern world are a function of landscapes and biomes in remote geological timeframes. The Middle Eastern petroleum systems formed on the shallow continental shelf-margins of the Jurassic Era Tethys Sea, and much of the global coal resource originated in tropical forests located in equatorial regions of Carboniferous Era Pangea (which tectonic movements subsequently redistributed) [42, 43]; these carbon deposits supplied much of the fossil fuel energy which has driven industrialisation and the Great Acceleration but from the anthropogenic perspective, were randomly spatially distributed. One of the key outcomes of this was that nations and regions endowed (by chance) with fossil fuel resources (or were able to obtain them e.g., through having land or sea infrastructure to import energy resources) had a distinct advantage in industrialising; this was a key factor in the nations which achieved this early (e.g., Britain, USA, Germany). Later, nations and regions with large reserves (particularly of oil) have been able to leverage this to gain geopolitical influence or have been the target of action (geopolitical and/or military) by others to gain access to and control of such resources [44, 45]. In particular early adopters of such technology and energy use (such as Britain) were able to use this to their advantage and building on previous exploitation of energy sources— notably slaves—conquered and colonised countries to gain access to further resources. Electrification (enabled by fossil fuel use and later supplemented by renewables and nuclear) represents a further paradigm change in that it has ‘smoothed’ out the influence of geography and portability. This is because any energy users connected to electricity grids (though not locally-generated electricity) are effectively disconnected from any concerns over portability, as electricity is available instantaneously without the requirements to transfer and process fuels and materials, and geographical constraints to local resource availability is reduced by the long-distance transmission of electrical power. Fuels and energy flows still have to be accessed and converted to generate the power, but these functions are largely centralised in many areas of the world and are therefore largely not immediately apparent to the user. Portability manifests as the requirement to generate high voltages and minimise losses from transmission of the power (though again this is largely ‘invisible’ from the users’ perspective). The spread of electrical

20

N. KING AND A. JONES

grids to a large proportion of the world in recent decades has made energy (in a form which can be used for a large range of applications) available to a large number of users for which geography had previously been a major constraint [46]. Energy Return on Investment One key measure of the changing nature of human energy use over time is Energy Return on Investment (EROI, also Energy Returned on Energy Invested, EROEI). In its simplest form, this is a dimensionless ratio of the quantity of energy yielded by an energy source or technology to the quantity of energy required to obtain that energy or enable its usage; it is therefore a measure of net energy (yield minus costs).4 The energy sources used by humans through most of history have had positive but relatively low EROI values, for example humans in early societies had to use a portion of their limited supply of (endosomatic) energy to gather biomass to fuel fires, however the thermal energy yielded by the fire was significantly greater than this input. Additionally, the energy yielded could be leveraged to assist in obtaining further (endosomatic) energy such as by cooking of food to enhance nutrient gain. However, for most of human history the ratio of yield to cost was relatively low, and the general dearth of ‘surplus’ (or discretionary) energy was an overarching constraint on the rate of growth and change achievable by human societies. This paradigm of limited energy returns and general low EROI values (which persisted for the majority of human history) was finally transformed with the flourishing of fossil fuel use and the Industrial Revolution. This is because the energy yielded by fossil fuels (at the start of the industrial era; see below) was vastly greater (up to two orders of magnitude) than had to be invested to obtain the fuels, which is as a result of fossil fuels having a high energy density and being concentrated in reservoirs. This is a function of ancient biological (primarily photosynthesis and anaerobic breakdown) and geological processes (consolidation and transformation, i.e., coalification and catagenesis) collecting and concentrating energy bearing materials over very long periods of time, which

4 The EROI concept can include significant complexity in the form of consideration of the where the boundaries of energy systems are assumed to be (what level of energy input are considered before energy is yielded to the final user) along with the efficiency of energy conversions at different stages, the role of material embedded energy, etc. For the purposes of this subsection, only the simplest description of EROI is considered.

1

INTRODUCTION

21

have provided reserves of energy which are essentially ‘ready to use/free at point of use’ for humans. Although fossil fuels have to be extracted (e.g., through mining and drilling) and processed (e.g., through refining), these processes are minor (in terms of energy input) in comparison to that which would be required to generate synthetic equivalents ‘from scratch’. It is this gain from accumulation and concentration of energy through natural processes over geological time (akin to the trickle charging of a battery) which has afforded human civilisation very high EROI resources in the form of fossil fuels [47]. Portable and high EROI fossil fuels have underpinned the dramatic growth in human population and the physical extent of civilisation in the modern era (i.e., via the Great Acceleration) as a result of the large amount ‘discretionary energy’ that they have yielded. However, different human organisations and societies through time have always sought to obtain resources via the ‘low hanging fruit’ principle; those resources which are most readily accessible and easiest to utilise have been preferentially utilised first, and fossil fuels have been no exception to this. The downside of the ‘low hanging fruit’ principle is that as depletion of finite stocks of resources continues, the remaining reserves become progressively harder to access. For fossil fuels, this means greater energy investment being required in non-conventional exploration and production such as drilling deeper wells, extracting lower quality ores from mines, or transporting fuels from more distant locations. The result of this greater energy investment is that the EROI value of all fossil fuels has been and continues to experience an ongoing and irreversible decline. EROI estimates vary although the mean current global EROI for coal, nuclear and wind are comparable (at around 20:1 – meaning twenty times as much energy can be usefully used as is required to access it), with gas, geothermal and solar having lower EROI. However, hydroelectric EROI is significantly larger at over 75:1 [48] but of course far less portable as a fuel. More recent estimates of societal EROI (EROI estimated at the final energy stage where it is consumed such as petrol or electricity) have shown a dramatically lower EROI for fossil fuels. Final energy stage EROI for electricity generated from coal or gas could be as low as 6:1 [49] which is comparable to solar photovoltaic and lower than that for wind. Therefore, the historic EROI advantage of fossil fuels over other sources of energy appears to have come to an end.

22

N. KING AND A. JONES

Maintenance of complex societies requires a minimum EROI value to sustain the far-from-equilibrium thermodynamic conditions in which these societies exist (i.e., to counteract entropic effects and allow additional complexity to be built). This is why substantial growth in overall human society was only possible when fossil fuels were yielding consistently high EROI values. As fossil fuels undergo progressively greater depletion and more energetically intensive extraction (e.g., deep ocean drilling, hydraulic fracturing and tar sand mining) must be increasingly pursued globally, their EROI value will continue to drop. As noted above, the other energy sources brought into use at scale using fossil energy (renewables and nuclear) have different EROI characteristics; generally, the discretionary energy yield may be lower and more homogenous than that obtained historically with fossil fuels. Therefore, the EROI of the aggregated energy sources underpinning civilisation may inevitably reach a new equilibrium value [50–53]. Despite the fundamental importance of EROI as a means to identify energy sources and technologies which are able to provide energy surpluses, and a measure of their ability to underpin complexity, the importance of EROI may in the modern context be secondary to geography and portability considerations. This is because the spatial spread of portable energy to widely distributed users (not just those in proximity to resources) has been fundamental to the transformation of energy (and more general civilisational) paradigms. An energy resource with very high EROI but which is not amenable to being stored or portable (by transport or transmission) would have little capacity to be transformational. Similarly, some modern technological uses of energy-dense resources reduce their EROI at the point of use (e.g., use of natural gas for electricity generation and long distance transmission, which incurs losses at each stage), but the gains in transcending geography through portability and high utility value of the end product are the key to making this system useful. Complexity, Energy and the Superorganism A characteristic and general trend of human civilisation through time has been growing complexity. This has been subject to short term ‘noise’ and reversals as societies at regional scale have decreased in complexity, but the global trend has been a steady increase over the course of history. Complexity in this context is defined as the totality of socio-economic,

1

INTRODUCTION

23

technological and organisational structures (physical and social), relationships and strategies that are deployed by human societies to solve problems, which may include construction/maintenance of infrastructure, control of the distribution of wealth, and enforcement of laws [54, 55]. As access to energy has grown (and attained exponential characteristics in recent timeframes), and the number of human nodes and agents (e.g., individuals, institutions, trading organisations, and their interactions) comprising global civilisation has increased, autopoietic behaviour has emerged as networks (i.e., exchanging information, energy, materials), have grown and complexified concurrently. When viewed as a collective, human civilisation can be considered to have become (especially in recent timeframes) a self-sustaining and -organising structure that may be labelled as the ‘superorganism’ driven by an inexorable ‘growth imperative’ [30]. Applying this description, the superorganism is the aggregation of humans, technology and all the systems interlinking these agents, into a self-organising whole that resembles an organism in that it seeks to ‘ingest’ high-exergy resources and ‘metabolise’ them in the myriad industrial processes undertaken by civilisation, and to ‘excrete’ the resulting high-entropy wastes (e.g., carbon dioxide, chemical pollutants) as it drives towards continued growth. This system is not under the direct control of any agency or other control mechanism and is an emergent phenomenon of very large numbers of human and automated interactions at different scales and will likely therefore continue to maintain itself and grow. The ‘organism’ analogy can be taken further by highlighting that the global economy appears to have undergone a transition or reorganisation during the approximate midpoint of the Great Acceleration such that it achieves sublinear scaling of resource consumption (in line with Kleiber’s Law) relative to size, in a manner equivalent to living organisms beyond a certain size and mass threshold. The globalisation of economic activity may in this analogy be akin to the development of an organism’s circulatory system (i.e., development of globe-spanning and intricate transport networks), and indicates the tendency towards growth is strong and will likely continue in the absence of deliberate intervention and/or the collapse of supporting flows of energy and materials [13, 30, 56–61].

24

N. KING AND A. JONES

Current Global Energy Situation—The Global ‘ Energy Bind’ The preceding subsections describe the evolution of human energy use over time in terms of succession and interdependence, and also the overarching phenomena and patterns which have emerged from this long and complex sequence of events and trends. One key aspect which has emerged in recent timeframes is the increasing risk of future instabilities and interruptions to the current global energy paradigm as a result of a number of factors. This increasingly precarious situation arises from the very large and increasing energy requirements to maintain and grow complex civilisation at global scale, which includes increasing costs on maintenance where existing complexity becomes increasingly energetically expensive to sustain due to constant and cumulative entropic effects [62]. At the same time energy availability is experiencing complex, interdependent constraints. This situation is the global ‘energy bind’. Renewables and nuclear energy entered the human energy mix following a prolonged period of time during which fossil fuels dominated energy consumption at global scale. However, even after several decades of nuclear and renewable energy use (including periods of rapid technological advancement and recent dramatic growth in installed capacity) fossil fuels still dominate the global energy mix. This indicates that due to the relentless rises in overall energy use globally over the course of the Great Acceleration (driven by the compulsion and imperative of the growth of the ‘superorganism’), along with the effects of the Jevons Paradox, new energy sources are adding to the total but have not yet created the conditions for a robust transition away from fossil fuel dominance. As long as overall energy demand and use continues to grow at least as fast as new alternative energy capacity can be brought online, fundamental changes to the energy system are unlikely to occur [63–65]. Energy technologies and the characteristics of the energy they output also bring up significant challenges. A wide range of the materials and technological items which underpin the function of complex, hightechnology society rely on industrial process heat (defined as thermal energy used directly in the preparation or treatment of materials) as a key part of their manufacturing processes (notably ammonia, cement and steel, but also key processes such as oil refining, food processing and microchip manufacture [66, 67]) This is largely reliant on the particular characteristics of fossil fuels (affordability, availability, storability, high energy density and being ‘sacrificial’ in the heat generation process) which

1

INTRODUCTION

25

are difficult to replicate with other energy sources, due mainly to material and/or technological maturity constraints [68, 69]. This is therefore a major driver of global energy demand (comprising two thirds of industrial energy demand and one fifth of total global energy consumption) and a large singular contributor to total GHG emissions (making up the majority of industrial CO2 emissions) [70], and these needs also compete with processes which require fossil fuels as non-energetic feedstocks (notably plastics, of which 8.3 billion tonnes has been produced globally to date [71]). Global civilisation has built up over time (and most intensively in recent decades) into a spatially very large and complex state on which more than eight billion people rely continuously (directly or indirectly) for survival. The main energy source (fossil fuels) which has underpinned the emergence of these systems and structures was ideally suited given its (initially) high EROI, but after approximately two hundred years of continually growing extraction and use, global reservoirs are experiencing intensifying depletion. Alternative energy sources have emerged, but global energy demand has grown in conjunction with these innovations, and they also have fundamentally different characteristics in terms of EROI and ability to support certain energetic processes. The ‘energy bind’ as described above is therefore that maintenance of societal complexity is an imperative (in order to minimise the risk of large-scale disruption and potential collapse), but the energetic means to do this are becoming increasingly more difficult to attain. As such, collective humanity faces a bind in how to effectively manage this multi-faceted energy challenge. The following chapter undertakes a systemic analysis of how the global energy future may evolve.

References 1. Future Earth. (2021). Global Risks Perceptions Report 2021. https://fut ureearth.org/wp-content/uploads/2021/12/GlobalRisksPerceptionsRepor t2021.pdf. Accessed 8 Jan 2023. 2. Lawrence, M., Janzwood, S., & Homer-Dixon, T. (2022). What Is a Global Polycrisis? And How Is It Different from a Systemic Risk? https://cascadeinstitute.org/wp-content/uploads/2022/04/Whatis-a-global-polycrisis-v2.pdf. Accessed 8 Jan 2023.

26

N. KING AND A. JONES

3. Post Carbon Institute. (2023). Welcome to the Great Unraveling: Navigating the Polycrisis of Environmental and Social Breakdown. https://www.postca rbon.org/publications/welcome-to-the-great-unraveling/. Accessed 8 July 2023. 4. D’Odorico, P., Davis, K. F., Rosa, L., Carr, J. A., Chiarelli, D., Dell’Angelo, J., Gephart, J., MacDonald, G. K., Seekell, D. A., Suweis, S., & Rulli, M. C. (2018). The Global Food-Energy-Water Nexus. Review of Geophysics, 56(3), 456–531. 5. Simpson, G. B., & Jewitt, G. P. W. (2019, February). The Development of the Water-Energy-Food Nexus as a Framework for Achieving Resource Security: A Review. Frontiers in Environmental Science, 7 , 08. 6. Hogeboom, R. J., Borsje, B. W., Deribe, M. M., van der Meer, F. D., Mehvar, S., Meyer, M. A., Özerol, G., Hoekstra, A. Y., & Nelsom A. D. (2021, March 18). Resilience Meets the Water–Energy–Food Nexus: Mapping the Research Landscape. Frontiers in Environmental Science. 7. Natalini, D., Bravo, G., & Jones, A. W. (2019). Global Food Security and Food Riots—An Agent-Based Modelling Approach. Food Security, 11, 1153– 1173. 8. Li, Y., Zhang, Y., & Yang, N. (2010). Ecological Network Model Analysis of China’s Endosomatic and Exosomatic Societal Metabolism. Procedia Environmental Sciences, 2, 1400–1406. 9. Woods, J., Williams, A., Hughes, J. K., Black, M., & Murphy, R. (2010). Energy and the Food System. Philosophical Transactions of the Royal Society B, 365(1554), 2991–3006. 10. Herbert, T. D., Lawrence, K. T., Tzanova, A., Peterson, L. C., CaballeroGill, R., & Kelly, C. S. (2016). Late Miocene Global Cooling and the Rise of Modern Ecosystems. Nature Geoscience, 9, 843–847. 11. Gowlett, J. A. J. (2016). The Discovery of Fire by Humans: A Long and Convoluted Process. Philosophical Transactions of the Royal Society B, 371, 20150164. 12. Wrigley, E. A. (2013). Energy and the English Industrial Revolution. Philosophical Transactions of the Royal Society A, 371, 20110568. 13. Hagens, N. J. (2020). Economics for the Future—Beyond the Superorganism. Ecological Economics, 169, 106520. 14. Gross, M. (2013). The Paradoxical Evolution of Agriculture. Current Biology, 23(16), R667–R670. 15. Larsen, C. S. (2006). The Agricultural Revolution as Environmental Catastrophe: Implications for Health and Lifestyle in the Holocene. Quaternary International, 150(1), 12–20. 16. Kunz, H., Hagens, N. J., & Balogh, S. B. (2014). The Influence of Output Variability from Renewable Electricity Generation on Net Energy Calculations. Energies, 7 , 150–172.

1

INTRODUCTION

27

17. Gignoux, C. R., Henn, B. M., & Mountain, J. L. (2011). Rapid, Global Demographic Expansions After the Origins of Agriculture. Proceedings of the National Academy of Sciences, 108(15), 6044–6049. 18. Alberta Culture and Tourism. (2023). Animal Power in Prehistoric Times. Available online: http://www.history.alberta.ca/energyheritage/energy/ human-and-animal-power/animal-power-in-prehistoric-times/default.aspx. Accessed 15 Jan 2023. 19. Pain, S. (2017). Power Through the Ages. Nature, 551, S134–S137. 20. Low Tech Magazine. (2009). Wind Powered Factories: History (and Future) of Industrial Windmills. Available online: https://www.lowtechmagazine. com/2009/10/history-of-industrial-windmills.html. Accessed 17 Jan 2023. 21. Guardian. (2023). The First Great Energy Transition: How Humanity Gave Up Whaling. Available online: https://www.theguardian.com/env ironment/2023/mar/11/the-first-great-energy-transition-how-humanitygave-up-whaling. Accessed 14 Mar 2023. 22. Perissi, I., & Bardi, U. (2021). The Empty Sea—The Future of the Blue Economy. Springer. 23. Slave Voyages. (2023). Trans-Atlantic Slave Trade—Estimates. Available online: https://www.slavevoyages.org/assessment/estimates. Accessed 16 Jan 2023. 24. Hughes, D. M. (2017). Energy Without Conscience: Oil, Climate Change, and Complicity (Chapter 1—Plantations Slaves, the First Fuel). Duke University Press, Duke University. 25. Johnstone, P., & McLeish, C. (2020). World Wars and the Age of Oil: Exploring Directionality in Deep Energy Transitions. Energy Research & Social Science, 69, 101732. 26. Science Daily. (2004). Humans May Surpass Other Natural Forces as Earth Movers. Available online: https://www.sciencedaily.com/releases/ 2004/07/040709083319.htm. Accessed 18 Jan 2023. 27. Cooper, A. H., Brown, T. J., Price, S. J., Ford, J. R., & Waters, C. N. (2018). Humans Are the Most Significant Global Geomorphological Driving Force of the 21st Century. The Anthropocene Review, 5(3), 222–229. 28. Erisman, J. W., Sutton, M. A., Galloway, J., Klimont, Z., & Winiwarter, W. (2008). How a Century of Ammonia Synthesis Changed the World. Nature Geoscience, 1, 636–639. 29. Douet, J. (2020). The International Committee for the Conservation of the Industrial Heritage—Thematic Study—The Heritage of the Oil Industry. Available online: https://ticcih.org/wp-content/uploads/2020/ 07/Oil-industry-thematic-report.pdf. Accessed 18 Jan 2023. 30. Murphy, T. W. (2021). Energy and Human Ambitions on a Finite Planet— Assessing and Adapting to Planetary Limits. University of California.

28

N. KING AND A. JONES

31. BP. (2022). BP Statistical Review of World Energy 2022 (71st ed.). Available online: https://www.bp.com/content/dam/bp/business-sites/en/global/ corporate/pdfs/energy-economics/statistical-review/bp-stats-review-2022full-report.pdf. Accessed 18 Jan 2023. 32. International Hydropower Association. (2022). A Brief History of Hydropower. Available online: https://www.hydropower.org/iha/discoverhistory-of-hydropower. Accessed 22 Jan 2023. 33. International Renewable Energy Agency. (2022). Wind Energy. Available online: https://www.irena.org/Energy-Transition/Technology/Windenergy. Accessed 26 Jan 2023. 34. NS Energy. (2020). Profiling the Top Geothermal Power Producing Countries in the World. Available online: https://www.nsenergybusiness.com/ features/top-geothermal-power-producing-countries/#. Accessed 31 July 2023. 35. International Energy Agency. (2022). Solar PV . Available online: https:// www.iea.org/reports/solar-pv. Accessed 26 Jan 2023. 36. World Nuclear Association. (2020). Outline History of Nuclear Energy. Available online: https://world-nuclear.org/information-library/currentand-future-generation/outline-history-of-nuclear-energy.aspx. Accessed 26 Jan 2023. 37. Millward-Hopkins, J., Steinberger, J. K., & Rao, N. D. (2020). Providing Decent Living with Minimum Energy: A Global Scenario. Global Environmental Change, 65, 102168. 38. Syvitski, J., Waters, C. N., Day, J., Milliman, J. D., Summerhayes, C., Steffen, W., Zalasiewicz, J., Cearreta, A., Gałuszka, A., Hajdas, I., Head, M. J., Leinfelder, R., McNeill, J. R., Poirier, C., Rose, N. L., Shotyk, W., Wagreich, M., & Williams, M. (2020). Extraordinary Human Energy Consumption and Resultant Geological Impacts Beginning Around 1950 CE Initiated the Proposed Anthropocene Epoch. Communications Earth & Environment, 1(1), 32. 39. Cavanagh, M., Ben-Yosef, E., & Langgut, D. (2022). Fuel Exploitation and Environmental Degradation at the Iron Age Copper Industry of the Timna Valley, Southern Israel. Scientific Reports, 12, 15434. 40. Steffen, W., Broadgate, W., Deutsch, L., Gaffney, O., & Ludwig, C. (2015). The Trajectory of the Anthropocene: The Great Acceleration. Anthropocene Review, 2, 81–98. 41. Melsted, O., & Pallua, I. (2018). The Historical Transition from Coal to Hydrocarbons: Previous Explanations and the Need for an Integrative Perspective. Canadian Journal of History, 53(3), 395–422. 42. Al-Husseini, M. I. (1997). Jurassic Sequence Stratigraphy of the Western and Southern Arabian Gulf. GeoArabia, 2(4), 361–382.

1

INTRODUCTION

29

43. Feulner, G. (2017). Formation of Most of Our Coal Brought Earth Close to Global Glaciation. Proceedings of the National Academy of Sciences, 114(43), 11333–11337. 44. National Bureau of Economic Research. (2014). Coal and the European Industrial Revolution. Available online: https://www.nber.org/system/ files/working_papers/w19802/revisions/w19802.rev0.pdf. Accessed 21 July 2023. 45. Salimi, M., & Amidpour, M. (2022). The Impact of Energy Transition on the Geopolitical Importance of Oil-Exporting Countries. World, 3(3), 607– 618. 46. Kalt, G., Thunshirn, P., & Haberl, H. (2021). A Global Inventory of Electricity Infrastructures from 1980 to 2017: Country-Level Data on Power Plants Grids and Transformers. Data in Brief, 38, 107351. 47. Schramski, J. R., Gattie, D. K., & Brown, J. H. (2015). Human Domination of the Biosphere: Rapid Discharge of the Earth-Space Battery Foretells the Future of Humankind. Proceedings of the National Academy of Sciences, 112(31), 9511–9517. 48. Hall, C. A. S., Lambert, J. G., & Balogh, S. B. (2014). EROI of Different Fuels and the Implications for Society. Energy Policy, 64, 141–152. 49. Brockway, P. E., Owen, A., Brand-Correa, L. I., & Hardt, L. (2019). Estimation of Global Final-Stage Energy-Return-on-Investment for Fossil Fuels with Comparison to Renewable Energy Sources. Nature Energy, 4, 612–621. 50. Murphy, D. J., Raugei, M., Carbajales-Dale, M., & Estrada, B. R. (2022). Energy Return on Investment of Major Energy Carriers: Review and Harmonization. Sustainability, 14(12), 7098. 51. Murphy, D. J. (2014). The Implications of the Declining Energy Return on Investment of Oil Production. Philosophical Transactions of the Royal Society A, 372, 20130126. 52. Brandt, A. R. (2017). How Does Energy Resource Depletion Affect Prosperity? Mathematics of a Minimum Energy Return on Investment (EROI). Biophysical Economics and Resource Quality, 2(1), 1–12. 53. Hall, C. A. S., Balogh, S., & Murphy, D. J. R. (2009). What Is the Minimum EROI That a Sustainable Society Must Have? Energies, 2(1), 25–47. 54. Tainter, J. A. (1988). The Collapse of Complex Societies. University of Cambridge Press. 55. Tainter, J. A., & Taylor, T. G. (2013). Complexity, Problem-Solving, Sustainability and Resilience. Building Research and Information, 42(2), 168–181. 56. Martin, J. R. (2005). Complex Systems and Exosomatic Energy Metabolism of Human Societies [Doctoral dissertation for the Programme in Environmental Sciences (Ecological Economics and Environmental Management)]. Universitat Autònoma de Barcelona, Spain.

