You are on page 1of 16

Applied Energy xxx (xxxx) xxx–xxx

Contents lists available at ScienceDirect

Applied Energy
journal homepage: www.elsevier.com/locate/apenergy

Industrial energy use and carbon emissions reduction in the chemicals


sector: A UK perspective

Paul W. Griffina, , Geoffrey P. Hammonda,b, Jonathan B. Normana
a
Department of Mechanical Engineering, University of Bath, Bath BA2 7AY, United Kingdom
b
Institute for Sustainable Energy and the Environment (I•SEE), University of Bath, Bath BA2 7AY, United Kingdom

H I G H L I G H T S

• Future decarbonisation of the UK Chemicals sector has been evaluated.


• The improvement potential of different technological interventions was assessed.
• 2050 ‘technology roadmaps’ were also developed for various alternative scenarios.
• Best practice technologies will prompt short-term energy and CO emissions savings.
2

• The prospects for longer-term, ‘disruptive technologies’ are far more speculative.

A R T I C L E I N F O A B S T R A C T

Keywords: The opportunities and challenges to reducing industrial energy demand and carbon dioxide (CO2) emissions in
Chemicals the Chemicals sector are evaluated with a focus on the situation in the United Kingdom (UK), although the lessons
Industrial energy analysis learned are applicable across much of the industrialised world. This sector can be characterised as being het-
Carbon accounting erogeneous; embracing a diverse range of products (including advanced materials, cleaning fluids, composites,
Enabling technologies
dyes, paints, pharmaceuticals, plastics, and surfactants). It sits on the boundary between energy-intensive (EI)
Improvement potential
and non-energy-intensive (NEI) industrial sectors. The improvement potential of various technological inter-
United Kingdom
ventions has been identified in terms of their energy use and greenhouse gas (GHG) emissions. Currently-available
best practice technologies (BPTs) will lead to further, short-term energy and CO2 emissions savings in chemicals
processing, but the prospects for the commercial exploitation of innovative technologies by mid-21st century are
far more speculative. A set of industrial decarbonisation ‘technology roadmaps’ out to the mid-21st Century are
also reported, based on various alternative scenarios. These yield low-carbon transition pathways that represent
future projections which match short-term and long-term (2050) targets with specific technological solutions to
help meet the key energy saving and decarbonisation goals. The roadmaps’ contents were built up on the basis of
the improvement potentials associated with various processes employed in the chemicals industry. They help
identify the steps needed to be undertaken by developers, policy makers and other stakeholders in order to
ensure the decarbonisation of the UK chemicals industry. The attainment of significant falls in carbon emissions
over this period will depends critically on the adoption of a small number of key technologies [e.g., carbon
capture and storage (CCS), energy efficiency techniques, and bioenergy], alongside a decarbonisation of the
electricity supply.

1. Introduction ‘greenhouse gas’ (GHG) emissions associated with industry [1]. Chem-
istry provides the fundamental basis for the synthesis of core inter-
1.1. Background mediate and end products in order to satisfy human needs. It supplies
inputs to matter transformation chains in other industrial sectors, e.g.,
The chemical and petrochemical industry represents the largest plastics, composite materials, industrial gases, fertilizers, and so on [2].
contributor to industrial energy demand worldwide. It accounts for These products are key to the modern global economy stretching from
about 10% of global total final energy consumption and 7% of agriculture to medicine, through fuels, plastics and synthetic textiles.


Corresponding author at: CDP - Global Environmental Reporting System, 71 Queen Victoria Street, London EC4V 4AY, United Kingdom.
E-mail address: paul.griffin@cdp.net (P.W. Griffin).

http://dx.doi.org/10.1016/j.apenergy.2017.08.010
Received 23 January 2017; Received in revised form 21 July 2017; Accepted 6 August 2017
0306-2619/ © 2017 The Authors. Published by Elsevier Ltd. This is an open access article under the CC BY license (http://creativecommons.org/licenses/BY/4.0/).

Please cite this article as: Griffin, P.W., Applied Energy (2017), http://dx.doi.org/10.1016/j.apenergy.2017.08.010
P.W. Griffin et al. Applied Energy xxx (xxxx) xxx–xxx

But the world landscape of the chemical industry has dramatically al- significant role in energy and GHG emission reductions going forward
tered in recent years. Asian chemical companies displayed a major [7]. Cefic [8] were aided by Ecofys, the energy and sustainability con-
growth trend from 1980 onwards, driven by their domestic markets, sultancy, in the development of their 2050 European chemicals
and presently accounts for half of the global market [2]. A surge of new roadmap to a competitive, low-carbon future. They emphasised the
investment in chemical processing plant then took place in the Middle importance of making changes to the sectoral fuel mix, particularly for
East after the turn of the Millennium, based on the natural (oil and heat generation and Nitrous Oxide (N2O) production. This would yield
natural gas) resources in that area [2]. However, large-scale shale gas again about 15% reduction in GHG emissions by 2050, although they
exploration and development in the United States of America (USA), noted that deeper cuts could be achieved via the decarbonisation of the
after 2006 in particular, has given their chemical sector a strong eco- power system and the adoption of CCS facilities [8]. Both the latter
nomic advantage when compared with competitors elsewhere [2–4]. options would be costly and necessitate technological breakthroughs.
The price of gas halved with an impact on competitiveness of a large Indeed, a novel feature of the Cefic study [8] was the focus on the
part of the global chemical industry (both in terms of natural gas adverse impacts that energy and climate policy costs are likely to have
feedstock and fuel). Companies from across the world are locating on European competiveness vis-à-vis chemicals production in the USA
themselves in the US to take advantage of this ‘revolution’ [3]: leading and other regions. The ECF were likewise concerned about price com-
to a potential ‘Golden Age of Gas’, according to the International Energy petitiveness [9], albeit in the context of the so-called energy policy
Agency (IEA) [4]. The IEA sees shale gas as contributing about 14% to trilemma: competitiveness, sustainability and security of supply. They
global gas production by 2035 [3,4]. studied the transition dynamics in the chemicals industry drawing on
The IEA [1] partnered with the International Council of Chemical the Cefic roadmap [8]. ECF argued [9] that substantial GHG emissions
Associations (ICCA) and DECHEMA (Germany’s Society for Chemical reduction could be achieved through process and energy efficiency
Engineering and Biotechnology) in order to produce a technology improvements, alongside greater resource ‘circularity’ or value chain
roadmap for energy and GHG reductions in the world chemicals sector collaboration. Thus, they suggested that by seeking out “cross-process,
out to 2050. This roadmap was developed in the context of the IEA 2 °C cross-company, cross-sector, and cross country abatement opportu-
Degree Scenario (2DS) for global warming. It focused on the particular nities” European price competiveness in the chemicals sector could be
role