30

N. KING AND A. JONES

57. Wissner-Gross, A. D., & Freer, C. E. (2013). Causal Entropic Forces. Physical Research Letters, 110, 168702. 58. Jarvis, A., & King, C. (2020). Energetic Regimes of the Global Economy— Past, Present and Future. Earth System Dynamics, Preprint esd-2020-59. 59. King, C. W. (2022). How Does Global Energy Consumption Scale with GDP and Mass? A Biophysical Perspective. Available online: https://www.resili ence.org/stories/2022-10-12/how-does-global-energy-consumption-scalewith-gdp-and-mass-a-biophysical-perspective/. Accessed 26 Oct 2022. 60. King, C. W. (2022). Interdependence of Growth, Structure, Size and Resource Consumption During an Economic Growth Cycle. Biophysical Economics and Sustainability, 7 , 1. 61. New England Complex Systems Institute. (2022). Complexity Rising: From Human Beings to Human Civilization, a Complexity Profile. Available online: https://necsi.edu/complexity-rising-from-human-beings-to-humancivilization-a-complexity-profile. Accessed 01 Feb 2023. 62. Garrett, T. J., Grasselli, M. R., & Keen, S. (2022). Lotka’s Wheel and the Long Arm of History: How Does the Distant Past Determine Today’s Global Rate of Energy Consumption? Earth System Dynamics, 13(2), 1021–1028. 63. York, R., & Bell, S. E. (2019). Energy Transitions or Additions? Why a Transition from Fossil Fuels Requires More Than the Growth of Renewable Energy. Energy Research & Social Science, 51, 40–43. 64. Garrett, T. J., Grasselli, M., & Keen, S. (2020). Past World Economic Production Constrains Current Energy Demands: Persistent Scaling with Implications for Economic Growth and Climate Change Mitigation. PLoS ONE, 15(8), e0237672. 65. Giampietro, M., & Mayumi, K. (2018). Unraveling the Complexity of the Jevons Paradox: The Link Between Innovation, Efficiency, and Sustainability. Frontiers in Energy Research, 6, 26. 66. Smil, V. (2022). The Modern World Can’t Exist Without These Four Ingredients. They All Require Fossil Fuels. Available online: https://time.com/617 5734/reliance-on-fossil-fuels/. Accessed 1 Aug 2023. 67. Williams, E. D., Ayres, R. U., & Heller, M. (2002). The 1.7 Kilogram Microchip: Energy and Material Use in the Production of Semiconductor Devices. Environmental Science and Technology, 36, 24, 5504–5510. 68. World Nuclear Association. (2020). Generation IV Nuclear Reactors. Available online: https://world-nuclear.org/information-library/nuclear-fuelcycle/nuclear-power-reactors/generation-iv-nuclear-reactors.aspx. Accessed 10 Jan 2023. 69. Yu, Z. (2022). Grand Challenges in Heat Decarbonisation. Frontiers in Thermal Engineering, 2, 940072.

1

INTRODUCTION

31

70. International Energy Agency. (2018). Clean and Efficient Heat for Industry. Available online: https://www.iea.org/commentaries/clean-andefficient-heat-for-industry. Accessed 10 Jan 2023. 71. Geyer, R., Jambeck, J. R., & Law, K. L. (2017). Production, Use, and Fate of All Plastics Ever Made. Science Advances, 3(7), e1700782.

CHAPTER 2

The Trifurcation of Energy Futures

Abstract Extrapolation of the current ‘energy bind’ from a systems perspective indicates that the global paradigm will likely follow one of three pathways; this trifurcation leads to branches with fundamentally distinct features. The first (Fossil-Seneca Branch) represents ‘BusinessAs-Usual’ in terms of continued fossil fuel (and nuclear fission) use; the combination of climatic changes combined with resource depletion gives rise to the risk of a Seneca collapse. The second (Continued Growth Branch) involves a transition to an energy system underpinned by nuclear fusion and/or renewables based technologies which could allow openended energy use growth. Both technologies have technical and resource constraints, but even if open-ended growth could be achieved it would likely result in multiple negative effects for global society. The third (Stabilisation Branch) involves stabilisation of energy use using renewables technologies; implementing this would involve significant systemic, resource and other challenges but would likely provide the optimal outcome for global civilisation. Keywords Energy futures · Trifurcation · Climate change · Depletion · Fossil fuels · Nuclear fission · Seneca Effect · Continued growth · Renewables · Nuclear fusion · Technosphere · Limits to growth · Degrowth

© The Author(s), under exclusive license to Springer Nature Switzerland AG 2023 N. King and A. Jones, Future Energy Options from a Systems Perspective, https://doi.org/10.1007/978-3-031-46448-5_2

33

34

N. KING AND A. JONES

This chapter considers the ‘energy bind’ which global civilisation faces (as described in the preceding chapter) from a systems perspective. This assessment extrapolates the current energy paradigm along different feasible pathways (‘energy futures’), in order to assess the high-level and long-term features and ‘modes’ of behaviour of these possible futures. This considers the features and characteristics of these pathways along with the potential consequences and outcomes of the trifurcation and subsequent evolution of the global energy system along these different pathways. The feasibility and desirability of these different modes (and potentially hybrids of these) is also assessed. The current energy paradigm is divided into three different ‘Branches ’, based on the current global paradigm trifurcating1 in future along three distinct, fundamentally different (in terms of underpinning technology, systemic features at global scale, and interaction with the wider Earth System) pathways. Their evolution over time is a function of their features and starting conditions, and in line with the chaotic behaviour of complex systems, may lead to very different final outcomes. These three pathways are: • The Fossil-Seneca Branch which sees society continue along a business-as-usual trajectory and remain dominated by the use of fossil fuels. • The Continued Growth Branch which sees society become reliant on renewable technologies whilst still pursuing open ended growth. • The Stabilisation Branch which sees society become reliant on renewable technologies but achieves a stabilisation in global energy demand while still delivering human prosperity.

The Fossil-Seneca Branch The Fossil-Seneca Branch of the trifurcating energy future represents ‘Business-As-Usual ’ (BAU) in terms of the global energy system continuing to be overwhelmingly based on fossil fuel energy, with year-on-year expansion in overall energy use. This will lead to fossil fuel consumption continuing to

1 Note that this trifurcation is distinct from (though indirectly related to) the ‘energy trilemma’ (security, equity/affordability, and sustainability).

2

THE TRIFURCATION OF ENERGY FUTURES

35

rise even where other energy sources (particularly renewables) are also undergoing capacity expansions, leading to continued growth in global greenhouse gas emissions contributed to ever increasing impacts from climate change. Climate Change and Depletion The flourishing of the extraction and use of fossil fuels at scale has represented a new paradigm in overall human energy use, and the innovations and transformations enabled by the fossil fuel transition have come to dominate virtually all aspects of modern civilisation. One particularly significant aspect is that the carbon contained in the fossil deposits was derived (via photosynthetic biomass accumulation and geological transformation) from the atmospheres of remote past geological timeframes. This is in contrast to modern biomass, which is derived from the contemporary atmosphere and is cycled by photosynthesis and combustion. The fossil carbon had been immobilised in the lithosphere and had not interacted with the global carbon cycle for extended periods of time (approximately tens to hundreds of millions of years), but combustion of these deposits (as fossil fuel) has released this ancient carbon into the contemporary atmosphere. This has resulted in the growing perturbation of the modern carbon cycle through the rapid and large-scale addition of this ‘new’ carbon [1]. The carbon emitted by anthropogenic activity, of which carbon dioxide (CO2 ) is the most volumetrically important, gives rise to a contemporary atmospheric concentration value of approximately 420 ppm.2 Shorterlived climate pollutants including methane (CH4 ) and nitrous oxides (NOx ) are also significant [2]. These emissions have resulted in a generalised and accelerating warming trend at the Earth’s surface which has increased in proportion to accumulating greenhouse gas (GHG) concentrations (though the climatic response to GHG is nonlinear; see below [3]). The total accumulated quantity of anthropogenic CO2 is approximately one and a half trillion tonnes, and by the early 2020s the annual global emissions of CO2 (from fossil fuel combustion and land use change3 ) was approximately 40 billion tonnes [4]. This forcing leads to a general increase in the energy budget of the Earth System (resulting 2 At the time of writing. 3 Note that the GHG from the Industrial Revolution onward accumulated in addition

to GHG from earlier human activity, namely deforestation and land use change undertaken

36

N. KING AND A. JONES

from additional atmospheric radiative forcing), which has resulted in the average global surface temperature increasing over the last two hundred years by approximately 1–1.2 °C (or approximately 1.5 °C if the surface equivalent potential temperature metric, which considers humidity and latent energy too, is applied), giving rise to temperatures which have not occurred on Earth for the last approximate hundred thousand years [5–8]. Further increases in global temperature are projected due to the ongoing and growing nature of global anthropogenic GHG emissions, which have the potential to cross critical thresholds on near-term timescales (1.5 °C during the 2030s and 2 °C during the 2040–2050s, in the worst case scenarios) [9]. This is an occurrence which has the potential to trigger enhancing feedback loops and consequently tipping points in different parts of the Earth System, which could cause climate changes to become self-amplifying, potentially leading to non-linear further increases in global temperature [10, 11]. The feedback loops and interactions the climate system is prone to are of both natural (>40 biotic and abiotic natural feedback mechanisms have been identified including increases in wildfire occurrence and changes in global ice cover and corresponding albedo) and anthropogenic (the number of these mechanisms is yet to be estimated, but many are likely to be novel in nature such as global-scale increases in the use of fossil-fuel powered air conditioning as ambient temperatures rise) origins, which operate on a range of different timescales (including mechanisms with decade-scale delays). These feedback mechanisms generate the potential for very significant average global temperature rises including runaway warming (up to approximately 10 °C) which in many cases are not accounted for sufficiently in existing climate models nor in the conservative range normally considered (in international governance discourse) [12–15]. Another significant effect could result from the generation of anthropogenic aerosols (from burning of coal and biomass, and use of high-sulphur fuels in shipping for example), which may have provided a long term cooling effect which has offset past climate forcing (though this may have varied spatially). The general warming may have exceeded the cooling effect after a certain point in time, and recent reductions in aerosol production (as overall coal combustion and the sulphur content of bunker oil has

by ancient societies, and these activities continue to contribute carbon to the atmosphere in parallel to that contributed by fossil carbon combustion.

2

THE TRIFURCATION OF ENERGY FUTURES

37

decreased) may have enhanced the net-warming effect and also generated regional ‘termination shocks’ [16–19]. An average rise of global temperatures of 1.5 °C above preindustrial temperatures is the internationally agreed goal for emissions reductions because this has been identified as the lower threshold for minimising the risk of the triggering tipping points and causing the runaway climatic effects described above. The Intergovernmental Panel on Climate Change (IPCC) has compiled approximately four hundred scenarios for limiting rises to 1.5 °C, but assessment of these has identified that only a small proportion actually provide the means to avoid significant overshoot of this limit. An even smaller proportion of these scenarios make realistic assumptions about the mitigation approaches and technologies that could be employed in achieving the goal; the implication is that there a high risk of climatic overshoot, and the consequences of this occurring could severely challenge the capabilities of global society to mitigate the root causes (potentially leading to a ‘doom loop’) [20, 21]. This anthropogenically-driven climate change likely represents a perturbation to the Earth System on a magnitude equivalent to major natural disruptive events in Earth’s history, in terms of quantity and rate of release of GHGs [22]. Increases in the frequency and intensity of climate events (such as droughts and storms) are likely to impact a range of different ecosystems and groups of organisms (for example tropical forests and arthropods), and contribute to extinctions [23, 24] given that previous mass extinction events were in the majority of cases (though not all; extra-terrestrial events may have contributed to the occurrence of some extinctions [25, 26]) due to rapid global scale climatic changes [27]. In addition to impacts on natural systems, there is a growing base of evidence that the consequences of such a severe and rapid perturbation of the climatic system are likely to be manifold and serious for human civilisation as well, given the complex and close coupling of the climatic system to multiple human systems and the natural systems on which civilisation relies. One of the primary risks to human societies at global scale arises from the potential severe and widespread disruptions to global agriculture. The growth of human populations to contemporary proportions has depended on the relatively stable climatic conditions of the Holocene Epoch. These predictable conditions underpinned reliable and increasing agricultural yields year on year over extended periods. The potential shift of climatic

38

N. KING AND A. JONES

conditions (either at global scale, or within the regions in which massscale industrial agricultural is undertaken most intensively) to a more variable and unpredictable state, in which local and regional climatic and other conditions may be prone to short term extreme variation, would be far less conducive to the continuation of monoculture, crop-based agriculture as it has been carried out for the last ten thousand years and at very large scale in the last two hundred [28–35]. In addition to the direct climatic impacts on the function of agriculture itself, climatic effects also have the potential to impact the secondary systems which support the global food system, namely the globe-spanning infrastructure and supply chains which transport and distribute bulk quantities of food produce from source to point of use or consumption [36–39]. The provision of reliable food supplies is fundamental to the survival of large-scale human populations (particularly in the growing urban regions of the world where food self-sufficiency is low or nonexistent) and therefore the stability of complex societies [7]. Climate change presents a real and growing threat to this. The increase in extreme weather events resulting from climate change also has the potential to generate direct and potentially catastrophic impacts on other forms of critical infrastructure globally, including residential and other buildings, communications assets, and power generation and distribution systems. This infrastructure damage may arise from increasing incidence of extreme wind, rain, drought or fire events, causing economic losses and loss of human life, and acting as a general destabilisation mechanism which insurance and other compensatory systems will increasingly be unable to manage [40–42]. Other key threats to complex societies arising from anthropogenic climate perturbation include the likelihood that significant areas of the Earth (particularly in heavily populated equatorial and sub-tropical regions) may progressively transition out of the ‘human climate niche’ in which ambient heat and humidity conditions are such that humans cannot physically and economically function and survive [43–45]. There is also potential for environmental and other systems under pre-existing stress (e.g., from chemical pollution) to be subject to tipping points upon the introduction of new stresses or discrete extreme events (i.e., increased heat), leading to rapid regime shift or even collapse [46–48]. The accumulating effects of climate change (particularly the risk of passing tipping points) constitutes a full-scale global emergency [49, 50]. However, climate change is not the only source of risk arising from fossil

2

THE TRIFURCATION OF ENERGY FUTURES

39

fuel dominance; depletion of fossil fuels (and other natural resources) constitutes an additional threat. The biotic origin of fossil fuels from the distant geologic past means that they constitute a stock of energy which accumulated very slowly over geological timescales and will not be replenished on human timescales, so they have been in a state in constant diminishment since large-scale extraction began. ‘Peak oil’ describes the point in time at which the rate of oil production (applicable to an individual reservoir, a country or region, or to the global scale) reaches its maximum value, and thereafter goes into an irreversible decline due to the reservoir in question becoming progressively depleted. Due to the complexities of assessing the extent of hydrocarbon reserves and resources, determining when global peak oil has or will occur has produced varied conclusions. There is however evidence that this occurred for ‘conventional oil’ (i.e., that extracted through standard methods) during approximately the mid-2000s, but global hydrocarbon production continued to expand beyond that point in time due to innovations in and widespread expansion of ‘unconventional oil’ extraction (e.g., fracking, the large-scale expansion of which was enabled in part by economic conditions following the late-2000s Global Financial Crisis). This progression follows the ‘low hanging fruit’ principle and as previously described, is a large part of the driver of declining EROI. From a depletion standpoint this is also only a temporary solution as ‘unconventional oil’ will be subject to the same depletion dynamics as previously experienced by ‘conventional oil’, and this may occur on much more rapid timescales [51–54]. At the very least, the increased energy input required to extract unconventional oil means that this historically cheap source of energy will become increasingly more expensive over the coming decades. This increase in cost will see large portions of the global economy dedicated to ongoing extraction and production of energy (as opposed to its ‘discretionary’ consumption) resulting in the slowing down of other sectors even if the overall economy can continue to grow. A move towards more expensive energy through continued fossil fuel use will also see a continued set of challenges associated with the geopolitics of resource extraction and certain geographies will remain unstable as countries and companies compete to access remaining fossil fuel reserves. The ongoing and accelerating drawdown of these fossil fuel reserves (of all types) has the potential to undermine the globe-spanning complex system (the superorganism) which is currently structurally reliant on

40

N. KING AND A. JONES

fossil fuel consumption (either directly to fuel systems such as transport networks, or indirectly as enabling inputs to systems such as agriculture) [44, 55, 56]. The ‘peak’ concept is applicable to all stock-based resources, so in addition to ‘peak oil’ global society potentially faces the effects of synchronous peaking of multiple key resources [57]. This depletion of global resources (with fossil fuels as one of the most fundamentally important) has the potential to drive a range of outcomes for societies, but several credible scenarios involve the potential for global instability and conflict as economic, geopolitical and biophysical stressors increase in proportion to and in lockstep with growing scarcities [58]. Characteristics and Status of Nuclear Fission Technology Nuclear technology emerged from the fossil-fuelled superorganism, and allowed humans to appropriate energy stored with a medium completely different to any used previously. Technological systems for initiating, sustaining and extracting energy from nuclear reactions under controlled conditions have only been developed to a level of technological maturity for nuclear fission, which involves the ‘splitting’ of heavy nuclei present in nuclear fuel (primarily uranium) by the addition of neutrons, forming two or more lighter nuclei and releasing energetic neutrons and gamma photons. Nuclear fuels4 have a far higher energy density (per unit mass or volume) than chemical fuels (approximately a million times greater [30]) and so large quantities of energy can be released from relatively small quantities of fuel. Due to the large total quantity of energy present within the stock of uranium present on Earth combined with its status as a mature and firm (non-intermittent) power source, nuclear technology (meaning fission-based systems; potential fusion systems are discussed separately in the Continued Growth Branch) has been suggested as a prime candidate to substantially displace fossil fuels [59], and thereby avert the climate change and depletion issues described above. This section assesses the history and (technical and economic) characteristics of nuclear fission technologies in the context of this Branch, to assess whether they may have the capacity to fundamentally alter it. In the initial stages of the development of nuclear fission technology following WWII, great transformational potential was anticipated; this

4 This applies to fuel used in both fission and fusion processes.

2

THE TRIFURCATION OF ENERGY FUTURES

41

was most famously captured in the 1954 statement “It is not too much to expect that our children will enjoy in their homes electrical energy too cheap to meter…” [60]. However, despite these optimistic early projections that nuclear technology would come to dominate global energy (or at least electricity) production, nuclear power was at the time of writing a relatively minor contributor to total global energy supplies; it contributed approximately 10% of global electricity production in 2021 and approximately 4% of global primary energy in 2019. In addition to these relatively minor contemporary contributions to the global energy mix, its proportional contribution to global energy output has experienced a 40% reduction relative to its 1990s peak [61, 62]. The fundamental dynamic of global nuclear fission capacity is that technological development and construction of capacity grew rapidly during the approximate period 1950s–1970s, but development and construction declined significantly after this period, and overall output from the global fleet of fission reactors subsequently started to decline (from the 1990s onwards) as they started to reach the end of their operational lifetimes. The factors and reasons underlying this initial rapid growth followed by prolonged decline are complex, but a central and pivotal factor is that the construction and generation costs of fission technologies have been continuously rising (by approximately 20% annually in the US) over recent decades [63, 64]. These cost rises are in turn attributable to a multitude of causes, but one key cross-cutting factor is that the ‘soft costs’ associated with the capital expenditure (capex) required to deliver new large nuclear plants worldwide has risen in the majority of locations. This is primarily due to increasingly demanding standards (as required by various national regulators to meet public acceptability criteria) of engineering design (e.g., for the reactor containment structures), along with purchasing, planning, scheduling and estimating/cost control. A number of inter-related factors which have contributed to the general rising costs and complexities of nuclear power (and therefore its deficient performance relative to early expectations) include: the impacts on public acceptability and perception as a result of major nuclear accidents, pollution incidents and military associations [65–69]; technical factors and public perception challenges associated with the rising demands for undertaking reactor decommissioning [70]; technological feasibility and long term capex and energy investment requirements; and siting and inter-generational equity challenges associated with radioactive waste

42

N. KING AND A. JONES

management, noting that mined geological repositories (the majorityaccepted approach for final radioactive waste management globally) are near-operational in a few nations, but facing ongoing challenges and uncertainties in many others [71–73]. An additional legacy factor which applies to the majority of the global nuclear industry and which has stymied its prospects in complex ways is technological lock in. The dominant fission technologies for civil power generation are subject to path dependency largely as a result of historical events combined with economic, intellectual property, national interest and other factors. This has resulted in light water reactors (LWRs) (and primarily the Pressurised Water Reactor, PWR) coming to numerically dominate the global reactor fleet, although a number of reason exist why these reactor types may not be the optimal technology for large civil power generation [74–76]. Related to this path dependency are rising costs associated with the fuel cycle; LWRs have a relatively inefficient fuel cycle in which a comparatively wasteful (and hazard-generating) ‘once-through’ use of uranium fuel has become in most nations more economical than recovery of unused fuel via reprocessing [77]. A factor which interlinks with many of those described above (and in particular the relatively inefficient use of uranium fuel by LWRs) and which is central to the meeting future energy demand, is the potential depletion of global uranium reserves [78, 79]. Fuels derived from mineral ores sourced from the Earth’s crust will always be subject to ‘hard’ total availability limits, and their extraction will also be subject to the ‘low hanging fruit’ and diminishing returns principles, and uranium will not be an exception. Although fuel costs have comprised a minor aspect of nuclear operational costs historically (at least relative to construction capex), current and future constraints to the availability of cheap uranium is likely to contribute to future rises in the operating costs of nuclear plants [80, 81] (in addition to the other rising costs, as described above) and to limit the overall potential for substantial expansion of global capacity [82]. An alternative terrestrial reserve of uranium is the global ocean, which contains the element in aqueous form. Although present at low concentrations (approximately 3 µg/l) this reserve is overall of much greater magnitude than crustal ore deposits (due to the overall volume of the ocean; 4 billion vs. almost 20 million tonnes). Accessing this much larger reserve of uranium has been suggested as fundamental to expanding

2

THE TRIFURCATION OF ENERGY FUTURES

43

nuclear fission to overall outputs competitive with fossil fuels, but extraction of this uranium presents significant technical challenges due to the high entropy state of its dissolved form. Although progress has been made in adsorbents which could selectively extract uranium from seawater effectively, fundamental practicality challenges around the scale of the infrastructure and energy input required (which would impact the EROI of the whole system) remain unresolved [83–86]. Fissile plutonium separated from spent nuclear fuel by reprocessing has been stockpiled in several nations and could be utilised by LWRs in the form of mixed oxide fuel; the global infrastructure to manufacture this form of nuclear fuel is however very limited, would present significant technical challenges and would in any case extend the globally available fuel stock by only an insignificant amount [87]. An emergent outcome of the confluence of increasing concerns about the growing prevalence and impacts of climate change, the intractability of nuclear cost rises (exemplified by delivery difficulties in contemporary large nuclear projects such as Flamanaville-3 and Vogtle-3/-4 [88, 89]), the technological lock-in which has generated path dependencies and constraints on technological flexibility, and the potential for uranium availability constraints, is interest and investment in development of new5 nuclear fission technologies, centred on ‘Generation IV’ concepts and Small- and Advanced-Modular Reactors (SMRs and AMRs, respectively). This technological development is a subset of the wider resurgence (‘renaissance’) in nuclear capacity construction, albeit on a far smaller scale than in the early stages of nuclear technology, in the first decades of the twenty-first century. This offers the possibility of disrupting a number of the collective drawbacks resulting from the global reactor fleet being LWR-dominated (noting that many of the SMR concepts use LWR technology; see below). The Generation IV technologies [90] (noting that some specific conceptual technologies are also described by their developers as ‘Generation V’ [91]) include reactor types which may be able to operate with fuel

5 A number of the concepts under development are modern successors of concepts proposed during the early phases of nuclear technology, the development which was suspended as LWRs reached maturity, gained economies of scale, and hence emerged as dominant.

44

N. KING AND A. JONES

cycles which utilise existing (LWR) spent nuclear fuel, thorium-based6 fuels [92], and to ‘breed’ additional fuel from non-fissile materials as they operate (noting that these reactor types have been proposed since the inception of nuclear fission technology and have been pursued through extensive research and development (R&D) programmes internationally, but these efforts have not successfully produced a technically or commercially viable system to date [93]). These technologies therefore offer the possibility of transforming the economics and efficiency of the nuclear fuel cycle, whilst improving other aspects such as efficiency, safety and load factors. SMRs/AMRs technologies have been proposed/developed which utilise different combinations of existing (primarily LWR) and Gen IV technologies,7 but with sizing and modularity which offer the possibility of more rapid and simplified (i.e., with many factory-built components which reduce construction-site burdens) construction, thereby addressing the origins of many of the fundamental cost challenges to which nuclear has recently been subject [94, 95]. A number of the Gen IV and/or SMR/AMR concepts also offer the potential to cater for applications which existing nuclear technologies cannot, namely process and district heating (to provide non-fossil fuel industrial process heat for bulk chemical and hydrogen production, and eliminating the need for gas fired domestic boilers, respectively) [96–98]. There are however a number of downsides, challenges and constraints to the transformative potential of Gen IV and SMR/AMR technologies. Firstly, the majority of these systems remain at low Technology Readiness Level (TRL; this is a method to estimate the maturity and deployability of technologies) [99], so may not have been demonstrated as safe and reliable at full operational scale and are likely have a large number of engineering challenges requiring resolution. Therefore, to develop one or more of these technologies to a level of maturity such that they could be licensed for commercial operations in multiple locations, and to then deploy these technologies such that they substantially replace/displace the

6 Thorium is significantly more abundant in the Earth’s crust then uranium, which could therefore alleviate (but not eliminate) depletion challenges. 7 Generally, a Gen IV reactor refers to a ‘full size’ reactor of this type (which may utilise non-conventional fuel such as thorium and/or undertake fuel breeding), and AMR as one of these systems built at smaller scale, whilst SMR refers to a smaller scale reactor using existing (primarily LWR) technologies. These categories are not fixed or formally defined, however.

2

THE TRIFURCATION OF ENERGY FUTURES

45

current global reactor fleet (and provide the basis for prolonged growth in energy output through provision of large amounts of new capacity), would require long-term and very substantial amounts of investment capital for development and construction (as well as for decommissioning and waste management of existing nuclear installations). The sources of the investment capital to drive such an expansion (which may require decadal commitment timelines) are highly uncertain, so national governments would most likely need to provide large amounts of partnership and financial support to drive such projects to final success, which also brings significant uncertainties (especially in Western countries) [100]. Additionally, analyses have indicated that the some SMR/AMR technologies may not offer significant performance advantages over existing nuclear technologies in some key respects such as overall radioactive waste production and may face public acceptance challenges if they undergo dispersed deployment as planned [101, 102]. Even with significant departures from and advancements over the current nuclear industry model with potential advanced and modular systems and other innovations, formidable capital investment and risk management would be required, and the inherent and fundamental factors that have and will continue to constrain the growth of fission technology will likely still apply. In particular, limits to the total (economically recoverable) global uranium resource, combined also with ever-present radioactive waste management, accident risk and public acceptability issues may continue to act as a significant drag on the potential for future growth (and particularly the very large-scale growth which would be needed to displace fossil fuels). Therefore, there are likely too many constraints and uncertainties associated with nuclear fission technologies to credibly change the outcomes of this Branch away from increasing costs and depleting resources also associated with fossil fuels (i.e., it could likely only delay and not avert the Fossil-Seneca paradigm; refer to the following section). To transcend these constraints, a technology would likely need to have fewer inherent limitations, and present lower societal risks whilst also relying on a fuel source of fundamentally different nature and provenance. The Seneca Effect Climate change and energy resource (fossil fuel and uranium) depletion have the potential to operate in conjunction, with the two phenomena

46

N. KING AND A. JONES

occurring simultaneously (or in close succession) and in synergy to impact global civilisation in a ‘pincer movement’. If the effects of fossil fuel depletion and/or increased costs were to manifest (i.e., as global energy shortages), it would likely result in emissions reductions [103], but the true extent of any such reductions are uncertain, and in any case the accumulation of past emissions means that climatic effects are increasingly ‘locked in’ even if emissions were to undergo rapid and deep reductions. There is a systems science description for this phenomenon of effects coinciding and operating synergistically to produce a more pronounced impact; the ‘Seneca Effect ’.8 This describes the situation in which the build-up of a given stock (such as socio-political complexity) within a system takes place slowly over an extended time period, but the collapse of that same stock occurs over a much shorter timeframe (an abrupt collapse after a period of steady and stable growth) [104]. The profile of a Seneca Effect-driven collapse is illustrated in Fig. 2.1, which demonstrates it in the context of total human energy use continuing on a BAU pathway. In this situation of continued, unmitigated fossil fuel use, energy use would continue to rise steadily until the egregious effects of climate change and depletion/cost increase combine9 to result in a collapse of energy use. In this context this would reflect a more general societal collapse to lower levels of socio-political (and other forms of) complexity. Globally dispersed nuclear fission technologies have the potential to generate additional risks in the event of a Seneca collapse; this is because the operation of nuclear reactors and supporting infrastructure (such as radioactive waste management facilities) require significant supporting socio-political and technological complexity, for example supervisory and regulatory institutions, global supply chains and power networks [105]. If these systems were to undergo significant degradation or failure, there could potentially be consequences (i.e., radiation releases to the environment) [106], which could potentially become an enhancing feedback exacerbating the effects arising separately from climate change and depletion. 8 The nomenclature of the Seneca Effect originates with the ancient Roman Stoic Philosopher Lucius Annaeus Seneca, who stated that “Fortune is of sluggish growth, but ruin is rapid”. 9 Other negative, systemic human impacts such as ecosystem destruction and chemical pollution would contribute to this as well, but these are not directly discussed due to energy focused nature of this discussion.

2

THE TRIFURCATION OF ENERGY FUTURES

47

FOSSIL-SENECA BRANCH ENERGY USAGE 6000 5000 4000 3000 2000 1000 0 1 3 5 7 9 11 13 15 17 19 21 23 25 27 29 31 33 35 37 39 41 43 45 47 49

Fig. 2.1 Fossil-Seneca Branch—illustration of system behaviour (Notes y-axis represent energy usage and the x-axis time units; all units are arbitrary and for the purpose of illustration of mode of system behaviour only. Energy usage is the metric applied to demonstrate collapse of human system function, but other measures [e.g., human population] could be used to demonstrate the effect as well)

The potential societal collapse considered in the Fossil-Seneca Branch is inherently impossible to accurately characterise or predict (as it is subject to the chaotic dynamics of complex systems), but equivalent events in historical societies suggest that different factors could act to make such an event occur over a range of different timescales, impacting systems and features (e.g., nations, total human population) to differential extents [41, 54, 107–109]. Although catastrophic collapse scenarios in which severe reductions in complexity occur rapidly (i.e., years or less) and simultaneously throughout the global system cannot be ruled out, it is more likely that the reduced availability of cheap energy would see economic growth in some countries falter, leading to increased incidents of failed states around the world [110]. These failed states would be subject to increasing climate change-induced disruption and damage

48

N. KING AND A. JONES

with little to no opportunities to adapt or mitigate these impacts causing cascading impacts to neighbouring countries. In states where increased energy costs can be more readily absorbed, increasing climate impacts acting in synergy with rising global instability would likely eventually create a destabilisation of energy (and other critical systems such as food) supply chains and the breaking of traditional geopolitical partnerships, resulting in the decay or collapse of globalisation. This would then likely lead to a collapse in energy usage on a global scale over a prolonged period (approximately decades—a century); the impacts of such a global scale event on human prosperity are impossible to foresee but are likely to be very far-reaching, complex, and potentially traumatic for many [111, 112], and could potentially involve significant human mortality [113]. Assessment of the Fossil-Seneca Branch Although inherently unpredictable, the potentially catastrophic scenarios which could occur as part of a climate change and depletion-induced Seneca ‘event’ are manifold. This is because there are multiple credible mechanisms by which egregious outcomes could result from a simultaneous departure from the narrow temperature range in which complex human societies have flourished and undermining of the mainstay of the global energy supply. Themes of how this could play out include general overwhelming of the adaptive capacity of human societies, reinforcing synchronous failures in different parts of the global system, and the triggering of discrete events which serve as ‘threat multipliers’ [114]. Overall, the Fossil-Seneca Branch of the trifurcating energy future would be a highly undesirable outcome for human civilisation, and significant effort should be directed to prevent this scenario, especially given that the current characteristics of the global energy system make it likely. Unfortunately, this branch of the trifurcation counts as BAU as this is currently the pathway that is supported globally by government policies and the overwhelming direction as set by collective business strategy.

The Continued Growth Branch The Continued Growth Branch of the trifurcating energy future describes scenarios in which there is a comprehensive, global scale transition to the use of new energy technologies. This will allow a full shift away from fossil

2

THE TRIFURCATION OF ENERGY FUTURES

49

fuelled BAU (thus avoiding a Seneca collapse) and the provision of energy supplies which are secure, abundant and have scope to be expanded in capacity, thus enabling human civilisation and its underpinning energy use to undergo continuous, ‘open-ended’ growth. However, many uncertainties exist for technology availability as well as the EROI available from existing technologies to sustain such growth. Technologies with Potential to Underpin Continued Growth Fossil fuel usage flourished and later came to dominate the global energy paradigm following the Industrial Revolution and provided the means for strong (and during recent timeframes, near-exponential) global growth over approximately two hundred years. However, continued global reliance on these fuels has the potential to introduce severe global instabilities from climate change and depletion, which indicates that their use can only ever be a temporary and evanescent phenomenon from the ‘open ended’ perspective.10 Several different energy technologies have also been introduced to the global energy mix during the period of fossil fuel dominance, namely renewables and nuclear power. However, despite decades of use and increasing technological development, investment and capacity increases, these technologies still contribute only a minority proportion of total global energy consumption and nuclear fission power, similarly to fossil fuels, is reliant on a finite and depletable resource. This Branch therefore considers the paradigm described above in terms of whether there might be the energetic basis for the phase out of fossil fuels and nuclear fission (thereby removing the systemic constraints they impose on human civilisation) whilst supporting continued general growth of the human endeavour (as fossil fuels have done to date). This would require energy technology (-ies) which utilises a very large and available energy resource (to circumvent the depletion constraint from the ‘open ended’ perspective) and do not emit GHG at rates which have the potential to (continue to) perturb prevailing climatic conditions (i.e., avoiding the climate change constraint). It also requires technologies for which future cost curves (the cost of fuel, development, deployment, use and waste management) are at least stable, if not declining.

10 In this context, ‘open ended’ refers to the longer term future on the scale of decadescenturies (i.e., up to approximately one thousand years).

50

N. KING AND A. JONES

The two broad energy technology groupings of nuclear fusion power and renewables have the potential to provide growing energy supplies on an ‘open ended’, continued growth basis over prolonged time periods. These two different technological grouping have some inherent, fundamental and contrasting characteristics which affect how they might interact with and underpin this scenario; this is expanded in the following bullets: • Nuclear Fusion Power—as opposed to nuclear fission which is the ‘splitting’ of heavy nuclei by the addition of neutrons, nuclear fusion is the ‘combination’ of two or more light nuclei, forming one or more heavier nuclei and releasing energetic subatomic particles. Anthropogenic nuclear fusion reactions have been achieved in the context of uncontrolled nuclear explosions and to very limited extents (i.e., number and duration of reactions) in more controlled experimental conditions, however no mature, net energy-producing fusion-based power systems were in existence at the time of writing. • Renewables—these are ‘flow’ based energy technologies, which means that they harvest a proportion of the energy contained within naturally-occurring flows or gradients of energy or energetic material. The dominant available flow of energy originates directly or indirectly from the Earth intersecting the flow of radiant energy from the Sun, with additional but smaller magnitude flows originating from the reservoir of primordial heat in the Earth’s interior and the gravitational interaction between the Earth and celestial bodies. The flows are constantly available and unchanging in magnitude on human-relevant timescales, so the renewables are not subject to depletion as stock-based reservoirs are, but are ‘rate-limited’ i.e., extraction from the flow can in theory (discounting the ‘real-world’ constraints effects of such extraction on the flow or medium in question) continue growing up to the maximum overall value of the flow (at which point all the energy in the flow would be appropriated), but no further (i.e., extraction would equilibrate with this maximum value). The total flow of solar energy arriving continuously at Earth (insolation; measured at the top of the atmosphere) is approximately 175 thousand trillion Watts (175 PW) [115], so the overall energetic magnitude available from this flow is very large, even if only a small fraction could be practically harvested.

2

THE TRIFURCATION OF ENERGY FUTURES

51

Due to the large energetic potential of both nuclear fusion power and renewables, both of these technology groupings have the potential to provide growing energy supplies on an ‘open ended’ basis. Both have inherent, large-scale constraints on the ‘ultimate’ quantities of energy they could supply, and therefore growth which they could support. However, on the approximately-defined timescales of ‘open ended’ growth, it is assumed that human civilisation would not be able to exhaust the total capacity limits of these technologies. Even if the current mean primary energy demand of global civilisation (18 trillion Watts; 18 TW [116]) were to expand by a very significant amount (two orders of magnitude over a decades-centuries timescale) this could likely be theoretically accommodated by the capacities of these technologies (the ‘real world’ biophysical and societal constraints to this actually being achieved are discussed in the following sections). The profile of continued growth is illustrated in Fig. 2.2, which illustrates energy consumption growing in an ‘open-ended’ mode. Characteristics, Potential and Constraints to Nuclear Fusion Technology Research into nuclear fusion as a controllable and useful energy source started in the early stages of the nuclear age following WWII (in parallel to the development of nuclear fission technologies). A number of different technological concepts for initiating, controlling and extracting energy from fusion reactions emerged, from which a subset were later identified as the most promising. All fusion systems must create conditions approximately analogous to that found in stellar cores (where stable fusion processes occur), namely extreme temperature and/or pressures, in order to achieve ‘ignition’ where fusion reactions take place and become selfsustaining. Creating the conditions for initiating fusion reactions is very energy intensive, so sustaining reactions such that the energy output is greater than the initial input (described as a ‘Q value’ of >1) is the key underpinning aim for achieving ignition. A number of fusion processes involving different combinations of light nuclei are possible, but the deuterium–tritium (DT) reaction occurs at the lowest temperature/pressure range so places the lowest (albeit still extreme) containment and control burden on the engineered systems in which it occurs, so the experimental systems aiming to achieving ignition and Q > 1 are based on this reaction. Two key feedstocks are required

52

N. KING AND A. JONES

CONTINUED GROWTH BRANCH ENERGY USAGE 20000 18000 16000 14000 12000 10000 8000 6000 4000 2000 0 1 3 5 7 9 11 13 15 17 19 21 23 25 27 29 31 33 35 37 39 41 43 45 47 49

Fig. 2.2 Continued growth branch—illustration of system behaviour (Notes yaxis represent energy usage and the x-axis time units; all units are arbitrary and for the purpose of illustration of mode of system behaviour only. The gradient of the line at different points in time and the energy consumption levels attained are illustrative only)

to fuel a DT-based fusion reactor; deuterium (a naturally occurring stable isotope of hydrogen) which directly participates in fusion reactions, and lithium, which is used to ‘breed’ tritium (the other hydrogen isotope which is unstable and therefore radioactive; ‘breeding’ occurs via transmutation under neutron irradiation, which can be supplied by nuclear reactions11 ). Deuterium is highly abundant on Earth as it is present in natural waters at a concentration of approximately 0.03 kg m−3 (making the magnitude of the total deuterium resource on Earth on the order of 100 million trillion kg) and therefore has near-ubiquitous distribution (in contrast to the fuel for fission systems). Lithium is present in moderate 11 These neutrons could be externally supplied either by fusion reactions (i.e., meaning that fusion power could become semi self-sustaining; the release of neutrons in fusion reactions could breed tritium fuel as a feedstock for future reactions) or by fission reactions in a hybrid system. In either case, the tritium would be located such it would be irradiated i.e., in a ‘blanket’ surrounding the fusion chamber.

2

THE TRIFURCATION OF ENERGY FUTURES

53

concentrations in Earth’s crust (significantly greater than uranium and thorium [117]) and in natural waters (seawater and geothermal brines), though is more heterogenous in its distribution [118, 119]. It is these characteristics of the fuel feedstock for the DT fusion process being abundant (extremely abundant in the case of deuterium), well-distributed and relatively available (though constraints to obtaining the feedstocks are discussed below), which could in theory provide the underpinning basis for the Continued Growth Branch scenario of open-ended energy supply growth. Despite the potential magnitude of the total fusion fuel resource, no nuclear fusion system able to sustain fusion reactions (and therefore able to achieve Q > 1), and much less able to output electricity on a commercially viable basis, has yet been achieved. However, R&D into viable systems is active globally, and is focused on the following broad approaches to initiating and sustaining fusion reactions: Magnetic Confinement Fusion (MCF), in which low pressure DT plasma confined by magnetic fields is heated to temperatures at which fusion can occur, using different reactor types (tokamaks, stellarators and reversed field pinch (RFP) devices); Inertial Confinement Fusion (ICF), in which laser or ion beams are focused on a millimetre-scale cryogenic mass (‘pellet’) of DT in order to implode and compress/heat the pellet (under its own inertia) to densities at which fusion can occur; and Magnetised Target Fusion (MTF), in which DT plasma is magnetically confined, with compressional heating provided by laser, electromagnetic or mechanical means (this approach combines some of the advantages of MCF/ MTF whilst reducing some of the more stringent requirements e.g., for prolonged plasma stability) [118]. Fusion R&D has been underway in different locations since the 1950s, funded and directed by a range of government, academic and private interests. MCF research has been conducted via tokamak (e.g., JET, MAST and TFTR to date, and STEP in future) and stellarator (LHD, 7-X and TJII) experiments, and is the subject of a major collaborative international effort in the form of ITER (International Thermonuclear Experimental Reactor) in southern France. This aims to demonstrate feasibility (but will not generate electricity) of large tokamak fusion systems (such as the construction and integration of very large cryogenic magnets) by generating a continuous Q value of ≥ 10 for up to 10 minutes (planned to occur during the late 2020s–2030s). ICF research

54

N. KING AND A. JONES

has been conducted primarily via the Lawrence Livermore National Laboratory National Ignition Facility (NIF) in California in the USA, in which approximately two hundred energetic lasers are focused on a DT pellet target. In 2022–2023 the NIF claimed to have achieved ignition and a Q value of approximately 1.5 during experimental runs. MTF research is currently led by the private company General Fusion, which aims to construct a demonstration plant in Culham in the UK. In this device magnetically-confined plasma will be injected into a compression chamber containing a liquid metal (maintained in a vortex by rapid spinning of the chamber) which will undergo mechanical compression using pneumatic pistons to collapse the vortex, generating fusion conditions in the plasma trapped in the centre of the vortex [118, 120]. Although investment in fusion R&D is significant and growing (approximately US$5 billion in private investment during 2022, and a total anticipated investment in the ITER experiment of approximately US$20 billion [121, 122]), significant scientific, engineering, logistical, political and other challenges remain to be solved. Effective solutions will be required before a commercially deployable fusion system (i.e., able to maintain a high, stable and controllable Q value over indefinite periods) is likely to emerge from one or more of the research and development efforts described above. Examples of key unresolved challenges include: (for the ITER tokamak) materials design and optimisation including for plasma-facing components (particularly the tritium-generating blanket) under expected extreme thermal, mechanical and electrodynamic loads; integration of fuel and thermodynamic cycles; quality control for very high specification components; high cooling and parasitic power demands of auxiliary systems; and materials able to withstand high flux, high energy neutron irradiation (and the subsequent management of large quantities of highly activated radioactive waste materials), and (for the NIF) economic target (pellet) design to allow ignition to occur reliably over a large number of events; and practical designs for the ‘fusion engine’ to extract and convert energy from the ignition events [123–128]. Separate but related and similarly critical issues arise from the challenges surrounding the provision of the DT fuel which will be required at industrial scale. Existing chemical techniques are available for the large scale production of deuterium from natural water feedstocks [129] (noting however that there will be significant energy input required for this at industrial scale, which will impact on the overall EROI of fusion

2

THE TRIFURCATION OF ENERGY FUTURES

55

systems), but it is the provision of tritium that may present more critical constraints. It is intended that lithium-containing ‘blankets’ will be used to breed tritium in operational fusion reactors, making them (partly) self-sustaining. However, some significant challenges and constraints must be overcome to enable this aspect of fuelling. Firstly, the feedstock for tritium ‘breeding’ must specifically be the isotope Li-6, which comprises only approximately 8% of natural lithium; isotopic separation is required to produce enriched Li-6, but this is not currently available at industrial scale. Some potentially scalable enrichment approaches have been suggested, but these remain at low TRL [130]. Experiments have also indicated that it may be challenging to make fusion systems truly selfsustaining due to inherent inefficiencies and entropic effects associated with tritium collection, storage and usage. Finally, in the experimental stages tritium must be sourced externally but global supplies of (nonmilitary12 ) tritium are highly constrained (being sourced at scale primarily from Heavy Water Reactors (HWRs) such as the CANDU-type13 and are also constantly decreasing due to radioactive decay) and the ITER project may consume the majority of the existing supply before 2030, such that later fusion reactors (of all types) would have very little available for experiments or operations [131, 132]. As with the nuclear fission technologies described above, the factors described above present fundamental and considerable challenges for fusion technologies to have the capacity to underpin the Continued Growth Branch. Even if the technical, fuel sourcing and other challenges associated with technology maturation were to be surmounted to produce a reliable, safe and technically and commercially viable fusion system or systems, the capex investment necessary to support deployment of large numbers of these systems worldwide, such that they create a continuously expanding energy base, would be very large. It is for reasons such as these that fusion has ‘remained 50 years away’ for decades [118], and there are strong indications that fundamental challenges will persist for the foreseeable future, even with significant R&D efforts.

12 A proportion of the global tritium supply is ring-fenced for military use i.e., to maintain thermonuclear warheads in operational condition. 13 The tritium from these reactors types arises primarily from neutron capture by deuterium nuclei; this is in incidental effect and neutron irradiation of deuterium would not be a practical source of tritium.

56

N. KING AND A. JONES

Renewables As noted, the different renewables technologies have undergone very significant global capacity increases in recent decades, but even with these increases, renewables have to date achieved only an insignificant fraction of their total potential. As noted above, the total insolation reaching Earth on a constant basis is 175 PW [115], a proportion of which can be harvested directly in primary form via solar power, or indirectly in secondary form via the harvesting of kinetic energy of environmental media (primarily wind and hydroelectric power, which result from the solar-driven climatic system). The harvestable proportions of this are a fraction of this total, but still large: the solar resource in North Africa (one of the areas of the world with the highest solar irradiance values, but only a fraction of the total global resource) could be approximately 70 TW (multiples of current global power demand) [115]. The mean total global near surface kinetic energy (wind) dissipation value is 336 TW (an order of magnitude greater than current global power demand) [116]; and extractable (enhanced) geothermal power at global scale may be on the order of 200 TW (approximately an order of magnitude larger than current global power demand) [133]. Although renewables comprise complex mechanical systems which interact with dynamic and high energy natural environments (wind turbines), incorporate highly structured, chemically complex materials to efficiently convert insolation to electrical power (PV panels), and interact with high temperature corrosive steam (geothermal plants) these technologies are still significantly less complex and capital intensive, and require less regulation, than nuclear technologies. This difference in inherent complexity is likely a factor (along with others such as potential direct competition [134]) in the rapid technological development and accelerating deployment of renewables in parallel to ongoing reductions in deployment and generation costs, whilst the cost of nuclear technologies has risen in the same timeframes. The capital investment costs of renewable generating systems have undergone very significant reductions globally since approximately 2010, with the global average levelised cost of electricity generation for PV falling by 77%, onshore wind by 35% and battery storage by 85% during the period 2010–2018 [135]. The underlying causes of these shift are complex, but can is largely be attributed to rapid technological advances arising from investments in R&D, increasing economies of scale of

2

THE TRIFURCATION OF ENERGY FUTURES

57

manufacture and installations, accruing experience in development and deployment, and government interventions such as subsidies and feed-in tariffs which provided initial impetus to the expansion. Comparing renewable costs with fossil fuel costs is complex; the bulk of costs for renewables are the up-front capital investment requirements (with some maintenance costs), which as indicated above have rapidly reduced over a relatively short space of time and will continue to do so. Meanwhile the costs for fossil fuels are a mix of up-front capital investment costs (and maintenance) and lifetime fuel costs [136]. Therefore, any comparison of costs of electricity need to consider the lifetime of assets (a longer lifetime means up-front costs can be spread over a longer period, while increasing fuel costs), and needs to take into consideration any likely changes to those up-front investment costs. These changes in the economics of renewables had not been anticipated in the majority of energy scenario planning undertaken historically, but forward-looking analyses which take account of these shifts have indicated the potential for tipping points to drive ‘breakthrough effects’ , in which the deployment of renewables systems undergoes (further) nonlinear upwards shifts alongside further reduction in investment costs. The acceleration in deployment/reductions in costs for renewables for grid systems as have already occurred may be the initial manifestations of this effect [137, 138]. This also implies that further breakthroughs in future, such as potential increases in the EROI achievable by renewables through avoidance of ‘curtailment ’ using storage systems [139] (the practicality of which is being increasingly underpinned by innovations such as the potential use of conventional and economic systems and materials to scale up storage capacities [140]) and demand management, may be possible. The global renewables resource is demonstrably very large and there have been significant successes in deploying renewables at global scale, however there are a number of significant potential constraints to further large scale increases in the deployment of renewables globally (which are of a different nature to that affecting nuclear power; these are explored in the context of the Stabilisation Branch). Therefore, achieving ‘open ended’ growth likely represents an inherently large challenge independent of the types of technology underpinning it. The nature of open ended growth in terms of externalities and other likely consequences is nonetheless explored in the following section.

58

N. KING AND A. JONES

Prospects and Consequences of Continued Growth This section explores a hypothetical scenario in which the constraints discussed in the previous (and following) sections are surmounted,14 thus allowing exploration of the following question: what might be the primary implications for humanity and the Earth System of open-ended, unfettered growth underpinned by an (essentially) unlimited energy base? Global Economic and Social Paradigms Two centuries of near-continuous economic growth (since the inception of the Industrial Revolution) has resulted in a drastic reduction in the proportion of the world’s population living in extreme poverty, even with the approximate eightfold increase in population which occurred in parallel during that time [141]. Although this is a considerable achievement in terms of human development overall, the proceeds of this prolonged period of growth have not been distributed equitably, and it has also generated a wide range of externalities. The extant economic paradigm might be described as a zero or negative sum situation, with the proceeds of economic growth having benefitted a relatively small subset of the human population disproportionately (zero sum; gains for some have been made at the expense of others). This results directly from a key failing with the capitalistic, market driven system which has come to increasingly dominate the global economic organisation and leads directly, by design, to inequality. Capitalism by definition applies an enhancing feedback mechanism to concentrate wealth towards individuals and organisations which already hold capital, and over the course of the Great Acceleration, large-scale operation of this mechanism has steadily increased inequality at many scales. Regional and national income inequality (measured in terms of core and peripheral regions of the global system) has tripled, a phenomenon which has been driven by mechanisms such as ongoing flows of resources from south to north [142, 143], and is replicated 14 This is assumed to be via fusion technology developed to operate commercially and reliably with a steady and viable fuel supply (i.e., with deuterium separated in industrial quantities at an acceptable energy cost, and tritium produced in sufficient quantities using lithium blankets), potentially supplemented by Gen IV and/or SMR/AMR fission systems, and/or deployment of multiple renewables systems at scales sufficient to harness large scale solar and wind resources. The practical details of achieving this are not provided; this assumption is only to enable the though experiment.

2

THE TRIFURCATION OF ENERGY FUTURES

59

at local level with widespread instances of poverty and extreme wealth capture by a small ‘elite’ of super-wealthy [144, 145]. Externalities have manifested primarily in the form of cumulative global climate change, biodiversity loss and pollution, impacting the Earth System and biosphere. The contemporary economic paradigm has operated effectively whilst the impacts of the externalities on growth have been relatively minor, but these have increasingly been producing feedbacks on the function of societies as they become cumulative and synergistic. The consequence of this may be that the gains of this subset of the population may be eventually overwhelmed by the externalities via societal degradation and collapse, which would be negative sum where all gains are eventually lost. In particular, unequal societies can be inherently unstable and over time, with external shocks, this instability can lead to collapse through social protests and civil unrest. One notable exception may be India where the caste system has been remarkably resilient to change over time. An alternative would be a positive sum situation in the benefits from (further) growth could be distributed more equitably, and negative effects (from past and ongoing growth) could be effectively ameliorated (gains achieved by all). ‘Fully Automated Luxury Communism’ (FALC) is a high level concept for potentially achieving a positive sum economic paradigm, in which limitless energy (in line with the Continued Growth Branch), expanding resource horizons and technological innovations would provide universal abundance along with redundancy of human labour [146]. Possibilities under this paradigm include existing pollution and scarcity emergent from growth being solved simply by virtue of energy availability. Examples of this may include: industrial-scale seawater desalination driven by abundant electricity (‘too cheap to meter’) which could alleviate droughts and other shortages globally; unlimited process energy which could enable the establishment of a truly circular recycling economy; Direct Air Capture (DAC)-based CO2 removal from the atmosphere driven by scrubbers which could operate at industrial scale without cease; and meeting of power and cooling demands for ever-greater levels of high-performance computing and artificial intelligence (AI) which could be tasked with automated societal optimisation [147, 148]. Concepts such as these tend to focus on the commodity and the material aspects of future societies enabled by advanced technologies, but do not provide underpinning detail about consequences from human and societal perspectives. Previous major re-organisations of society (such as occurred after the Agricultural and Industrial Revolutions) led to drastic

60

N. KING AND A. JONES

phase shifts in the basis of societies, and a shift to a society based on FALC (or equivalent) would potentially generate shifts of a similar magnitude. These shifts could include continuations and extrapolations of current trends and phenomena, including: ongoing capture of wealth by ‘elites’ (despite the egalitarian and positive sum intent of the concept); magnification of innate human behaviours and biases (such as conflict and inequality arising from aggression and tribalism) and unforeseen externalities (including societal instabilities arising from internet ‘information overload’). Human societies are complex, adaptive, dynamic systems so it is not possible to determine how societies and economies (organised according to FALC or other models) would evolve under ‘open ended’ growth, but extrapolation of trends such as those described above is likely a reasonable assumption. Of these, the potential for societal instabilities and collapse dynamics to be driven by continued ‘elite’ resource appropriation is a notable risk [149]. Physical and Digital Footprint The accelerating physical and spatial expansion of human activity and influence over the surface of the Earth has been one of the most direct consequences of economic growth to date. This has generally manifested as the overproduction of a large range of manufactured materials, the totality of which is labelled as the ‘Technosphere’. This comprises anthropogenic material actively moving through human built systems, and additionally the significant ‘residue layer’ of accumulating, high entropy material of which only a small component cycles back into the active system. The ‘Technosphere’ is estimated to have a total mass of approximately 30 trillion tonnes, exceeding that of the biosphere, and its total material fluxes rival or exceed total global fluxes of geological material transported by Earth System processes [150, 151]. One of the largest components of the ‘Technosphere’ is plastics (second only to bulk construction materials such as cement). Plastic has undergone very rapid growth in production since the 1950s, and of the approximate 8 billion tonnes ever produced the vast majority (~80%) has been landfilled or has spread entropically into the environment, where it has entered a large range of ecosystems and environmental niches and has the potential to generate significant ongoing harm [152]. In addition to the spread of polluting agents such as plastics, human influence has also taken other forms including urbanisation, agriculture

2

THE TRIFURCATION OF ENERGY FUTURES

61

and resource extraction. These all tend to alter and impair natural environment function such that at global scale, human action has extensively destroyed and displaced ecosystems and other natural features (such that less than 3% of the Earth’s land surface is now considered to be ‘functionally intact ’ with full ecological integrity unaffected by human influence). This has also led to a 69% decline in wildlife populations around the world during the last approximate 50 years [153, 154]. All of these phenomena were unintended consequences of the energy and material surpluses provided by fossil fuels, so open-ended economic growth underpinned by larger energy resources could be robustly expected to continue to add mass to the ‘Technosphere’ and thereby generate a continuation of its negative effects. Even if a significant proportion of energy consumption were devoted to amelioration of anthropogenic mass, as suggested by FALC, the existing quantities are so large this could likely not be readily remediated. Following its origins in the late twentieth century, the internet has undergone exponential growth in terms of numbers of users and total data traffic, such that by the 2020s, it has become extensively integrated into most aspects of global society. As a phenomenon emergent from earlier technologies (electrical grids, digital computers and communications networks), it has undoubtedly brought beneficial changes to human societies through unprecedented connectivity, exchange of ideas across time and space, efficiencies in information and people management, cataloguing and searchability of information, and the streamlining of processes such as logistics. However, the internet, and particularly social media (a sub-phenomenon of the internet which has expanded and accelerated user interconnectivity) has increasingly introduced destabilising tendencies of a type which were not possible prior to the inception of the internet into societies across the world. These phenomena arise partly from the hyper-connectivity the internet enables allowing the spread of ideas and ideologies and the formation of non-geographically based communities which ‘coagulate’ and ‘cohere’ due to algorithms promoting their connection. This trend has increasingly challenged the authority of the nation-state. In parallel, the sheer quantity and accessibility of information on the internet, exacerbated also by the algorithms of social media (which are programmed to exploit human cognitive biases), has driven rapid growth in misinformation and ‘fake news’ of many types. These phenomena threaten to weaken the collective consent and agreement as what constitutes authoritative information

62

N. KING AND A. JONES

and what is indeed true and correct, as well as potentially reducing the attention spans and cognitive abilities of humans at population scale [155– 157]. The increasing prevalence of AI in cyberspace and other domains is likely the next ‘stage’ in this process and has the potential to introduce significant new risk paradigms [158]. The dependency and escalation of commitment to the internet (and increasingly AI) to date (due to a large degree to the migration of most economic, governmental and knowledge storage activity to that domain) means that open-ended growth would most likely require a continuation of these trends, along with the harms they have the potential to bring [159]. Energy Accumulation Natural systems in which there are energy flows and gradients experience energy-dissipating events (e.g., earthquakes in tectonic systems, lightning within atmospheric storm systems), which often follow powerlaw distributions in terms of their magnitude. Most importantly, the return frequencies of larger, more energetic events are distributed with certain intervals. If human civilisation is considered to be a thermodynamic system dissipating energy gradients (and seeking Maximum Entropy Production akin to natural systems [160]), and war (i.e., organised conflict between human groups in which large quantities of energy are expended) is considered to be one of the most effective entropyproducing mechanisms through destruction of infrastructure [161], equivalent and similar effects may be observable. Analysis of historical (and recent) datasets has indicated that the frequency and severity of conflicts (i.e., deaths and economic costs) may follow a universal statistical pattern, namely a Poisson arrival process with a power-law distribution applying to severity [162–164]. On this basis, if the total amount of energy available in the human system was greatly increased through open ended growth, the total number of wars of all sizes (including large and very destructive conflicts) occurring globally over time could potentially increase from the contemporary prevalence of approximately one hundred discrete conflicts to a much larger number [165]. This perspective aligns with the idea that the dynamics of human civilisation is dictated at large scale as much by thermodynamics as by human agency (as captured by the Superorganism concept). The final negative aspect of open-ended growth presented here concerns another unavoidable effect of thermodynamics; namely the release of waste heat, which may accumulate in the Earth System (noting

2

THE TRIFURCATION OF ENERGY FUTURES

63

that this is separate to but additional to the accumulation of thermal energy in the Earth System due to the changes in atmospheric radiative forcing as a result of GHGs). Waste heat (which cannot be used as part of ‘energy services’ i.e., useful work) is generated as a function of thermodynamic laws and comprises the predominant fraction of energy output from human systems (67% of the total energy consumption on the USA in 2021 [166]). This applies primarily to ‘stock’ based energy technologies which release stored energy in the form of heat (so is applicable to all forms of nuclear power) but also ‘flow’ based energy technologies. The energetic flows in the Earth System such as sunlight occur independently of human appropriation of a fraction of the flow, but this collection and use will generate waste heat (i.e., solar cells may absorb 90% of incident insolation, but only 18% is converted to electricity; the rest is lost as waste heat [167]). The total power output (and therefore waste heat generation) of current global civilisation is trivial compared to the power of natural energy flows (primarily insolation; 18 TW vs 175 PW or approximately four orders of magnitude), but in the Continued Growth Branch scenario the power output could potentially grow to a more appreciable fraction of natural flows, and the waste heat would be cumulative within the (partly) closed Earth System, so would inevitably and eventually lead to a corresponding rise in temperature (potentially up to 3 °C on a thousand year timescale) [168, 169]. This level of heat rise would be commensurate to those which could contribute to the scenarios described above, and would be additional to any GHG-driven temperature rises. Assessment of the Continued Growth Branch The Continued Growth Branch presents a scenario in which combinations of fusion and/or renewables technologies are successfully developed to technical maturity (in the case of fusion technologies) and deployed at scale globally such that they underpin open ended energy and economic growth. These scenarios incorporate a range of uncertainties which are primarily linked to the solving of technical and societal challenges associated with progression of development of the technologies, overcoming systemic constraints, and obtaining the resources/securing the capital to underpin global deployment of extensive nuclear and/or renewables infrastructure. Even if surmounting these challenges were readily feasible and attainable, the resulting indefinite growth may to lead to a range of egregious effects and externalities, some of which are extrapolations

64

N. KING AND A. JONES

of phenomena which have arisen from fossil-fuelled growth to date, and others which may comprise new systemic effects for global society. Open-ended growth of the human endeavour is a theoretical construct; natural systems can only grow for limited periods before attaining some form of equilibrium (or undergoing collapse). The equivalent constraints on the growth of human civilisation have been recognised for some time (a position which has been reiterated and confirmed more recently) [170, 171]. Overall, there are open questions as to whether Continued Growth Branch would be attainable, and even if the challenges associated with it could credibly be overcome, it would likely not be a desirable long-term outcome for human civilisation.

The Stabilisation Branch The Stabilisation Branch of the trifurcating energy future describes scenarios in which there is a comprehensive, global scale transition to the use of new energy technologies. This will allow a full shift away from fossil fuelled BAU (thus avoiding a Seneca collapse) and a shift to a paradigm of long-term stabilisation in the overall rate of global energy use. Long-Term Stability of Complex Systems A fundamental feature and behaviour of many complex systems is occurrence of ‘adaptive cycles ’ in which endogenous dynamics drive constant shifts and progression through foreseeable cyclical states (namely growth, conservation, collapse and renewal) [172, 173]. However, a number of natural complex systems are observed to be exceptions to this tendency, by attaining stable (or quasi-stable) states over prolonged time periods. These states are observed to emerge primarily as a result of reaching equilibrium (or near-equilibrium) conditions in alignment and accordance with long-term, steady (or near steady; varying at sufficiently slow rates to permit the system in question to adapt and align concurrently) value environmental conditions or exogeneous inputs. Examples of complex systems reaching this long-term equilibrium include: • Total biosphere mass—the biosphere comprises approximately 550 billion tonnes of carbon [174], which is likely to be commensurate with (close to or at) the planetary carrying capacity (in terms of overall biomass). This is set (primarily) by the total amount of

2

THE TRIFURCATION OF ENERGY FUTURES

65

global photosynthesis that can take place in aggregate, which is in turn limited by the available solar energy input at planetary scale. The complex terrestrial biosphere has evolved (since its emergence at the start of the Phanerozoic Eon approximately 550 million years ago) and stabilised in terms of total mass around the total exergy value which the Sun has continuously input to the Earth System over time. The solar constant has varied over geological time, with solar luminosity being approximately 70% of its current value in early Earth history, and 95% at the start of the Phanerozoic [175]. The condition of stable environmental homeostasis has likely emerged as a result of feedback loops operating between life and the Earth environment. These have allowed the total mass of the biosphere to remain broadly constant whilst accommodating and adapting to the long-term energetic variance (and increasing in internal complexity and diversity within the mass limit over time). This phenomenon is labelled as ‘Gaia’ [176]. • Stellar energy output stability—stars of all types, including the Sun, attain a state of long-term output stability (or quasi-stability, for some star types) for as long as they remain on the main sequence. This occurs through stars maintaining conditions of hydrostatic and thermal equilibrium, which emerges from a balance of feedbacks between gravitation compression (which tends to compress and heat the stellar core) and radiative pressure (which tends to counteract the compression and lower the heating of the stellar core). The fusion which occurs in stars is as a result of what might be labelled as ‘gravitation confinement’ (in contrast to anthropogenic approaches as described above), with compressional heating of the core increasing the rate of fusion events, and the radiative pressure lowering this rate. It is the continual interplay of these ‘ambient’ conditions which leads to the stabilisation around a fusion output value which maintains the equilibrium conditions of the star over extended periods of time [177]. If the human system is to attain a condition of long-term stability, it will need to emulate the fundamental characteristics of the systems described above. Namely, it will need to stabilise around (either at or below) the limiting value of freely available and surplus energy (i.e., that not required for the function of biosphere and/or to maintain vital Earth System processes such as the climate system). The potential to do

66

N. KING AND A. JONES

this may be supported by the observed tendency for different types of natural complex systems to exhibit similar and equivalent features and selforganisation across large ranges of scales [178], so anthropogenic systems may be able to achieve similar replication of systemic features. Energy Technologies and Stabilisation The key technologies with large energy capacity potentials have different characteristics which interact with the requirements of open-ended growth in different ways. These technologies also have different characteristics in the context of long-term stabilisation, which is captured in the following bullets: • ‘Stock’ based energy systems (whether fossil carbon or fissile/fusible atomic nuclei, but particularly the latter) could in theory stabilise in terms of energy output and remain at that level for a prolonged period. The emergence of such a paradigm would however not be as a result of interactions within a complex systems (via interaction equivalent to Lotka-Volterra dynamics [179]). Instead, this would have to be as a result of deliberate intent and decisions made by human agency, which the Superorganism concept indicates might not be readily possible. Secondly, even the stabilised and constant output of a ‘stock’ based system would eventually be subject to depletion (even if the drawdown of the reservoir of fuel were a low, stabilised value), so could not be a long-term solution akin to the examples given previously. • ‘Flow’ based energy systems (based on renewables technologies) are inherently suited to stabilisation. This is because these technologies harvest energy from natural flows and are limited by the total amount which they can appropriate by the maximum total magnitude of the flow. They can extract that much and no more (and in practice will only be able to extract less than the maximum value in accordance with physical laws such as the Betz limit [180]). The flows in question are (primarily) based on solar energy output so from human perspectives will continue indefinitely. Once extractive technologies are aligned with flow values, they can continue without effective limit. Here resources required to capture these flows (such as minerals for renewable infrastructure) will need to be managed through a circular economy or be substituted for resources that are

2

THE TRIFURCATION OF ENERGY FUTURES

67

also subject to ‘flow’ on timescales that match demand. Ergo, to provide a stable, (truly) open ended energy future for human civilisation, a transition to a (predominantly) renewables-based energy system will be required.

A Renewables-Based Civilisation The stabilisation of the human energy paradigm around the energy limit available from solar input-derived environmental energy flows is represented in Fig. 2.3. The stabilised, environmental flow-limited civilisation represented in Fig. 2.3 would be based on 100% (or near 100%) renewables system, which is a paradigm which has been increasingly studied as renewables technologies have penetrated further into the global energy system in

RENEWABLES STABILISATION BRANCH ENERGY USAGE 1800 1600 1400 1200 1000 800 600 400 200 0 1 3 5 7 9 111315171921232527293133353739414345474951535557596163656769717375 ENERGY USAGE

ENERGY INPUT

Fig. 2.3 Renewables Stabilisation Branch—illustration of system behaviour (Notes y-axis represent energy usage and the x-axis time units; all units are arbitrary and for the purpose of illustration of mode of system behaviour only. The gradient of the line at different points in time and the energy consumption levels attained are illustrative only. The ‘Energy Input’ level is illustrative and is not representative of any physical values)

68

N. KING AND A. JONES

recent decades. A significant body of research supports the position that it would be possible to transition to a 100% renewables-based system on timelines and at costs that are eminently feasible at global scale [181– 183], whilst other research has identified challenges to achieving and operating such a system at large scale [184]. These positions must be considered in light of the Stabilisation Branch scenario. Challenges for the Renewables Transition The following subsections explore and analyse some of the fundamental systemic challenges to implementing a global transition to a renewablesbased energy system. Systemic Challenges The transition described in this scenario would require a foundational shift in many aspects of the organisation of societies globally. Therefore, the biophysical, economic, social and technological systemic aspects need to be fully understood in terms of feasibility and implications. A number of Integrated Assessment Models (IAMs) have been developed to assess the interlinking environmental, economic and societal systems along with the uncertainties and knowledge gaps. For example, the MEDEAS modelling framework is an IAM which assesses the sustainable energy transition15 by addressing limitations in predecessor IAMs. These earlier IAMs were found to incorporate a number of divergences from robust biophysical realities and assumptions, including (but not limited to): sequential model structures with limited internal feedbacks; differential accounting for thermodynamics depending on whether scientific or economic-based models were favoured; that future energy transitions are purely demand-driven transformations with only monetary (and not energy and material) constraints; and lack of systemic consideration of the implications of EROI [185, 186]. The MEDEAS framework addresses these shortfalls (particularly the low prevalence of feedback mechanisms) through modelling, assessment and computation of factors including (but not limited to): biophysical constraints to energy availability; mineral and energy investments and 15 The ‘sustainable energy transitions’ is the term commonly used to describe a large scale shift away from fossil fuel based systems, and primarily refers to renewables based systems.

2

THE TRIFURCATION OF ENERGY FUTURES

69

scarcities; net energy available to society; and sectoral economic structure (input–output analysis). The MEDEAS framework is spatially and temporally flexible in that it can be applied at world, regional and country scales and to time horizons of between 2060–2100. An initial application of the framework was to simulate four energy transition scenarios which extrapolated current trends at global scale into future decades (i.e., BAU scenarios, with continued use of fossil fuels to varying extents into the future as renewables transition progresses). One of these scenarios is a reference scenario and three explore some key variables within the model. These variables are related to: fossil fuel resource availability (total abundance and maximum extractability); renewable energy source techno-ecological resource potential (intermittency and integration in the energy mix— ranging from 60–80% of the global total, seasonality and uneven spatial distribution, lower energy density and land use, mineral and materials use, and environmental impacts); and the climate change damage function (when the global average surface temperature change reaches +1.75 °C, GDP gains are cancelled out). Energy availability and climate change damages arising from these effects therefore act as ‘limits to growth’ within the MEDEAS framework. The two ‘limits to growth’ in different combinations produced results that differed significantly from the outputs of the predecessor IAMs. Omission of these limits generated outputs broadly similar to those of other IAMs, but where they were permitted to operate and interact within the model, completely new dynamics were output. The BAU scenario with highest renewables penetration into the global energy mix was found to drive large mineral, energy and land-use demands, causing a persistent global economic recession due to energy scarcity and climate damages (which is commensurate with the economic impacts of energy scarcity and climate change in lower penetration scenarios). This model output indicates the potential for a widespread systemic global socioeconomic and environmental crisis (driven by the transition to a highly renewables-dependent energy system) which could lead to shift to a ‘regional competition’ geopolitical situation (i.e., fragmentation of the world into competing power blocks), with potential to lead to conflict (or even collapse) scenarios. Additional novel insights that the dynamical assessment enabled by the MEDEAS framework highlighted were that potential mineral scarcities and societal net energy availability may require trade-offs between

70

N. KING AND A. JONES

urgent climate mitigation, and the viability of the energy systems which may provide that mitigation. An outcome of this could be a drive for more rapid deployment of alternative energy systems which would in turn reduce the whole-system EROI. This highlights the eminence of biophysical factors over monetary costs. Rapid transition to renewables systems would also require re-materialisation of the economy which could have the effect of nullifying historic trends for efficiency improvement in the energy system. It is noted that the MEDEAS model may also be applied to assess different scenarios including different policy interventions, but the scenarios described above are the limiting worst cases for BAU, which have minimal emphasis on sustainability [187]. Re-materialisation The high-technology industrial paradigm which currently underpins global human civilisation is material-intensive and therefore entirely dependent on the extraction of non-renewable, natural mineral and other resources to function. The MEDEAS modelling framework highlights that a transition to a renewables-based energy system would align with that paradigm by requiring an enabling ‘re-materialisation’. That is the large scale construction of a renewables-based system would require the material intensity of human civilisation to rise further. The existing industrial ecosystem which underpins complex modern societies is the product of development over approximately one hundred years, enabled by abundant fossil fuels, minerals and financial credit [56]. The future transition to an industrial system underpinned in the majority by renewables would require the input of very large quantities of mineral resources to build the required mass-scale renewables systems. It is increasingly recognised that for partly or fully-renewables based system to be viable, storage systems (e.g., battery storage plants and pumped hydro facilities) or installation of ‘overcapacity’ (additional generation capacity to compensate for spatial variations) to smooth out the inherent variability and intermittency of renewables technologies [188, 189] will be required. The scale of these measures will need to be proportional to the overall size of any given renewables system, and this supporting infrastructure will significantly contribute to the overall materials requirement. Crucially, these material requirements would need a ramp up in resource extraction and processing on a much more rapid timescale than previous transitions, which will also be required at a point in history characterised by depleting fossil energy resources, a fragile and heavily

2

THE TRIFURCATION OF ENERGY FUTURES

71

indebted global financial system, growing human numbers and energy demand, and generally degrading environmental conditions. Fossil fuels have been extracted on the basis of the ‘low hanging fruit’ principle, and the same is true of mineral ores. The highest quality deposits have been exploited preferentially which has resulted in the grade of raw materials being extracted declining over time [78]. This trend has obligated greater energy and water input for processing (a trend which has been accelerated by technological developments requiring higher material purities), greater waste rock production, and social costs. Extractive activity is primarily powered by fossil fuels, and the increasing costs/ declining EROI of fossil fuels globally further exacerbates the problem of decreasing mineral grades. Market and economic instruments (e.g., supply and demand management) have been deployed to manage this issue to date, but this approach is limited in the potential scope of its effectiveness, and without cheap fossil energy input the cost of extracting and processing increasingly depleting reserves will rise (likely leading to reduced overall production rates). A globe-spanning renewables-based energy system would require the mass input of certain metals and other materials (including, but not limited to, steel, concrete, lithium, cobalt, nickel and rare earth metals). A number of these materials are not currently undergoing extraction at rates which would theoretically support such a transition. Several of these minerals have never been required in bulk quantities for any applications previously, so infrastructure for their extraction does not currently exist. Given this context, there is a risk that the remaining overall global reserves of some resources may be insufficient to fully support the transition. There is the potential for ‘new frontiers’ for mineral extraction to be opened to contribute towards the renewables transition, notably deep-sea mining of polymetallic nodules and other deposits. However, the feasibility, economics and environmental impacts of such extraction remain uncertain [190]. The reuse of previously extracted key materials through mass recycling/circular economy systems could potentially ‘close the cycle’ to a large extent and remove much of the need to mine virgin materials. Despite recent advances [191] many technologies are at low TRLs and have not yet been constructed at scale [192–194]. Therefore, to fully unlock the potential of a circular (energy) economy, much more investment is required in a short space of time.

72

N. KING AND A. JONES

Externalities of Renewables The transition to a renewables-based energy system would be a materially intensive process. In common with all engineered systems, renewable technologies have finite operational lifetimes and so will produce quantities of end-of-life waste commensurate with the material input required for their construction and operation. The end-of-life waste management burden of decommissioned renewables system at global scale is already of significant magnitude with approximately 10% penetration into the global energy system [195]. This will increase in proportion to future further penetration of renewables. Wind energy is projected to produce 43 million tonnes of waste annually by 2050, and for redundant solar panels which, in contrast to the largely inert material produced from wind turbines, contain a range of toxic substances such as lead and cadmium, the global total in 2016 was 250 thousand tonnes [196]. A circular economy for renewables could largely ameliorate this waste production, but much of the requisite infrastructure and technologies to support reuse and recycle are not yet available and will likely not be until the ‘first generation’ of redundant equipment starts coming ‘on stream’ at scale such that the investment for construction of the required recycling infrastructure can be justified. Pending the availability of these enablers, landfilling will likely remain the main destination for the majority of renewables waste. The natural energy flows which renewable energy systems exploit are less energy-dense and more diffuse than the fuels underpinning ‘stock’ based systems and therefore the collection infrastructure (and hence land use) is large. The land occupied by the expanding infrastructure of a renewables transition could therefore potentially become a significant future driver of land use globally and therefore a significant competitor with anthropogenic food production systems (though in certain circumstances there may not be competition), and with the natural habitats and biomes underpinning biodiversity. In the latter case, impacts may include direct biomass removal (to make way for physical structures such as foundations) and land cover change (which may be particularly acute in the desert habitats suited to large PV and CSP systems), habitat fragmentation (e.g., clearing of forests) and degradation/disruption of ecosystem services (e.g., disruption of pollinators). There is also the potential for impacts to ecosystems to be greater than just the physical spatial area occupied. The ‘zone of influence’ of infrastructure can extend over several kilometres and therefore the ‘virtual footprint’ of renewables systems could become significant [197–200].

2

THE TRIFURCATION OF ENERGY FUTURES

73

The global economic system is currently dominated by the capitalistgrowth paradigm, and many of the mitigations and responses suggested for systemic challenges confronting civilisation can be defined as being ‘techno-optimist’ (or ‘eco-modernist’). That is, they are based on and apply technological and economic/market-based instruments which align with the overarching paradigm and make optimistic assumptions about the continuation of economic growth and the stability of current anthropogenic systems [201]. Therefore, an indirect externality of a renewables transition may be to create the conditions for continued techno-optimist views to thrive. This could manifest as hedging on energy technologies (renewable or otherwise) and applications (increased consumptive behaviours) which could drive effects such as those described above, and be contrary to and undermining of transformative economic and social efforts. In the context of a renewables transition, outcomes such as this could potentially drive the global system towards the dynamics described in the worst case MEDEAS scenarios. Constraints to New Frontiers for Renewables The preceding sections describe ‘conventional’ renewables. These harvest natural energy flows available at or near the Earth’s surface using ‘standard’ infrastructure and systems (e.g., concrete foundations in the ground or seabed, connections to established power grids). The energy flows themselves in this environment are of a magnitude which precludes them from being a realistic constraint to the expansion of renewables, but the constraints and externalities described in the preceding sections could start to act as ‘brakes’ on the development of the system to underpin a full transition. One feasible mitigation to these constraints could be to expand renewables infrastructure beyond the immediate Earth surface to locations where flows are available but the constraints described above apply to a lesser extent. The first of these possibilities is exploitation of insolation outside of Earth’s atmosphere, which is labelled as ‘space-based solar power’. This would utilise solar power satellites which would be placed in low earth orbit (or beyond) to collect solar radiation. The collected energy would be converted to tight-beam microwave radiation to allow transmission of the power to dedicated receiving stations on the Earth’s surface. Advantages of this would (besides avoiding the constraints and externalities described) include: increased energy gain relative to surface based systems as the energy harvesting could take place within a volume of space much

74

N. KING AND A. JONES

greater than Earth’s surface; the lack of clouds/air/dust to limit the efficiency of the PV arrays; the direction of the collected energy to users and locations as required; and the use of several existing and mature technologies. However, there are numerous challenges related to the scalability, economics and environmental impact of launching the required orbital equipment and building the ground-based infrastructure, the large maintenance costs of space-based collectors, and the likely low EROI of the overall system due to the multiple energy conversions which would be required [202–204]. The second possibility relates to expansion of geothermal energy to locations where the energy gradient is much greater than those currently exploited in ‘conventional’ geothermal systems. Specifically, this would directly target extreme geologic environments where high temperature Earth materials are located close to the surface at large scale, such as active volcanic calderas (e.g., Yellowstone). Such environments present significant energy gradients which could in principle support generation of very large quantities of power. However, practical exploitation of such a resource would require technological capabilities and novel engineering approaches which are at very low TRL and would therefore present large risks of both engineering failure, and also perturbation of complex natural systems with the potential unintended consequences (e.g., initiation of seismic or volcanic activity) [205]. Overall, there are likely to be significant constraints on the feasibility (primarily related to investment costs, scalability and the maturity of the required technologies) of large scale deployment of alternative renewables technologies, which may be more limiting than the constraints on ‘conventional’ renewables. Therefore, there is likely very little scope for non-conventional technologies to supplement or replace the technologies which currently underpin the renewables system in the near-term. Enablers and Opportunities for the Renewables Transition The following subsections explore and analyse some of the factors which may contribute towards the implementation of a global transition to a renewables-based energy system in terms of systemic enablers and favourable conditions, and what may contribute to making it a true ‘stabilisation’.

2

THE TRIFURCATION OF ENERGY FUTURES

75

Degrowth and Changes to Per Capita Energy Use The global economy has undergone a prolonged period of nearuninterrupted growth, but the gains in terms of human welfare have been and continue to be heterogeneous at global scale. It is a central assumption of economic models that that global economy will continue to increase in magnitude over coming decades to enable a continued rise in human welfare for still-developing regions of the world, and to grow or at least maintain the general existing levels of wealth of already-developed regions. This assumption is central to the scenarios for a renewables transition described above, and many of the negative effects described (the requirement to re-materialise, to reduce the EROI of the global energy system and to potentially introduce collapse dynamics to the global system by expanding global renewables infrastructure) therefore result from this. As such, a key enabling factor for the renewables transition could be to reduce the total amount of material, infrastructure and capital required to underpin it by reducing the overall energy output required of the system. Increases in per capita energy usage have historically provided significant improvements to multiple metrics of quality of life, but there is a clear pattern of these gains reaching a plateau beyond a certain value. Societies which strive to continually grow per capita but which already have consumption-based, high-energy lifestyles are essentially wasting energy resources without substantially improving citizen’s lives. This situation exists in parallel to other societies globally having insufficient energy resources to meet quality of life metrics for many of their citizens. Therefore, there is a case for a convergence such that developed societies reduce unnecessary energy usage and less developed societies experiencing scarcity could increase theirs to attain quality of life improvements. The global Gini coefficient, a measure of income or wealth inequality, potentially needs to reduce by up to a factor of two [206]. A deliberate reduction in the energetic and material footprint of human civilisation at global scale would fall within the conceptual framework of ‘degrowth’, defined as “…voluntary and equitable downscaling of the economy towards a sustainable, just, and participatory steady-state society” which aims to make controlled changes to consumption before potential ‘hard’ resource constraints are encountered by global society [207, 208]. An energy consumption value of 79 GJ/capita/year has been calculated as sufficient to provide sufficiency living standards globally. This is approximately in line with global energy use in 1960, a time when many areas of the world experienced good performance across key health, economic

76

N. KING AND A. JONES

and environmental metrics. However, with increased energy efficiency since 1960 the actual quality of life obtained by this level of energy usage would be substantially higher than the quality of life experienced at that time. A framework which has been suggested for achieving this is ‘Donut Economics ’, which defines the ‘social foundation’ (12 basic human necessities including healthcare and education) and the ‘ecological ceiling’ (nine boundaries which define planetary overshoot including biodiversity loss and nitrogen/phosphorus loading). An economy is defined as prosperous under this model when the foundations are met for all people without exceeding any aspect of the ceiling (‘a safe and just space for humanity’); this has been assessed as attainable for many nations with existing levels of development and globalisation [209, 210]. A shift in energy usage of this type could potentially reduce overall future energy demands and could potentially be delivered by renewables systems with an EROI value commensurate with contemporary fossil fuels [211, 212]. This could potentially de-risk the transition to a renewablesbased energy system by diminishing the total physical extent of the system/infrastructure needed and therefore overall materials requirements. However, uncertainties over the extent of the degrowth which would be optimal (beyond just per capita energy figures) and the nature of the deep and radical systemic changes which would likely be required to enable transition to a smaller, steady state economic system, could present significant risks, and would likely be highly challenging to implement politically and socially [213–215]. One particular risk which has been suggested is that societies in a steady state or degrowing economic system could experience stagnation in which human wellbeing could be compromised. However, a key counterargument to that is that Earth’s biosphere has operated for its whole duration within an overall energy limit (total solar energy input) but has also achieved ever greater levels of complexity and diversity within that overarching constraint over time. This has been achieved through continuous reordering (via evolutionary processes) of its constituent mass along with incremental increases in complexity at multiple scales, as permitted by the expansive ‘phase space’ within the planetary solar constant [216] (noting however that there are ultimate constraints and limits to these complexity increases [217]). If human civilisation were to also exist within an equivalent overall energy constraint, it is reasonable to assume that similar increases in complexity, diversity and continuously available ‘variety and newness’ (in different ‘arenas’ such as social and technological) would

2

THE TRIFURCATION OF ENERGY FUTURES

77

in theory also be possible, allowing human societies to continue flourishing and similarly avoid stagnation under a degrowth and/or steady state paradigm. Renewables Circular Economy Recycling infrastructure for the redundant and end-of-life equipment and materials used in renewables systems has yet to be fully developed, and thus economic large-scale deployment of any such technologies remains in the future. However, significant technological steps and advances have been made that represent progress towards the development of the systems which would be required to start implementing circularity of resource use. Examples include chemical approaches for the recovery of epoxy resin materials from redundant wind turbine blades (which is able to produce virgin-grade material for new blades [218]), and targeted recovery of valuable materials (such as gold and rare earths) from redundant PV solar panels for resale, with an aim to make the whole recycling process more financially viable [219]. Advances such as these, working in aggregate and scaling with time, could potentially form the basis for future closedcycle systems that could transform the material basis of renewable energy systems. Interactions with the Food System The contemporary global food system has a high dependency on fossil fuel energy input (through both direct and indirect routes) and the large gains in global food production since the 1960s were underpinned to a large degree due to ramping up of this energetic input [220, 221]. This reliance on depleting fossil resources, along with the vulnerability of agricultural system to climate change (which it contributes to itself through emissions [222]) creates future risks (in line with those described in Chapter 1) such as the potential for large and potentially highly destabilising and risk-generating ‘calorie gaps’ to emerge where global demand start to outstrip production [223]. The renewables transition offers the potential for multi-faceted transformations of the food system, which could include: renewables providing direct energetic inputs to versatile, scalable and sustainable food production technologies such as precision fermentation (which have the potential to substantially displace highly damaging livestock agriculture globally) [224, 225]; and may also allow approaches such as the beneficial (rather

78

N. KING AND A. JONES

than exclusive as described above) co-location of solar PV and arable agricultural systems (‘agrivoltaics ’), which could provide mutual benefits for agriculture in certain climates by providing shade to enhance crop growth, whilst increasing the conversion efficiency of the panels by reducing their operating temperature [226]. Potential Hybrid Energy Systems A ‘hybrid’ energy system involves the concurrent operation of different generation technologies such that the different characteristics and strengths of the different technologies are leveraged (via tight coupling of the systems for particular applications) in order to gain overall benefit. In the context of a renewables-dominated system, this might primarily refer to provision of power for balancing purposes (to compensate for intermittency, alongside energy storage, overcapacity and demand management) and for non-electrical energy requirements, such as supply of high quality process heat (for e.g., hydrogen and steel production), desalination and district heating [227]. There are currently constraints and limitations to certain industrial processes being powered by renewables technologies, particularly the generation of high quality process heat [69], though progress has been made to date in decarbonising steelmaking (via electrical and hydrogenfired approaches, both of which could feasibly be underpinned by renewables technologies) and direct generation of lower-grade process heat via solar technologies [228, 229]. The relatively low TRL status of these approaches means that continued R&D is required to progress them to the stage at which they could be versatile, effective and scalable, and as such there could be scope for renewables systems to be supplemented by other technologies for niche applications on an interim or time-limited basis, to aid the transition to a fully renewable system. Given the risks and limitations associated with fossil fuel usage, any such hybrid system would involve the use of nuclear technology. Fusionbased systems would not be suitable due to the lack of near-term deployable technology, so the technologies which would have the greatest potential to fulfil this role are fission technologies. Candidates to fulfil this role are mature Gen III+ systems (recently-deployed examples such as the EPR have projected operational lifetimes of 60 years [230], so will in theory by available over a prolonged period and would be best suited to the system balancing application) and Gen IV systems (including SMRs/ AMRs) at higher TRLs which could be deployed at smaller scale in the

2

THE TRIFURCATION OF ENERGY FUTURES

79

near-term (approximately within the next decade). However, some of the Gen IV systems best suited to applications such as industrial process heat and hydrogen production (e.g., High Temperature Reactors, using novel thermochemical processes [96, 231]) are still at lower TRL. Overall, there could be benefits to operating a hybrid energy system to ensure certain specialist but vital (to the operation of technological civilisation, notably cement and steel production) processes are not curtailed or interrupted in the near-term by renewables technologies not having reached the level of maturity necessary to interchange with current fossil-fuel based systems (and also in a lesser system balancing role). The application of nuclear fission technologies in these roles would however likely only be desirable and effective if cost and availability risks could be kept low (i.e., if this was not reliant on systems at low TRL which would require large financial investments and time to bring to operation) and only for limited scale and duration (to ensure full stabilisation is ultimately achieved). Assessment of the Renewables Stabilisation Branch The overarching scenario presented in this branch is defined by human civilisation aligning its overall energy consumption around (likely well below) the limit set by (mainly) solar-powered environmental flows to become fully (or primarily) renewables-powered. This presents the potential for human civilisation to achieve long term stability (relative to exploitation of stock-based energy sources; natural analogues provide the steer for this), however significant complexities for its implementation have been identified. Systemic analysis using a specialist Integrated Assessment Model indicates that mineral, energy and land-use requirements of the transition could generate global economic recession as a result of energy scarcity and climate damages. The mineral requirements of the transition in particular represent a potential limit to the deployment of the necessary infrastructure, and waste management, ecological impacts and lack of credible new frontiers for renewables represent other significant constraints. Enablers and advantages from the renewables transition include significant cost reductions in the last decade, and the potential of tipping points to generate further reductions, direct food production and potential synergies with agriculture, and the potential for establishment of a recycling system to reduce the material intensity of the system.

80

N. KING AND A. JONES

The most significant opportunity to ameliorate the fundamental challenges associated with the renewables transition (notably materials supply) may be the reduction and convergence of global energy consumption to achieve economic degrowth (and eventually a steady state economy). This could potentially reduce the total infrastructure required such that hard limits are less apparent, or do not come into play. Overall, the Stabilisation Branch involves a number of significant potential complexities and challenges, including systemic constraints to achieving the transitions, but if these difficulties could be overcome and benefits and synergies realised, a renewables-based steady state energy system, combined with implementation of degrowth, could represent the optimal chance for human civilisation to attain long-term stability [232].

References 1. Zeebe, R. E., & Zachos, J. C. (2013). Long-term Legacy of Massive Carbon Input to the Earth System: Anthropocene Versus Eocene. Philosophical Transactions of the Royal Society A, 371, 20120006. 2. Pörtner, H.-O., Roberts, D. C., Adams, H., Adelekan, I., Adler, C., Adrian, R., Aldunce, P., Ali, E., Ara Begum, R., Bednar-Friedl, B., Bezner Kerr, R., Biesbroek, R., Birkmann, J., Bowen, K., Caretta, M. A., Carnicer, J., Castellanos, E., Cheong, T. S., Chow, W., … Zaiton Ibrahim, Z. (2022). IPCC Assessment Report 6: Technical Summary. In H.-O. Pörtner, D. C. Roberts, E. S. Poloczanska, K. Mintenbeck, M. Tignor, A. Alegría, M. Craig, S. Langsdorf, S. Löschke, V. Möller, A. Okem, & B. Rama (Eds.), Climate Change 2022: Impacts, Adaptation and Vulnerability. Contribution of Working Group II to the Sixth Assessment Report of the Intergovernmental Panel on Climate Change. Cambridge University Press. 3. Deng, J., Dai, A., & Xu, H. (2020). Nonlinear Climate Responses to Increasing CO2 and Anthropogenic Aerosols Simulated by CESM1. Journal of Climate, 33(1), 281–301. 4. Our World in Data. (2023). CO 2 Emissions. Available online: https://our worldindata.org/co2-emissions. Accessed 4 Feb 2023. 5. IPCC. (2021). Summary for Policymakers. In V. Masson-Delmotte, P. Zhai, A. Pirani, S. L. Connors, C. Péan, S. Berger, N. Caud, Y. Chen, L. Goldfarb, M. I. Gomis, M. Huang, K. Leitzell, E. Lonnoy, J. B. R. Matthews, T. K. Maycock, T. Waterfield, O. Yelekçi, R. Yu, & B. Zhou (Eds.), Climate Change 2021: The Physical Science Basis. Contribution of Working Group I to the Sixth Assessment Report of the Intergovernmental Panel on Climate Change. Cambridge University Press. 6. Etminan, G., Myhre, G., Highwood, E. J., & Shine, K. P. (2016). Radiative

2

7.

8.

9.

10.

11.

12.

13.

14.

15.

16.

17.

THE TRIFURCATION OF ENERGY FUTURES

81

Forcing of Carbon Dioxide, Methane, and Nitrous Oxide: A significant Revision of the Methane Radiative Forcing. Geophysical Research Letters, 43(24), 12614–12623. Song, F., Zhang, G. J., Ramanathan, V., & Leung, L. R. (2022). Trends in Surface Equivalent Potential Temperature: A More Comprehensive Metric for Global Warming and Weather Extremes. Proceedings of the National Academy of Sciences, 119(6), e2117832119. Kaufman, D. S., & McKay, N. P. (2022). Technical Note: Past and Future Warming—Direct Comparison on Multi-century Timescales. Climate of the Past, 18(4), 911–917. Diffenbaugh, N. S., & Barnes, E. A. (2023). Data-Driven Predictions of the Time Remaining Until Critical Global Warming Thresholds Are Reached. Proceedings of the National Academy of Sciences, 120(6), e2207183120. McKay, D. I. A., Staal, A., Abrams, J. F., Winkelmann, R., Sakschewski, B., Loriani, S., Fetzer, I., Cornell, S. E., Rockström, J., & Lenton, T. M. (2022) Exceeding 1.5°C Global Warming Could Trigger Multiple Climate Tipping Points. Science, 377 (6611), eabn7950. Klose, A. K., Wunderling, N., Winkelmann, R., & Donges, J. F. (2021). What Do We Mean, ‘Tipping Cascade’? Environmental Research Letters, 16, 125011. Ripple, W. J., Wolf, C., Lenton, T. M., Gregg, J. W., Natali, S. M., Duffy, P. B., Rockström, J., & Shellnhuber, H. J. (2023). Many Risky Feedback Loops Amplify the Need for Climate Action. One Earth, 6(2), 86–91. Prentice, I. C., Williams, S., & Friedlingstein, P. (2015). Biosphere Feedbacks and Climate Change. Grantham Institute, Briefing Paper No. 12, Imperial College London. Jehn, F. U., Kemp, L., Ilin, E., Funk, C., Wang, J. R., & Breuer, L. (2022). Focus of the IPCC Assessment Reports Has Shifted to Lower Temperatures. Earth’s Future, 10(5), e2022EF002876. Hansen, J. E., Sato, M., Simons, L., Nazarenko, L. S., von Schuckmann, K., Loeb, N. G., Osman, M. B., Kharecha, P., Jin, Q., Tselioudis, G., Lacis, A., Ruedy, R., Russell, G., Cao, J., & Lo, J. (2022). Global Warming in the Pipeline. arXiv:2212.04474v2 Murakami, H. (2022). Substantial Global Influence of Anthropogenic Aerosols on Tropical Cyclones over the Past 40 Years. Science Advances, 8(19), eabn9493. Watson-Parris, D., Christensen, M. W., Laurensen, A., Clewly, D., Gryspeerdt, E., & Stier, P. (2022). Shipping Regulations Lead to Large Reduction in Cloud Perturbations. Proceedings of the National Academy of Sciences, 119(41), e2206885119.

82

N. KING AND A. JONES

18. Yuan, T., Song, H., Wood, R., Wang, C., Oreopoulos, L., Platnick, S. E., von Hippel, S., Meyer, K., Liight, S., & Wilcox, E. (2022). Global Reduction in Ship-Tracks from Sulfur Regulations for Shipping fuel. Science Advances, 8(29), eabn7988. 19. Nair, H. R. C. R., Budhavant, K., Manoj, M. R., Andersson, A., Satheesh, S. K., Ramanathan, V., & Gustafsson, Ö. (2023). Aerosol Demasking Enhances Climate Warming over South Asia. NPJ Climate and Atmospheric Science, 6(1), 39. 20. Warszawski, L., Kriegler, E., Lenton, T. M., Gaffney, O., Jacob, D., Klingenfeld, D., Koide, R., Costa, M. M., Messner, D., & Nakicenovic, N. (2021). All Options, Not Silver Bullets, Needed to Limit Global Warming to 1.5 °C: A Scenario Appraisal. Environmental Research Letters, 16, 064037. 21. Laybourn, L., Throp, H., & Sherman, S. (2018). 1.5°C—Dead or Alive?: The Risks to Transformational Change from Reaching and Breaching the Paris Agreement Goal. IPPR and Chatham House. Available online: http://www.ippr.org/research/publications/1-5c-dead-or-alive. Accessed 21 Feb 2023. 22. Zeebe, R. E., Ridgwell, A., & Zachos, J. C. (2016). Anthropogenic Carbon Release Rate Unprecedented During the Past 66 Million Years. Nature Geoscience, 9, 325–329. 23. Doughty, C. E., Keany, J. M., Wiebe, B. C., Rey-Sanchez, C., Carter, K. R., Middleby, K. B., Cheesman, A. W., Goulden, M. L., da Rocha, H. R., Miller, S. D., Malhi, Y., Fauset, S., Gloor, E., Slot, M., Menor, I. O., Crous, K. Y., Goldsmith, G. R., & Fisher, J. B. (2023). Tropical Forests Are Approaching Critical Temperature Thresholds. Nature. 24. Halsch, C. A., Shapiro, A. M., Fordyce, J. A., Nice, C. C., Thorne, J. H., Waetjen, D. P., & Forister, M. L. (2021). Insects and Recent Climate Change. Proceedings of the National Academy of Sciences, 118(2), e2002543117. 25. Ball, P. (2003). Gamma-ray Burst Linked to Mass Extinction. Nature. 26. Melott, A. L., Marinho, F., & Paulucci, L. (2019). Hypothesis: Muon Radiation Dose and Marine Megafaunal Extinction at the End-Pliocene Supernova. Astrobiology, 19(6), 825–830. 27. Song, H., Kemp, D. B., Tian, L., Chu, D., Song, H., & Dai, X. (2021). Thresholds of Temperature Change for Mass Extinctions. Nature Communications, 12(1), 4694. 28. Kornhuber, K., Lesk, C., Schleussner, C. F., Jägermeyr, J., Pfleiderer, P., & Horton, R. M. (2023). Risks of Synchronized Low Yields Are Underestimated in Climate and Crop Model Projections. Nature Communications, 14(1), 3528.

2

THE TRIFURCATION OF ENERGY FUTURES

83

29. Kornhuber, K., Coumou, D., Vogel, E., Lesk, C., Donges, J. F., Lehmann, J., & Horton, R. M. (2019). Amplified Rossby Waves Enhance Risk of Concurrent Heatwaves in Major Breadbasket Regions. Nature Climate Change, 10, 48–53. 30. Dasgupta, S., & Robinson, E. J. Z. (2022). Attributing Changes in Food Insecurity to a Changing Climate. Scientific Reports, 12(1), 4709. 31. Christian, J. I., Martin, E. R., Basara, J. B., Furtado, J. C., Otkin, J. A., Lowman, L. E. L., Hunt, E. D., Mishra, V., & Xiao, X. (2023). Global Projections of Flash Drought Show Increased Risk in a Warming Climate. Communications Earth & Environment, 4(1), 165. 32. Rogers, C. D. W., Kornhuber, K., Perkins-Kirkpatrick, S. E., Loikith, P. C., & Singh, D. (2022). Sixfold Increase in Historical Northern Hemisphere Concurrent Large Heatwaves Driven by Warming and Changing Atmospheric Circulations. AMS Journal of Climate, 35(3), 1063–1078. 33. Xu, Y., Chu, C., & Yao, S. (2021). The Impact of High-Temperature Stress on Rice: Challenges and Solutions. The Crop Journal, 9(5), 963–976. 34. Ditlevsen, P., & Ditlevsen, S. (2023). Warning of a Forthcoming Collapse of the Atlantic Meridional Overturning Circulation. Nature Communications, 14. Article Number: 4254. 35. Qing, Y., Wang, S., Yang, Z.-L., & Gentine, P. (2023). Soil MoistureAtmosphere Feedbacks Have Triggered the Shifts from Drought to Pluvial Conditions Since 1980. Communications Earth and Environment, 4, 254. 36. Bailey, R., & Wellesley, L. (2017). Chokepoints and Vulnerabilities in Global Food Trade. Chatham House Report, Royal Institute of International Affairs. 37. Walton, R., Miller, J. O., & Champagne, L. (2019). Simulating Maritime Chokepoint Disruption in the Global Food Supply. In Winter Simulation Conference (WSC) (pp. 1708–1718). 38. McKinsey Global Institute. (2020). Will Infrastructure Bend or Break under Climate Stress? Available online: https://www.mckinsey.com/capabi lities/sustainability/our-insights/will-infrastructure-bend-or-break-underclimate-stress. Accessed 15 Dec 2022. 39. Kinnunen, P., Guillaume, J. H. A., Taka, M., D’Odorico, P., Siebert, S., Puma, M. J., Jalava, M., & Kummu, M. (2020). Local Food Crop Production Can Fulfil Demand for Less Than One-Third of the Population. Nature Food, 1, 229–237. 40. Forzieri, G., Bianchi, A., Silva, F. B. E., Herrera, M. A. M., Leblois, A., Lavalle, C., Aerts, J. C. J. H., & Feyen, L. (2018). Escalating Impacts of Climate Extremes on Critical Infrastructures in Europe. Global Environmental Change, 48, 97–107.

84

N. KING AND A. JONES

41. Steel, D., DesRoches, C. T., & Mintz-Woo, K. (2022). Climate Change and the Threat to Civilization. Proceedings of the National Academy of Sciences, 119(42), e2210525119. 42. Vox. (2023). Climate Change Is Already Making Parts of America Uninsurable. Available online: https://www.vox.com/climate/23746045/statefarm-california-climate-change-insurance-wildfire-florida-flood. Accessed 1 Aug 2023. 43. Xu, C., Kohler, T. A., Lenton, T. M., Svenning, J.-C., & Scheffer, M. (2020). Future of the Human Climate Niche. Proceedings of the National Academy of Sciences, 117 (21), 11350–11355. 44. Lenton, T. M., Xu, C., Abrams, J. F., Ghadiali, A., Loriani, S., Sakschewski, B., Zimm, C., Ebi, K. L., Dunn, R. R., Svenning, J.-C., & Scheffer, M. (2023). Quantifying the Human Cost of Global Warming. Nature Sustainability, 6, 1237–1247. 45. Kroeger, C. (2023). Heat Is Associated with Short-Term Increases in Household Food Insecurity in 150 Countries and This Is Mediated By Income. Nature Human Behaviour, 7 , 1777–1786. 46. Willcock, S., Cooper, G. S., Addy, S., & Dearing, J. A. (2023). Earlier Collapse of Anthropocene Ecosystems Driven by Multiple Faster and Noisier Drivers. Nature Sustainability. 47. Rocha, J. C., Peterson, G., Bodin, Ö., & Levin, S. (2018). Cascading Regime Shifts Within and Across Scales. Science, 362(6421), 1379–1383. 48. King, A. D., & Harrington, L. J. (2018). The Inequality of Climate Change from 1.5 to 2°C of Global Warming. Geophysical Research Letters, 45(10), 5030–5033. 49. Ripple, W. J., Wolf, C., Gregg, J. W., Levin, K., Rockström, J., Newsome, T. M., Betts, M. G., Huq, S., Law, B. E., Kemp, L., Kalmus, P., & Lenton, T. M. (2022). World Scientists’ Warning of a Climate Emergency 2022. BioScience, 72(12), 1149–1155. 50. Bradshaw, C. J. A., Ehrlich, P. R., Beattie, A., Ceballos, G., Crist, E., Diamond, J., Dirzo, R., Ehrlich, A. H., Harte, J., Harte, M. E., Pyke, G., Raven, P. H., Ripple, W. J., Saltré, F., Turnball, C., Wackernagal, M., & Blumstein, D. T. (2021). Underestimating the Challenges of Avoiding a Ghastly Future. Frontiers in Conservation Science, 1, 615419. 51. de Castro, C., Miguel, L. J., & Mediavilla, M. (2009). The Role of Non Conventional Oil in the Attenuation of Peak Oil. Energy Policy, 37 (5), 1825–1833. 52. Eveloy, V., & Elshiek, H. (2022). Unconventional Oil Prospects and Challenges in the Covid-19 Era. Frontiers in Energy Research, 10, 829398. 53. Brecha, R. J. (2013). Ten Reasons to Take Peak Oil Seriously. Sustainability, 5(2), 664–694.

2

THE TRIFURCATION OF ENERGY FUTURES

85

54. Jaccard, M. (2020). The Citizen’s Guide to Climate Success—Overcoming Myths That Hinder Progress. Cambridge University Press. 55. Court, V., & Fizaine, F. (2017). Long-Term Estimates of the EnergyReturn-on-Investment (EROI) of Coal, Oil, and Gas Global Productions. Ecological Economics, 138, 145–159. 56. Bardi, U. (2014). Extracted: How the Quest for Mineral Wealth Is Plundering the Planet—The Past, Present, and Future of Global Mineral Depletion: A Report to the Club of Rome. Chelsea Green Publishing. 57. Seppelt, R., Manceur, A. M., Liu, J., Fenichel, E. P., & Klotz, S. (2014). Synchronized Peak-Rate Years of Global Resources Use. Ecology and Society, 19(4), 50. 58. World Economic Forum. (2023). Global Risks Report 2023—3. Resource Rivalries: Four Emerging Futures. Available online: https://www.wef orum.org/reports/global-risks-report-2023/in-full/3-resource-rivalriesfour-emerging-futures/. Accessed 26 July 2023. 59. Qvist, S. A., & Brook, B. W. (2015). Potential for Worldwide Displacement of Fossil-Fuel Electricity by Nuclear Energy in Three Decades Based on Extrapolation of Regional Deployment Data. PLoS ONE, 10(5), e0124074. 60. United States Atomic Energy Commission. (1954). Remarks Prepared by Lewis L. Strauss, Chairman, United States Atomic Energy Commission, for Delivery at the Founder’s Day Dinner, National Associated of Science Writers, on Thursday, September 16, 1954. Available online: https://www. nrc.gov/docs/ML1613/ML16131A120.pdf. Accessed 17 Jan 2023. 61. Mycle Schneider Consulting. (2022). The World Nuclear Industry Status Report 2022. Available online: https://www.worldnuclearreport. org/IMG/pdf/wnisr2022-lr.pdf. Accessed 18 Jan 2023. 62. Our World in Data. (2022). Nuclear Energy. Available online: https://our worldindata.org/nuclear-energy. Accessed 21 July 2023. 63. Eash-Gates, P., Klemun, M. M., Kavlak, G., McNerney, J., Buongiorno, J., & Trancik, J. E. (2020). Sources of Cost Overrun in Nuclear Power Plant Construction Call for a New Approach to Engineering Design. Joule, 4, 2348–2373. 64. J. P. Morgan Private Bank. (undated). The Rising Cost of Nuclear Power. Available online: https://privatebank.jpmorgan.com/content/dam/jpmwm-aem/global/pb/en/insights/eye-on-the-market/the-rising-cost-ofnuclear-power.pdf. Accessed 17 Jan 2023. 65. Hultman, N., & Koomey, J. (2013). Three Mile Island: The Driver of US Nuclear Power’s Decline? Bulletin of the Atomic Scientists, 69(3), 63–70. 66. Von Hippel, F. (2016). Chernobyl Didn’t Kill Nuclear Power. Scientific American, 314(4), 11.

86

N. KING AND A. JONES

67. Hacquin, A.-S., Altay, S., Aarøe, L., & Mercier, H. (2022). Disgust Sensitivity and Public Opinion on Nuclear Energy. Journal of Environmental Psychology, 80, 101749. 68. Kim, Y., Kim, W., & Kim, M. (2014). An International Comparative Analysis of Public Acceptance of Nuclear Energy. Energy Policy, 66, 475–483. 69. Haywood, L., Leroutier, M., & Pietzcker, R. (2023). Why Investing in New Nuclear Plants Is Bad for the Climate. Joule, 7 (8), 1675–1678. 70. International Atomic Energy Agency. (2023). Global Status of Decommissioning of Nuclear Installations. Available online: https://www-pub. iaea.org/MTCD/Publications/PDF/PUB2023_web.pdf. Accessed 31 July 2023. 71. Faybishenko, B., Birkholzer, J., Sassani, D., & Swift, P. (2016). International Approaches for Nuclear Waste Disposal in Geological Formations: Geological Challenges in Radioactive Waste Isolation—Fifth Worldwide Review. Available online: https://www.osti.gov/servlets/purl/1353043/. Accessed 31 July 2023. 72. Swift, P., & Newman, A. (2022). Deep Disposal of Radioactive Waste: Next Steps and Applicability to National Programs. Columbia University Center on Global Energy Policy. 73. King, N., & Jones, A. (2020). An Assessment of Civil Nuclear ‘Enabling’ and ‘Amelioration’ Factors for EROI Analysis. Sustainability, 12(20), 8414. 74. Nuclear Engineering International. (2018). Pressuring Light Water. WNE Special Edition—New Technology: Encouraging Excellence in Nuclear Energy. NEI Magazine. 75. Fouquet, R. (2016). Path Dependence in Energy Systems and Economic Development. Nature Energy, 1, 16098. 76. A Fine Theorem. (2011). Nuclear Power Reactors: A Study in Technological Lock-In, R. Cowan (1990). Available online: https://afinetheorem.wordpr ess.com/2011/11/16/nuclear-power-reactors-a-study-in-technologicallock-in-r-cowan-1990/. Accessed 17 Jan 2023. 77. Bunn, M., Holdren, J. P., Fetter, S., & van der Zwaan, B. (2017). The Economics of Reprocessing versus Direct Disposal of Spent Nuclear Fuel. Nuclear Technology, 150(3), 209–230. 78. Calvo, G., Mudd, G., Valero, A., & Valero, A. (2016). Decreasing Ore Grades in Global Metallic Mining: A Theoretical Issue or a Global Reality? Resources, 5(4), 36. 79. Hall, S., Coleman, M. (2013). Critical Analysis of World Uranium Resources: U.S. Scientific Investigations Report 2012–5239 (56pp.). USGS. 80. Dittmar, M. (2016). The End of Cheap Uranium. Science of the Total Environment, 1(461–462), 792–798.

2

THE TRIFURCATION OF ENERGY FUTURES

87

81. Abbott, D. (2011). Is Nuclear Power Globally Scalable? Proceedings of the IEEE, 99(10), 1611–1617. 82. Muellner, N., Arnold, N., Gufler, K., Kromp, W., Renneberg, W., & Liebert, W. (2021). Nuclear Energy—The Solution to Climate Change? Energy Policy, 155, 112363. 83. McKay, D. J. C. (2009). Sustainable Energy—Without the Hot Air. UIT. 84. Yuan, Y., Yu, Q., Cao, M., Feng, L., Feng, S., Liu, T., Feng, T., Yan, B., Guo, Z., & Wang, N. (2021). Selective Extraction of Uranium from Seawater with Biofouling-Resistant Polymeric Peptide. Nature Sustainability, 4, 708–714. 85. Guidez, J., & Gabriel, S. (2015). Extraction of Uranium from Seawater: A Few Facts. EPJ Nuclear Sciences & Technologies, 2, Article Number 10. 86. Dungan, K., Butler, G., Livens, F. R., & Warren, L. M. (2017). Uranium from Seawater—Infinite Resource or Improbable Aspiration? Progress in Nuclear Energy, 99, 81–85. 87. Hyatt, N. C. (2020). Safe Management of the UK Separated Plutonium Inventory: A Challenge of Materials Degradation. NPJ Materials Degradation, 4, Article number: 28. 88. American Nuclear Society. (2022). Another Delay, Cost Bump, for Flamanville-3. Available online: https://www.ans.org/news/article-3573/ another-delay-cost-bump-for-flamanville3/. Accessed 31 July 2023. 89. AP News. (2023). Georgia Nuclear Rebirth Arrives 7 Years Late, $17B over Cost. Available online: https://apnews.com/article/georgia-nuclearpower-plant-vogtle-rates-costs-75c7a413cda3935dd551be9115e88a64. Accessed 1 Aug 2023. 90. Gen IV International Forum. (2023). Generation IV Systems. Available online: https://www.gen-4.org/gif/jcms/c_59461/generation-iv-systems. Accessed 18 Jan 2023. 91. Nuclear Engineering International. (2022). Reinvented: The Dual Fluid Principle. Available online: https://www.neimagazine.com/features/featur ereinvented-the-dual-fluid-principle-9952766/. Accessed 18 Jan 2023. 92. Humphrey, U. E., & Khandaker, M. U. (2018). Viability of ThoriumBased Nuclear Fuel Cycle for the Next Generation Nuclear Reactor: Issues and Prospects. Renewable and Sustainable Energy Reviews, 97 , 259–275. 93. Cochran, T. B., Feiveson, H. A., Mian, Z., Ramana, M. V., Schneider, M., & von Hippel, F. N. (2010). It’s Time to Give Up on Breeder Reactors. Bulletin of the Atomic Scientists, 66(3), 50–56. 94. World Nuclear Association. (2022). Small Nuclear Power Reactors. Available online: https://www.world-nuclear.org/information-library/nuc lear-fuel-cycle/nuclear-power-reactors/small-nuclear-power-reactors.aspx. Accessed 18 Jan 2023.

88

N. KING AND A. JONES

95. Stewart, W. R., & Shirvan, K. (2022). Capital Cost Estimation for Advanced Nuclear Power Plants. Renewable and Sustainable Energy Reviews, 155, 111880. 96. Peakman, A., & Merk, B. (2019). The Role of Nuclear Power in Meeting Current and Future Industrial Process Heat Demands. Energies, 12(19), 3664. 97. Teräsvirta, A., Syri, S., & Hiltunen, P. (2020). Small Nuclear Reactor— Nordic District Heating Case Study. Energies, 13(15), 3782. 98. Yan, B. H., Wang, C., & Li, L. G. (2020). The Technology of Micro Heat Pipe Cooled Reactor: A Review. Annals of Nuclear Energy, 135, 106948. 99. Horizon 2020. (2014). Technology Readiness Levels. Available online: https://ec.europa.eu/research/participants/data/ref/h2020/wp/2014_2 015/annexes/h2020-wp1415-annex-g-trl_en.pdf. Accessed 31 July 2023. 100. NucNet. (2020). Generation IV / The Key Challenges in the Race for Commercialisation. Available online: https://www.nucnet.org/news/thekey-challenges-in-the-race-for-commercialisation-2-3-2020. Accessed 6 Feb 2023. 101. Krall, L. M., Macfarlane, A. M., & Ewing, R. C. (2022). Nuclear Waste from Small Modular Reactors. Proceedings of the National Academy of Sciences, 119(23), e2111833119. 102. Nuclear Energy Agency. (2021). Small Modular Reactors: Challenges and Opportunities. Available online: https://www.oecd-nea.org/upload/docs/ application/pdf/2021-03/7560_smr_report.pdf. Accessed 31 July 2023. 103. Höök, M., & Tang, X. (2013). Depletion of Fossil Fuels and Anthropogenic Climate Change—A Review. Energy Policy, 52, 797–809. 104. Bardi, U. (2017). The Seneca Effect: Why Growth Is Slow but Collapse is Rapid. Springer. 105. Brutschin, E., Cherp, A., & Jewell, J. (2021). Failing the Formative Phase: The Global Diffusion of Nuclear Power Is Limited by National Markets. Energy Research & Social Science, 80, 102221. 106. Weyler, R. (2017). Nuclear Power and the Collapse of Society. Available online: https://www.greenpeace.org/international/story/47815/nuclearpower-and-the-collapse-of-society/. Accessed 20 Aug 2023. 107. King, N., & Jones, A. (2021). An Analysis of the Potential for the Formation of ‘Nodes of Persisting Complexity.’ Sustainability, 13(15), 8161. 108. Bolson, N., Yutkin, M., Rees, W., & Patzek, T. (2022). Resilience Rankings and Trajectories of World’s Countries. Ecological Economics, 195, 107383. 109. Rees, W. E. (2023). The Human Ecology of Overshoot: Why a Major ‘Population Correction’ Is Inevitable. World, 4(3), 509–527. 110. Gopalakrishnan, T., & Miller, J. (2022). Energy Transition in Fragile States: A Critical Primer. Tufts University.

2

THE TRIFURCATION OF ENERGY FUTURES

89

111. Pasqualino, R., & Jones, A. W. (2020). Resources, Financial Risk and the Dynamics of Growth – Systems and Global Society. Routledge. 112. Crutchfield, P. (2022). Societal Collapse and Intergenerational Disparities in Suffering. Neuroethics, 15(3), 27. 113. Pearce, J. M., & Parncutt, R. (2023). Quantifying Global Greenhouse Gas Emissions in Human Deaths to Guide Energy Policy. Energies, 16(16), 6074. 114. Kemp, L., Xu, C., Depledge, J., Ebi, K. L., Gibbins, G., Kohler, T. A., Rockström, J., Scheffer, M., Schellnhuber, H. J., Steffen, W., & Lenton, T. M. (2022). Climate Endgame: Exploring Catastrophic Climate Change Scenarios. Proceedings of the National Academy of Sciences, 119(34), e2108146119. 115. Karl, T. R., & Trenberth, K. E. (2003). Modern Global Climate Change. Science, 302(5651), 1719–1723. 116. Possner, A., & Caldeira, K. (2017). Geophysical Potential for Wind Energy over the Open Oceans. Proceedings of the National Academy of Sciences, 114(43), 11338–11343. 117. LibreTexts Geosciences. (undated). Abundance of Elements in Earth’s Crust. Available online: https://geo.libretexts.org/Bookshelves/Geo logy/Fundamentals_of_Geology_(Schulte)/02%3A_Rock_Forming_Min erals/2.08%3A_Abundance_of_Elements_in_Earth’s_Crust. Accessed 1 Aug 2023. 118. World Nuclear Association. (2022). Nuclear Fusion Power. Available online: https://world-nuclear.org/information-library/current-and-futuregeneration/nuclear-fusion-power.aspx. Accessed 7 Feb 2023. 119. Butt, F. S., Lewis, A., Chen, T., Mazlan, N. A., Wei, X., Hayer, J., Chen, S., Han, J., Yang, Y., Yang, S., & Huang, Y. (2022). Lithium Harvesting from the Most Abundant Primary and Secondary Sources: A Comparative Study on Conventional and Membrane Technologies. Membranes, 12(4), 373. 120. Lawrence Livermore National Laboratory. (2022). National Ignition Facility Achieves Fusion Ignition. Available online: https://www.llnl.gov/ news/national-ignition-facility-achieves-fusion-ignition. Accessed 7 Feb 2023. 121. Fusion Industry Association. (2022). The Global Fusion Industry in 2022. Available online: https://www.fusionindustryassociation.org/about-fusionindustry. Accessed 8 Feb 2023. 122. Cho, A. (2014). Cost Skyrockets for United States’ Share of ITER Fusion Project. Available online: https://www.science.org/content/article/costskyrockets-united-states-share-iter-fusion-project. Accessed 8 Feb 2023.

90

N. KING AND A. JONES

123. Surrey, E. (2019). Engineering Challenges for Accelerated Fusion Demonstrators. Philosophical Transactions of the Royal Society A, 377 (2141), 20170442. 124. Hesch, K., Boccaccini, L. V., & Stieglitz, R. (2018). Blankets—Key Element of a Fusion Reactor—Functions, Design and Present State of Development. Kerntechnik, 83, 3. 125. World Nuclear News. (2023). ITER to Replace 23km of Cooling Pipes on Thermal Shield Panels. Available online: https://www.world-nuclearnews.org/Articles/ITER-to-replace-23km-of-cooling-pipes-on-thermal-s. Accessed 8 Feb 2023. 126. Bulletin of the Atomic Scientists. (2017). Fusion Reactors: Not What They’re Cracked Up to Be. Available online: https://thebulletin.org/2017/04/fus ion-reactors-not-what-theyre-cracked-up-to-be/. Accessed 8 Feb 2023. 127. Bulletin of the Atomic Scientists. (2018). ITER Is a Showcase … for the Drawbacks of Fusion Energy. Available online: https://thebulletin. org/2018/02/iter-is-a-showcase-for-the-drawbacks-of-fusion-energy/. Accessed 9 Mar 2023. 128. Nuckolls, J. H. (2010). Grand Challenges of Inertial Fusion Energy. Journal of Physics: Conference Series, 244, 012007. 129. Huang, F., & Meng, C. (2011). Method for the Production of DeuteriumDepleted Potable Water. Industrial & Engineering Chemical Research, 50(1), 378–381. 130. Giegerich, T., Battes, K., Schwenzer, J. C., & Day, C. (2019). Development of a Viable Route for Lithium-6 Supply of DEMO and Future Fusion Power Plants. Fusion Engineering and Design, 149, 111339. 131. Sawan, M. E., & Abdou, M. A. (2006). Physics and Technology Conditions for Attaining Tritium Self-Sufficiency for the DT Fuel Cycle. Fusion Engineering and Design, 81(8–14), 1131–1144. 132. Clery, D. (2022). Out of Gas—A Shortage of Tritium Fuel May Leave Fusion Energy with an Empty Tank. Science, 376(6600). 133. Aghahosseine, A., & Breyer, C. (2020). From Hot Rock to Useful Energy: A Global Estimate of Enhanced Geothermal Systems Potential. Applied Energy, 279, 115769. 134. Sovacool, B. K., Schmid, P., Stirling, A., Walter, G., & MacKerron, G. (2020). Differences in Carbon Emissions Reduction Between Countries Pursuing Renewable Electricity Versus Nuclear Power. Nature Energy, 5, 928–935. 135. He, G., Lin, J., Sifuentes, F., Liu, X., Abhyankar, N., & Phadke, A. (2020). Rapid Cost Decrease of Renewables and Storage Accelerates the Decarbonization of China’s Power System. Nature Communications, 11(1), 2486.

2

THE TRIFURCATION OF ENERGY FUTURES

91

136. Mercure, J.-F. (2012). FTT: Power: A Global Model of the Power Sector with Induced Technological Change and Natural Resource Depletion. Energy Policy, 48, 799–811. 137. Xiao, M., Junne, T., Haas, J., & Klein, M. (2021). Plummeting Costs of Renewables—Are Energy Scenarios Lagging? Energy Strategy Reviews, 35, 100636. 138. SystemIQ. (2023). The Breakthrough Effect: How to Trigger a Cascade of Tipping Points to Accelerate the Net Zero Transition. Available online: https://www.systemiq.earth/wp-content/uploads/2023/01/TheBreakthrough-Effect.pdf. Accessed 6 Mar 2023. 139. Ghiassi-Farrokhfal, Y., Keshav, S., & Rosenberg, C. (undated). An EROIBased Analysis of Renewable Energy Farms with Storage. Available online: https://ece.uwaterloo.ca/~cath/eenergy14.pdf. Accessed 17 Mar 2023. 140. Chanut, N., Stefaniuk, D., Weaver, J. C., Zhu, Y., Shao-Horn, Y., Masic, A., & Ulm, F.-J. (2023). Carbon-Cement Supercapacitors as a Scalable Bulk Energy Storage Solution. Proceedings of the National Academy of Sciences, 120(32), e2304318120. 141. Our World in Data. (2023). World Population Living in Extreme Poverty, World, 1820 to 2015. Available online: https://ourworldindata.org/ grapher/world-population-in-extreme-poverty-absolute. Accessed 15 Feb 2023. 142. Hickel, J. (2017). Is Global Inequality Getting Better or Worse? A Critique of the World Bank’s Convergence Narrative. Third World Quarterly, 38(10), 2208–2222. 143. Hickel, J., Dorninger, C., Wieland, H., & Suwandi, I. (2022). Imperialist Appropriation in the World Economy: Drain from the Global South Through Unequal Exchange, 1990–2015. Global Environmental Change, 73, 102467. 144. Hickel, J., & Sullivan, D. (2023). Capitalism, Global Poverty, and the Case for Democratic Socialism. Available online: https://monthlyreview.org/ 2023/07/01/capitalism-global-poverty-and-the-case-for-democratic-social ism/. Accessed 20 Aug 2023. 145. Oxfam. (2023). Survival of the Richest: How We Must Tax the Super-Rich Now to Fight Inequality. Available online: https://oxfamilibrary.openrepos itory.com/bitstream/handle/10546/621477/bp-survival-of-the-richest160123-en.pdf. Accessed 20 Feb 2023. 146. Bastani, A. (2020). Fully Automated Luxury Communism—A Manifesto. Verso Books. 147. British Broadcasting Corporation. (2022). How Limitless Green Energy Would Change the World. Available online: https://www.bbc.com/future/ article/20221006-what-would-happen-if-we-had-limitless-green-energy. Accessed 15 Feb 2023.

92

N. KING AND A. JONES

148. Future Business Tech. (2022). What If We Had Unlimited Electricity? Available online: https://www.futurebusinesstech.com/blog/what-if-wehad-unlimited-electricity. Accessed 16 Feb 2023. 149. Motesharrei, S., Rivas, J., & Kalnay, E. (2014). Human and Nature Dynamics (HANDY): Modeling Inequality and Use of Resources in the Collapse or Sustainability of Societies. Ecological Economics, 101, 90–102. 150. Zalasiewicz, J., Williams, M., Waters, C. N., Barnosky, A. D., Palmesino, J., Rönnskog, A., Edgeworth, M., Neal, C., Cearreta, A., Ellis, E. C., Grinevald, J., Haff, P., Ivar do Sul, J. A., Jeandel, C., Leinfelder, R., McNeill, J. R., Odada, E., Oreskes, N., Price, S. J., … Wolfe, A. P. (2017). Scale and Diversity of the Physical Technosphere: A Geological Perspective. The Anthropocene Review, 4(1), 9–22. 151. Elhacham, E., Ben-Uri, L., Grozovski, J., Bar-On, Y. M., & Milo, R. (2020). Global Human-Made Mass Exceeds All Living Biomass. Nature, 588(7838), 442–444. 152. Filella, M., Turner, A., & Arp, H. P. H. (2021). Editorial: Plastics in the Environment: Understanding Impacts and Identifying Solutions. Frontiers in Environmental Science, 9, 699971. 153. Plumptre, A. J., Baisero, D., Belote, R. T., Vázquez-Domínguez, E., Faurby, S., J¸edrzejewski, W., Kiara, H., Kühl, H., Benítez-López, A., LunaAranguré, C., Voigt, M., Wich, S., Wint, W., Gallego-Zamorano, J., & Boyd, C. (2021). Where Might We Find Ecologically Intact Communities? Frontiers in Forests and Global Change, 4, 626635. 154. World Wildlife Fund. (2022). Living Planet Report 2022—Building a Nature-Positive Society. In R. E. A. Almond, M. Grooten, D. Juffe Bignoli & T. Petersen (Eds.). WWF. 155. University of Pennsylvania. (2014). Things Fall Apart: How Social Media Leads to a Less Stable World. Available online: https://knowledge.wharton. upenn.edu/article/how-social-media-leads-to-a-less-stable-world. Accessed 17 Feb 2023. 156. Scientific American. (2020). Information Overload Helps Fake News Spread, and Social Media Knows It. Available online: https://www.scientificameri can.com/article/information-overload-helps-fake-news-spread-and-socialmedia-knows-it. Accessed 17 Feb 2023. 157. Harari, Y. N. (2015). Sapiens: A Brief History of Humankind. Vintage. 158. Doya, K., Ema, A., Kitano, H., Sakagami, M., & Russell, S. (2022). Social Impact and Governance of AI and Neurotechnologies. Neural Networks, 152, 542–554. 159. Smith, R. (2023). Finding Paths from Our Predicament of a Poisoned World. BMJ, 380, 342.

2

THE TRIFURCATION OF ENERGY FUTURES

93

160. Kleidon, A., Malhi, Y., & Cox, P. M. (2010). Maximum Entropy Production in Environmental and Ecological Systems. Philosophical Transactions of the Royal Society B, 365(1545), 1297–1302. 161. Conflict and Environment Observatory. (2014). Conflict Rubble: A Ubiquitous and Under-Studied Toxic Remnant of War. Available online: https://ceobs.org/conflict-rubble-a-ubiquitous-and-under-stu died-toxic-remnant-of-war/. Accessed 3 Aug 2023. 162. Newman, M. E. J. (2005). Power Laws, Pareto Distributions and Zipf’s Law. Contemporary Physics, 46(5), 323–351. 163. Clauset, A. (2019). On the Frequency and Severity of Interstate Wars. In N. P. Gleditsch (Eds.), Lewis Fry Richardson: His Intellectual Legacy and Influence in the Social Sciences. Pioneers in Arts, Humanities, Science, Engineering, Practice (Vol. 27). Springer. 164. Becerra, Ó., Johnson, N., Meier, P., Restrepo, J., & Spagat, M. (2006, August 31–September 3). Natural Disasters, Casualties and Power Laws: A Comparative Analysis with Armed Conflict. In 2006 Annual Meeting of the American Political Science Association (APSA). 165. Geneva Academy. (2023). Today’s Armed Conflicts. Available online: https://geneva-academy.ch/galleries/today-s-armed-conflicts. Accessed 22 Aug 2023. 166. Lawrence Livermore National Laboratory. (undated). Energy Flow Charts: Charting the Complex Relationship among Energy, Water and Carbon. Available online: https://flowcharts.llnl.gov/. Accessed 3 Aug 2023. 167. Parthiban, R., & Ponnambalam, P. (2022). An Enhancement of the Solar Panel Efficiency: A Comprehensive Review. Frontiers in Energy Research, 10, 937155. 168. Chaisson, E. J. (2011). Long-Term Global Heating from Energy Usage. Eos, Transactions, American Geophysical Union, 89(28), 253–254. 169. Steiglechner, P. (2021). Estimating Global Warming from Anthropogenic Heat Emissions [Master’s Thesis, Universität Potsdam]. 170. Meadows, D. H. (1972). The Limits to Growth. Universe Books. 171. Herrington, G. (2021). Update to Limits to Growth: Comparing the World3 Model with Empirical Data. Journal of Industrial Ecology, 25(3), 614–626. 172. Sundstrom, S. M., & Allen, C. R. (2019). The Adaptive Cycle: More Than a Metaphor. Ecological Complexity, 39, 100767. 173. Zu Castell, W., & Schrenk, H. (2020). Computing the Adaptive Cycle. Scientific Reports, 10(1), 18175. 174. Bar-On, Y. M., Phillips, R., & Milo, R. (2018). The Biomass Distribution on Earth. Proceedings of the National Academy of Sciences, 115(25), 6506– 6511.

94

N. KING AND A. JONES

175. Feulner, G. (2012). The Faint Young Sun Problem. Reviews of Geophysics, 50(2). https://doi.org/10.1029/2011RG000375. 176. Dyke, J. G., & Weaver, I. S. (2013). The Emergence of Environmental Homeostasis in Complex Ecosystems. PLoS Computational Biology, 9(5), e1003050. 177. Radboud University. (2017). Chapter 7—Stellar Models and Stellar Stability. Available online: https://www.astro.ru.nl/~onnop/education/ stev_utrecht_notes/chapter7-8.pdf. Accessed 23 Feb 2023. 178. Vazza, F., & Felutti, A. (2020). The Quantitative Comparison Between the Neuronal Network and the Cosmic Web. Frontiers in Physics, 8, 491. 179. Davis, J. D., Olivença, D. V., Brown, S. P., & Voit, E. O. (2022). Methods of Quantifying Interactions among Populations Using Lotka-Volterra Models. Frontiers in Systems Biology, 2, 1021897. 180. Coelho, P. (2022). The Betz Limit and the Corresponding Thermodynamic Limit. Wind Engineering, 47 (2), 491–496. 181. Jacobson, M. Z., von Krauland, A.-K., Coughlin, S. J., Dukas, E., Nelson, A. J. H., Palmer, F. C., & Rasmussen, K. R. (2022). Low-cost Solutions to Global Warming, Air Pollution, and Energy Insecurity for 145 Countries. Energy and Environmental Science, 8(15), 3343–3359. 182. Breyer, C., Khalili, S., Bogdanov, D., Ram, M., Oyewo, A. S., Aghahosseini, A., Gulagi, A., Soloman, A. A., Keiner, D., Lopez, G., Østergaard, P. A., Lund, H., Mathiesen, B. V., Jacobson, M. Z., Victoria, M., Tske, S., Pregger, T., Fthenakis, V., Raugei, M., … Sovacool, B. K. (2022). On the History and Future of 100% Renewable Energy Systems Research. IEEE Access, 10, 78176–78218. 183. Way, R., Ives, M. C., Mealy, P., & Farmer, J. D. (2022). Empirically Grounded Technology Forecasts and the Energy Transition. Joule, 6(9), 2057–2082. 184. Denholm, P., Arent, D. J., Baldwin, S. F., Bilello, D. E., Brinkman, G. L., Cochran, J. M., Cole, W. J., Frew, B., Gevorgian, V., Heeter, J., Hodge, B.-M.S., Kroposki, B., Mai, T., O’Malley, M. J., Palmintier, B., Steinberg, D., & Zhan, Y. (2021). The Challenges of Achieving a 100% Renewable Electricity System in the United States. Joule, 5(6), 1331–1352. 185. Solé, J., Samsó, R., García-Ladona, E., García-Olivares, A., BallabreraPoy, J., Madurell, T., Turiel, A. Osychenko, O., Álvarez, D., Bardi, U., ˇ Baumann, M., Buchmann, K., Capellán-Pérez, Í., Cerný, M., Carpintero, Ó., De Blas, I., De Castro, C., De Lathouwer, J-D., Duce, … Theofilidi, M. (2020). Modelling the Renewable Transition: Scenarios and Pathways for a Decarbonized Future Using Pymedeas, a New Open-Source Energy Systems Model. Renewable and Sustainable Energy Reviews, 132, 110105.

2

THE TRIFURCATION OF ENERGY FUTURES

95

186. de Blas, I., Miguel, L. J., & Capellán-Pérez, I. (2019). Modelling of Sectoral Energy Demand Through Energy Intensities in MEDEAS Integrated Assessment Model. Energy Strategy Reviews, 26, 100419. 187. Capellán-Pérez, I., de Blas, I., Nieto, J., de Castro, C., Miguel, L. J., Carpintero, Ó., Mediavilla, M., Lobejón, L. F., Ferreras-Alonso, N., Rodrigo, P., Frechoso, F., & Álvarez-Antelo, D. (2020). MEDEAS: A New Modeling Framework Integrating Global Biophysical and Socioeconomic Constraints. Energy & Environmental Science, 13, 986–1017. 188. Ruhnau, O., & Qvist, S. (2022). Storage Requirements in a 100% Renewable Electricity System: Extreme Events and Inter-Annual Variability. Environmental Research Letters, 17 , 044018. 189. Fragaki, A., Markvart, T., & Laskos, G. (2019). All UK Electricity Supplied by Wind and Photovoltaics—The 30–30 Rule. Energy, 169, 228–237. 190. Miller, K. A., Thompson, K. F., Johnston, P., & Santillo, D. (2018). An Overview of Seabed Mining Including the Current State of Development, Environmental Impacts, and Knowledge Gaps. Frontiers in Marine Science, 4, 418. 191. World Economic Forum. (2022). 3 Ways the Circular Economy Is Vital for the Energy Transition. Available online: https://www.weforum.org/age nda/2022/02/3-ways-circular-economy-renewables-energy-transition/. Accessed 4 Mar 2023. 192. Michaux, S. P. (2021). The Mining of Minerals and the Limits to Growth. Geological Survey of Finland, Report number: 16/2021. 193. International Energy Agency. (2022). World Energy Outlook Special Report—The Role of Critical World Energy Outlook Special Report Minerals in Clean Energy Transitions. Available online: https://www.iea.org/rep orts/the-role-of-critical-minerals-in-clean-energy-transitions/mineral-req uirements-for-clean-energy-transitions. Accessed 2 Mar 2023. 194. Moreau, V., Dos Reis, P. C., & Vuille, F. (2019). Enough Metals? Resource Constraints to Supply a Fully Renewable Energy System. Resources, 8(1), 29. 195. Our World in Data. (2022). Energy Mix. Available online: https://ourwor ldindata.org/energy-mix. Accessed 4 Aug 2023. 196. Green Journal. (2020). How Much Waste Is Produced From Renewables vs. Fossil Fuels? Available online: https://www.greenjournal.co.uk/2020/06/ how-much-waste-is-produced-from-renewables-vs-fossil-fuels/. Accessed 5 Mar 2023. 197. Capellán-Pérez, I., de Castro, C., & Arto, I. (2017). Assessing Vulnerabilities and Limits in the Transition to Renewable Energies: Land Requirements under 100% Solar Energy Scenarios. Renewable and Sustainable Energy Reviews, 77 , 760–778.

96

N. KING AND A. JONES

198. MacKay, D. J. C. (2013). Solar Energy in the Context of Energy Use, Energy Transportation and Energy Storage. Philosophical Transactions of the Royal Society A, 371, 20110431. 199. Niebuhr, B. B., Sant’Ana, D., Panzacchi, M., van Moorter, B., Sandström, P., Morato, R. G., & Skarin, A. (2022). Renewable Energy Infrastructure Impacts Biodiversity Beyond the Area It Occupies. Proceedings of the National Academy of Sciences, 119(48), e2208815119. 200. Murphy-Mariscal, M. M., Grodsky, S. M., & Hernandez, R. R. (2018). 20—Solar Energy Development and the Biosphere. In A Comprehensive Guide to Solar Energy Systems with Special Focus on Photovoltaic Systems (391–405). Academic Press. 201. Earth.org. (2021). Techno-Optimism: Why Money and Technology Won’t Save Us. Available online: https://earth.org/techno-optimism/. Accessed 4 Aug 2023. 202. Space.com. (2022). Can Space-Based Solar Power Really Work? Here Are the Pros and Cons. Available online: https://www.space.com/space-solarpower-pros-cons. Accessed 6 Mar 2023. 203. Cash, I. (2017). CASSIOPeiA Solar Power Satellite. In 2017 IEEE International Conference on Wireless for Space and Extreme Environments (WiSEE) (pp. 144–149). 204. Ryan, R. G., Marais, E. A., Balhatchet, C. J., & Eastham, S. D. (2022). Impact of Rocket Launch and Space Debris Air Pollutant Emissions on Stratospheric Ozone and Global Climate. Earth’s Future, 10, e2021EF002612. 205. Arciuolo, T. F., & Faezipour, M. (2022). Yellowstone Caldera Volcanic Power Generation Facility: A New Engineering Approach for Harvesting Emission-Free Green Volcanic Energy on a National Scale. Renewable Energy, 198, 415–425. 206. Millward-Hopkins, J., & Oswald, Y. (2023). Reducing Global Inequality to Secure Human Wellbeing and Climate Safety: A Modelling Study. The Lancet Planetary Health, 7 (2), E147–E154. 207. Weiss, M., & Cattaneo, C. (2017). Degrowth—Taking Stock and Reviewing an Emerging Academic Paradigm. Ecological Economics, 137 , 220–230. 208. Seibert, M. K., & Rees, W. E. (2021). Through the Eye of a Needle: An Eco-Heterodox Perspective on the Renewable Energy Transition. Energies, 14(15), 4508. 209. Raworth, K. (2017). Doughnut Economics: Seven Ways to Think Like a 21stCentury Economist. Random House Business. 210. Roy, A., Basu, A., & Dong, X. (2021). Achieving Socioeconomic Development Fuelled by Globalization: An Analysis of 146 Countries. Sustainability, 13(9), 4913.

2

THE TRIFURCATION OF ENERGY FUTURES

97

211. Jackson, R. B., Ahlström, A., Hugelius, G., Wang, C., Porporato, A., Ramaswami, A., Roy, J., & Yin, J. (2022). Human Well-Being and Per Capita Energy Use. Ecosphere, 13(4), e3978. 212. White, E., & Kramer, G. J. (2019). The Changing Meaning of Energy Return on Investment and the Implications for the Prospects of Post-fossil Civilization. One Earth, 1(4), 416–422. 213. Trainer, T. (2021). Degrowth: How Much Is Needed? Biophysical Economics and Sustainability, 6(2), 5. 214. Büchs, M., & Koch, M. (2019). Challenges for the Degrowth Transition: The Debate About Wellbeing. Futures, 105, 155–165. 215. Hickel, J., Kallis, G., Jackson, T., O’Neill, D., Schor, J. B., Steinberger, J. K., Victor, P. A., & Ürge-Vorsatz, D. (2022). Degrowth Can Work— Here’s How Science Can Help. Nature, 612, 400–403. 216. Kauffman, S. (2022). Is There a Fourth Law for Non-Ergodic Systems That Do Work to Construct Their Expanding Phase Space? Entropy, 24(10), 1383. 217. Liu, G., Yang, Z., Giannetti, B. F., Casazza, M., Agostinho, F., Pan, J., Yan, N., Hao, Y., Zhang, L., Almeida, C. M. V. B., Gonella, F., Ilgiati, S., & Brown, M. T. (2021). Energy Constrains to Increasing Complexity in the Biosphere. The Innovation, 2(4), 100169. 218. Vestas. (2023). Vestas Unveils Circularity Solution to End Landfill for Turbine Blades. Available online: https://www.vestas.com/en/media/ company-news/2023/vestas-unveils-circularity-solution-to-end-landfill-forc3710818. Accessed 5 Mar 2023. 219. MIT Technology Review. (2021). Solar Panels Are a Pain to Recycle. These Companies Are Trying to Fix That. Available online: https://www.techno logyreview.com/2021/08/19/1032215/solar-panels-recycling/. Accessed 5 Mar 2023. 220. Woods, J., Williams, A., Hughes, J. K., Black, M., & Murphy, R. (2010). Energy and the Food System. Philosophical Transactions of the Royal Society of London B, 365(1554), 2991–3006. 221. International Energy Agency. (2022). How the Energy Crisis Is Exacerbating the Food Crisis. Available online: https://www.iea.org/commentar ies/how-the-energy-crisis-is-exacerbating-the-food-crisis. Accessed 4 Aug 2023. 222. Ivanovich, C. C., Sun, T., Gordon, D. R., & Ocko, I. B. (2023). Future Warming from Global Food Consumption. Nature Climate Change, 13(3), 297–302. 223. Gro Intelligence. (2017). How Can We Avoid a Food Crisis That’s Less Than a Decade Away. Available online: https://www.gro-intelligence.com/ins ights/214-trillion-calories. Accessed 14 Mar 2023.

98

N. KING AND A. JONES

224. Monbiot, G. (2022). Regenesis: Feeding the World Without Devouring the Planet. Penguin Books. 225. Järviö, N., Maljanen, N.-L., Kobayashi, Y., Ryynänen, T., & Tuomisto, H. L. (2021). An Attributional Life Cycle Assessment of Microbial Protein Production: A Case Study on Using Hydrogen-Oxidizing Bacteria. Science of the Total Environment, 776, 145764. 226. Williams, H. J., Hashad, K., Wang, H., & Zhang, K. M. (2023). The Potential for Agrivoltaics to Enhance Solar Farm Cooling. Applied Energy, 332, 120478. 227. International Atomic Energy Agency. (2023). Nuclear–Renewable Hybrid Energy Systems. IAEA Nuclear Energy Series, No. NR-T-1.24, Vienna, Austria. 228. Global Energy Monitor. (2023). Pedal to the Metal—It’s Time to Shift Steel Decarbonization into High Gear. Available online: https://globalene rgymonitor.org/wp-content/uploads/2023/07/GEM_SteelPlants2023. pdf. Accessed 22 Aug 2023. 229. Schoeneberger, C. A., McMillan, C. A., Kurup, P., Akar, S., Margolis, R., & Masanet, E. (2020). Solar for Industrial Process Heat: A Review of Technologies, Analysis Approaches, and Potential Applications in the United States. Energy, 206, 118083. 230. World Nuclear Association. (2021). Advanced Nuclear Power Reactors. Available online: https://world-nuclear.org/information-library/nuclearfuel-cycle/nuclear-power-reactors/advanced-nuclear-power-reactors.aspx. Accessed 22 Aug 2023. 231. Juárez-Martínez, L. C., Espinosa-Paredes, G., Vázquez-Rodríguez, A., & Romero-Paredes, H. (2021). Energy Optimization of a Sulfur-Iodine Thermochemical Nuclear Hydrogen Production Cycle. Nuclear Engineering and Technology, 53(6), 2066–2073. 232. Nitzbon, J., Heitzig, J., & Parlitz, U. (2017). Sustainability, Collapse and Oscillations in a Simple World-Earth Model. Environmental Research Letters, 12(7), 074020.

CHAPTER 3

Conclusions

Abstract Human civilisation has evolved over time to utilise a succession of different energy types and sources, which led eventually to phenomena such as the ‘Great Acceleration’ and the global ‘energy bind’, in which civilisation has been made vulnerable by its dependence on growing energy (primarily fossil fuel) flows. The collective ‘energy future’ will likely trifurcate into one of three Branches (Fossil-Seneca; Continued Growth; and Stabilisation) with very different characteristics; of these, the Stabilisation Branch likely offers the optimal way forward for human civilisation in terms of addressing the ‘energy bind’. The current fossil fuel system likely has an ‘inertia’ which may present systemic risks and challenges to undertaking a full transition to a renewables-based system, but if it can be achieved the prospect of long-term stability may offer a solution to the ‘Great Filter’; this will however require that near term problems are given due attention. Keywords Energy bind · Water-Energy-Food nexus · Systemic inertia · Longtermism · Great Filter · Exo-civilisations

The appropriation and use of exosomatic energy is a uniquely human phenomenon and is of particular importance in the context of the ‘WaterEnergy-Food Nexus ’, as it the critical resource on which humans have © The Author(s), under exclusive license to Springer Nature Switzerland AG 2023 N. King and A. Jones, Future Energy Options from a Systems Perspective, https://doi.org/10.1007/978-3-031-46448-5_3

99

100

N. KING AND A. JONES

become reliant to maintain a planet-spanning civilisation. The types of energy humans use, and the means of appropriation, have undergone evolution and fundament transformations over time. The taming of fire (biomass combustion) was the first stage of the control and application of energy outside of human metabolism and this was a crucial enabling step towards agriculture, which increased the total amount of food energy available and allowed control of animal metabolism. As human numbers grew and civilisations developed, natural energy flows (primarily wind and water) were utilised at increasingly large scales, alongside human slavery and continued use of various types of biomass (including whale oil). The most fundamental energy revolution started with the rise in the use of fossil biomass. This involved the combustion of first coal, and later oil and natural gas at hitherto unheard-of scales and for new industrial applications, and later electricity generation. New technological means of exploiting natural energy flows and energy from nuclear fuels have been the most recent additions to the human energy mix. A key characteristic of this evolution over time has been the addition of new energy sources and technologies to the pre-existing ones, which has allowed a continuous rise in the total amount of energy used. This phenomenon has been enabled by the high net energy provided by fossil fuels and the self-organisation of the Superorganism. The most recent stage has been the ‘Great Acceleration’, during which human civilisation has rapidly grown to its current state, but which has also led to an ‘energy bind’. This arises from complex human civilisation having become dependent on fossil fuels, but with the addition of energy sources preventing a full transition occurring, global society is left vulnerable as a result of this dependence. The analysis presented in this book presents and describes the systemic characteristics of three separate ‘Branches’ for the collective human energy future; civilisation will likely trifurcate into one of these outcomes due to thermodynamics and limits to system behaviours, summarised as follows: ● First (Fossil-Seneca) Branch—this describes ‘Business-As-Usual’, with continued energy use growth based predominantly on fossil fuel energy. This branch may lead to a ‘Seneca collapse’, in which the combined and synergistic effects of fossil fuel depletion, increased costs of energy and severe climate change overwhelm human adaptive capacities and generate global reinforcing synchronous failures.

3

CONCLUSIONS

101

● Second (Continued Growth) Branch—this describes a transition away from a fossil fuelled energy system to a nuclear fusion or renewables technologies-based system which is able to exploit very abundant energy reserves and flows that could allow open ended energy use growth. Fusion-based technologies are technologically immature, would be expensive to build, and would be subject to fuel constraints, even though global reserves are in theory very abundant. Renewables are likely to face a range of systemic and materials-based constraints to open ended growth. This open-ended growth is also likely to generate a range of egregious societal effects. ● Third (Stabilisation) Branch—this describes a transition away from a fossil fuelled energy system to a renewables-based system which emulates the long term stability of natural system which equilibrise around external inputs. There would be significant challenges associated with a global transition to a fully renewables-based system, primarily around systemic impacts on the global economic system and material requirements, but if combined with degrowth to reduce energy demand it may be feasible. If realised, it could provide long term stability in parallel to providing other significant benefits. Overall, the Renewables Stabilisation Branch likely offers the optimal way forward for human civilisation in terms of addressing the ‘energy bind’, with the other Branches described likely leading to collapse or having low feasibility or desirability. Indeed, while the Fossil-Seneca branch has its own negative feedback loop which invariably leads to collapse, the level of societal intervention and re-engineering of social norms required to make the Continued Growth path safe for societies (that is to avoid human civilisation from using unending energy access to destroy itself) is likely to be far greater than the effort required to unlock the potential of the Stabilisation branch. There would be significant challenges in moving away from the fossil fuel-based system due to civilisation being currently so heavily reliant on it, but if the transition could be made and stability achieved, there would be benefits that could be realised and gained over very long timescales. Specifically, it could permit the long term persistence of civilisation, and potentially could mitigate some of the critical risks that apply to humanity’s survival in the context of the ‘Great Filter’.

102

N. KING AND A. JONES

Indeed, the Stabilisation Branch could lead to human prosperity as we refocus societal goals away from growth as a key aim and move towards aspects of wellbeing and human flourishing as the desirable outcome.

Fossil Fuel Systemic Inertia The analysis undertaken in this study provides a strong underpinning for a transition of the global energy system to a renewables based steady (or near-steady) state, which would likely need to occur concurrently with a downshift in global energy intensity. However, for the transition to the renewables-stabilised future to occur, a transition away from the fossil fuelled present must also occur, which itself may involve significant challenges not fully captured in the preceding discussions. Fossil fuels currently provide the vast majority of the primary energy used globally. This situation has persisted and remained extant through huge changes in global energy usage (in terms of total quantity, and applications/new sources of energy being pioneered being pioneered) over prolonged periods. This implies that the very high penetration of fossil fuels into the global energy system creates a systemic ‘inertia’ or ‘lock-in’ which will likely provide myriad forms of resistance to large scale, systemic shifts (such as to renewables dominating). A key manifestation of the dominance of fossil fuels in the energy system is the unrealistic, flawed and frequently self-defeating solutioneering proposed for the mitigation of climate change harms. Direct air capture (DAC) is a ‘negative emissions’ technology which operates by directly capturing and immobilising gaseous CO2 from air, allowing drawdown of the stock of this gas in the atmosphere [1]. Notwithstanding open questions about the energy requirements and scalability of DAC technologies to drawdown CO2 at rates commensurate with annual global emissions [2, 3], it is likely that avoiding GHG emissions in the first instance may produce better climate outcomes than allowing emissions and then later drawing them down using DAC [4]. Geoengineering (specifically Solar Radiation Management, SRM) is another proposed approach to mitigate climate change by reducing the solar energy flux reaching the Earth’s surface. Some of the proposals for achieving this have very low feasibility [5] and/or introduce significant risks [6]. Carbon Capture and Storage has also failed to reach commercial deployment in recent decades, despite international climate change agreements making this a central tenet of achieving emissions reductions [7].

3

CONCLUSIONS

103

In all of these cases, significant energies and efforts are proposed to be expended to enable the continuation and perpetuation of fossil fuelbased energy systems, when transitioning away from them may involve lesser complexity, lower risks and more optimised outcomes. However, even this position (by the fossil fuel industry and its supporters) of proposing diversionary approaches represents an advancement over historically held positions of outright climate change denial. The suppression of the outputs of studies (as part of concerted efforts to avoid acknowledgement of the harms of fossil fuel products) indicates the determination of the industry to maintain a dominant position in the energy system [8]. Recent global events have indicated that the fossil fuel energy system also has considerable scope to act adaptively to seek to maintain its position. Following the invasion of Ukraine by the Russian Federation in 2022, the natural gas supply to Europe underwent a rapid and large scale reduction. However, natural gas and other fossil use persisted due to co-ordinated transnational efforts to obtain and distribute alternative supplies [9, 10]. This indicates that adaptive responses, achieved through the effective application of the distributed intelligence of adaptive systems (responding to phase shifts in conditions and other emergent phenomena which may be a manifestation of the Superorganism) could provide wide scope for fossil fuel systems to find innovative solutions to maximise persistence. Conversely, the same behaviour may also provide the basis for the global energy system to re-organise to seek out long-term stable equilibrium conditions as described in the Stabilisation Branch. Innovative approaches to national and international policy and leadership, such as integration of systems science and risk-based approaches, will be required to influence whether energy system evolution will tend towards maintenance of incumbent fossil fuel systems, or towards the renewables transition [11]. A key underpinning aspect of the renewables transition will likely be a requirement for concurrent degrowth of the energy and material intensity of human civilisation. In the competition between continued fossil-fuelled high energy use, and renewables-supported lower energy use, low cost and complexity approaches may start to increasingly prevail by default over expensive fossil fuelled technologies (as solutions for managing fundamental societal challenges). Examples of this include: the potential widespread use of external paints and films for radiative passive cooling of dwellings and urban areas (which will also reduce overall global radiative forcing, if sufficiently widely used [12]) in place of energy-intensive

104

N. KING AND A. JONES

air conditioning (the increasing use of which could create an enhancing climate change feedback loop [13]); and promotion and application of effective, locally implemented water saving schemes [14] in favour of pursuing highly complex, expensive and difficult-to-deliver water transfer ‘megaprojects ’ to supply arid regions [15]. If shifts towards simpler, lowerenergy solutions such as these were to start occurring globally in multiple settings to increasing degrees with time, enhancing feedbacks may start to take effect which may start to drive more general systemic changes (in line with the ‘breakthrough effects ’). These shifts may also be increasingly driven by growing public awareness, dissatisfaction and protest against with the extant system at global scale [16, 17]. Particular risks which could likely apply to a rapid drawdown in fossil fuel use relate to the dependency of current agricultural systems on energetic inputs from fossil fuels, and to the formation of ‘stranded assets ’ in the fossil fuel industry. The contemporary global food system has developed (particularly since the advent of the ‘green revolution’) to require fossil fuel inputs at multiple stages in order to maintain outputs. This path dependency means that the growth of the human population to current levels has largely been enabled and created by food-borne fossil fuel energy [18, 19]. Therefore, in the absence of fossil or substitute energy input, current levels of human population could likely not be supported. The replacement of depleting fossil energy input with renewable (potentially supplemented by nuclear) energy is as such an imperative and could be achieved at least in part via the food technologies described Chapter 2. Fossil fuel extraction, processing and distribution is a highly complex and material/energy intensive, globally distributed industry which holds large inventories of valuable assets (e.g., offshore drilling rigs, refineries), with the ‘equity risk ownership’ spread through a large number of companies and private investors (including pension funds). A collapse of future fossil fuel profit expectations in the event of large demand reductions (i.e., as part of a full scale energy transition) could cause a net transfer of stranded asset risk to investors, which has the potential to drive significant financial and economic instabilities/risks globally. Control of the risks associated with phase-out of fossil fuels by the international community would therefore be required to stabilise potential knock-on effects such as these and ensure they don’t become ‘blockers’ to a transition [20].

3

CONCLUSIONS

105

Overall, there are likely multiple mechanisms and pathways through which the fossil fuel system may seek to maintain its position in the global energy system. There are also attendant risks in its termination, but systemic effects may assist in enabling this transition which is imperative to avoiding collapse scenarios and achieving long term stability.

Longer Term Perspectives The key differentiator between the different branches of the energy trifurcation is the potential to enable long-term stability for human civilisation, or alternatively to lead to near or medium term pervasive challenges and failures. In the context of ‘open ended’, ‘long term’ is assumed to refer to persistence of organised human societies in at least some areas of the planet on the timescales of decades to centuries. This would require that the energy (and other related key systems, notably those of the waterenergy-food nexus) continue to operate without the development and/ or strengthening of destabilising feedbacks which apply to the functioning of global civilisation. However, successfully achieving a renewables based stabilisation of human civilisation might have implications over even greater timescales. The concept of ‘longtermism’ has recently emerged as a philosophy or moral framework which considers the human future over very long timeframes, with an emphasis on the interests of humans in distant future timeframes (specifically, the much larger number of humans that may exist in remote future timeframes relative to the present). Although it has been criticised on the basis of the lesser emphasis it gives to current and near term timeframes (and the issues which are most pressing and relevant e.g., the different facets of the ‘polycrisis’ [21]), under this framework ensuring that global-scale catastrophes are avoided, and that civilisation persists, is paramount. Decisions and actions undertaken in contemporary timeframes will have a large influence on the future, so from this perspective ensuring that collective humanity pursues the correct Branch of the energy trifurcation is therefore also of paramount importance [22]. Another perspective could be that achieving long-term stability for civilisation could contribute towards avoidance of the ‘Great Filter’. This is a cosmological concept which arises from the apparent lack of evidence of extraterrestrial life or intelligence (such as observed extrasolar planets, or radio signals) despite the vastness of the observable universe, which indicates that there could be a ‘universal’ tendency towards civilisations

106

N. KING AND A. JONES

encountering existential threats and disasters the longer they endure and/ or become technologically advanced. Human and hypothetic ‘exo-civilisations’ have in common that they undertake organised resource-harvesting and operate within closed planetary environments, which will generate feedbacks that drive the planetary system away from its initial state. These feedbacks and the responses to them may lead to a range of different outcomes ranging from collapse (which in the human context may arise from phenomena such as climate change, pandemics, asteroid impacts, nuclear warfare, and environmental toxification) to stabilisation and persistence of the civilisation (achieved through foresight of the impacts of the perturbation and change in resource-harvesting strategy). The Stabilisation Branch of the energy trifurcation presented in this study represents the latter scenario, in that recognising the hazards of rapid destabilisation and/or long-term growth and diverting from these pathways could mitigate the existential risks as described by the ‘Great Filter’, particularly climate change [23–25]. Maintaining a stabilised human civilisation (and by extension, the Earth’s biosphere [26]) into the long and very long term will however require that near term problems are given due attention.

References 1. Erans, M., Sanz-Pérez, E. S., Hanak, D. P., Clulow, Z., Reiner, D. M., & Mutch, G. A. (2022). Direct Air Capture: Process Technology, Technoeconomic and Socio-political Challenges. Energy & Environmental Science, 15, 1360–1405. 2. Realmonte, G., Drouet, L., Gambhir, A., Glynn, J., Hawkes, A., Köberle, A. C., & Tavoni, M. (2019). An Inter-model Assessment of the Role of Direct Air Capture in Deep Mitigation Pathways. Nature Communications, 10(1), 3277. 3. Recharge News. (2021). The Amount of Energy Required by Direct Air Carbon Capture Proves It Is an Exercise in Futility. Available online: https://www.rechargenews.com/energy-transition/the-amount-of-energyrequired-by-direct-air-carbon-capture-proves-it-is-an-exercise-in-futility/2-11067588. Accessed 13 Mar 2023. 4. Zickfield, K., Azevedo, D., Mathesius, S., & Matthews, H. D. (2021). Asymmetry in the Climate-Carbon Cycle Response to Positive and Negative CO2 Emissions. Nature Climate Change, 11, 613–617. 5. Bromley, B. C., Khan, S. H., & Kenyon, S. J. (2023). Dust as a Solar Shield. PLOS Climate, 2(2), e0000133.

3

CONCLUSIONS

107

6. Parker, A., & Irvine, P. J. (2018). The Risk of Termination Shock From Solar Geoengineering. Earth’s Future, 6(3), 456–467. 7. Martin-Roberts, E., Scott, V., Flude, S., Johnson, G., Haszeldine, R. S., & Gilfillan, S. (2021). Carbon Capture and Storage at the End of a Lost Decade. One Earth, 4(11), 1569–1584. 8. Supran, G., Rahmstorf, S., & Oreskes, N. (2023). Assessing ExxonMobil’s Global Warming Projections. Science, 379(6628), eabk0063. 9. Oxford Institute for Energy Studies. (2023). Ukraine Invasion: What This Means for the European Gas Market. Available online: https://www.oxf ordenergy.org/publications/ukraine-invasion-what-this-means-for-the-eur opean-gas-markets/. Accessed 6 Feb 2023. 10. The Guardian. (2023). How Putin’s Plans to Blackmail Europe over Gas Supply Failed. Available online: https://www.theguardian.com/world/ 2023/feb/03/putin-russia-blackmail-europe-gas-supply-ukraine. Accessed 6 Feb 2023. 11. Pearson, R., & Bardsley, D. K. (2022). Applying Complex Adaptive Systems and Risk Society Theory to Understand Energy Transitions. Environmental Innovation and Societal Transitions, 42, 74–87. 12. Li, X., Peoples, J., Yao, P., & Ruan, X. (2021). Ultrawhite BaSO4 Paints and Films for Remarkable Daytime Subambient Radiative Cooling. ACS Applied Materials & Interfaces, 13(18), 21733–21739. 13. Cary Institute of Ecosystem Studies. (2020). Air Conditioning. Available online: https://www.caryinstitute.org/news-insights/blog-translational-eco logy/air-conditioning. Accessed 12 Mar 2023. 14. Tsai, Y., Cohen, S., & Vogel, R. M. (2011). The Impacts of Water Conservation Strategies on Water Use: Four Case Studies. Journal of American Water Resources Association, 47 (4), 687–701. 15. Shumilova, O., Tockner, K., Thieme, M., Koska, A., & Zarfl, C. (2018). Global Water Transfer Megaprojects: A Potential Solution for the WaterFood-Energy Nexus? Frontiers in Environmental Science, 6, 150. 16. Bugden, D. (2020). Does Climate Protest Work? Partisanship, Protest, and Sentiment Pools. Socius, 6, 2378023120925949. 17. Gregersen, T., Andersen, G., & Tvinnereim, E. (2023). The Strength and Content of Climate Anger. Global Environmental Change, 82, 102738. 18. Neff, R. A., Parker, C. L., Kirschenmann, F. L., Tinch, J., & Lawrence, R. S. (2011). Peak Oil, Food Systems, and Public Health. American Journal of Public Health, 101(9), 1587–1597. 19. Yes Magazine. (2016). Without Fossil Fuels, a New Population Puzzle—So How Many People Can the Planet Really Support? Available online: https:// www.yesmagazine.org/issue/life-after-oil/2016/03/22/without-fossilfuels-a-new-population-puzzle. Accessed 14 Mar 2023.

108

N. KING AND A. JONES

20. Semieniuk, G., Holden, P. B., Mercure, J.-F., Salas, P., Pollitt, H., Jobson, K., Vercoulen, P., Chewpreecha, U., Edwards, N. R., & Viñuales, J. E. (2022). Stranded Fossil-Fuel Assets Translate to Major Losses for Investors in Advanced Economies. Nature Climate Change, 12, 532–538. 21. Aeon. (2021). Against Longtermism. Available online: https://aeon.co/ess ays/why-longtermism-is-the-worlds-most-dangerous-secular-credo. Accessed 27 Jan 2023. 22. MacAskill, W. (2022). What We Owe the Future. Basic Books. 23. Frank, A., Carroll-Nellenback, J., Alberti, M., & Kleidon, A. (2018). The Anthropocene Generalized: Evolution of Exo-civilizations and Their Planetary Feedback. Astrobiology, 18(5), 503–518. 24. Jiang, J. H., Rosen, P. E., Lu, K., Fahy, K. A., & Obacz, P. (2023). Avoiding the “Great Filter”: Extraterrestrial Life and Humanity’s Future in the Universe. Journal of Humanities & Social Sciences, 6(2), 59–66. 25. Jiang, J. H., Rosen, P. E., & Fahy, K. A. (2021). Avoiding the “Great Filter”: A Projected Timeframe for Human Expansion Off-World. Galaxies, 9(3), 53. 26. Jebari, K., & Sandberg, A. (2022). Ecocentrism and Biosphere Life Extension. Science and Engineering Ethics, 28(6), 46.

Index

A adaptive capacity, 48 adaptive cycles , 64 adaptive responses, 103 Agricultural Revolution, 5 agrivoltaics , 78 animal metabolism, 6 Anthropocene, 16 autopoietic behaviour, 23

B biomass, 4, 8 “breakthrough effects ”, 57 Business-As-Usual , 34

C capital expenditure (capex), 41 Carbon Capture and Storage, 102 carbon dioxide, 35 circular economy, 72 climate change, 36 coal, 9 collapse, 46

commercially viable, 53 complexification, 2 complex societies, 38 concentrated solar power, 13 constraints, 51 consumerism, 11 conventional oil, 39

D deep-sea mining, 71 degrowth, 75 depletion, 22 desalination, 78 deuterium, 52 Direct Air Capture, 59 discretionary energy, 21 DT-based fusion, 52

E Earth System, 34 economic growth, 11 economic substitutes, 18 economies of scale, 12

© The Editor(s) (if applicable) and The Author(s), under exclusive license to Springer Nature Switzerland AG 2023 N. King and A. Jones, Future Energy Options from a Systems Perspective, https://doi.org/10.1007/978-3-031-46448-5

109

110

INDEX

ecosystems, 37 electrical grids, 20 electrification, 10 elite, 59 emergent, 18 Endosomatic energy, 2 energy appropriation, 15 ‘energy bind’, 24 energy flows, 6 ‘energy futures’, 34 energy paradigm, 6, 14, 34 Energy Return on Investment, 20 energy surpluses, 22 energy transition, 5 energy transition scenarios, 69 enhancing feedbacks, 9 entropic effects, 22 entropy, 43 exosomatic energy, 2 externalities, 57 extraction, 16 F fossil fuels, 8 ‘Fully Automated Luxury Communism’, 59 G Gaia, 65 Generation IV, 43 Geoengineering, 102 geological timeframes, 35 Geothermal energy, 13 geothermal heat, 12 global carbon cycle, 35 globalisation, 23 global primary energy, 41 global primary energy use, 11 Great Acceleration, 16 Great Britain, 8 ‘Great Filter’, 105

greenhouse gas (GHG) emissions, 12

H Haber-Bosch process, 10 Holocene, 37 ‘human climate niche’, 38 human slavey, 7 Hybrid Energy Systems, 78 hydroelectric power, 11

I ignition, 51 industrial heat, 10 Industrial Revolution, 8 inequality, 58 Inertial Confinement Fusion (ICF), 53 insolation, 50 Integrated Assessment Models, 68 interdependence, 24 Intergovernmental Panel on Climate Change, 37 intermittency, 70 internal combustion engines, 9 International Thermonuclear Experimental Reactor, 53 internet, 61

J Jevons Paradox, 24

L Light Water Reactor, 14 limiting value, 65 ‘limits to growth’, 69 ‘lock-in’, 102 longtermism, 105 ‘low hanging fruit’ principle, 21

INDEX

M Magnetic Confinement Fusion (MCF), 53 Magnetised Target Fusion, 53 mass extinction events, 37 MEDEAS, 68 mineral ores, 42 misinformation, 61 modularity, 44

R radiative forcing, 36 radioactive waste management, 42 recycling, 71 Re-Materialisation, 70 renewable energy, 11 research and development, 44 reservoir, 39 runaway climatic effects, 37

N National Ignition Facility, 54 Natural gas, 10 ‘negative emissions’ technology, 102 negative sum, 59 nodes, 23 nuclear power, 14

S scarcities, 69 Seneca Effect, 45 Small- and Advanced-Modular Reactors, 43 societal complexity, 17 Solar energy, 13 ‘space-based solar power’, 73 sperm oil, 7 stabilisation, 64 stagnation, 76 steam engines, 9 storage systems, 70 ‘stranded assets ’, 104 succession, 24 superorganism, 23 synchronous peaking, 40 systemic change, 11 Systemic Inertia, 102

O oil, 9 open-ended, 49

P path dependency, 42 pathways, 34 Peak oil, 39 ‘phase space’, 76 photovoltaic, 13 polycrisis, 2 portability, 17 positive sum, 59 power-law distributions, 62 precision fermentation, 77 Pressurised Water Reactor (PWR), 42 public acceptability, 41 pyrotechnology, 4

Q Q value, 51

T technological innovations, 9 technological lock in, 42 Technology Readiness Level, 44 ‘techno-optimist’, 73 ‘Technosphere’, 60 thermodynamic system, 62 thorium, 44 tipping points, 36 trifurcating, 34 tritium, 52

111

112

INDEX

U unconventional oil, 39 uranium, 42

V variability, 70

W Waste heat, 63 Water–Energy–Food Nexus , 2 waterwheels, 6 whale oil, 7 Windmills, 6 wind power, 6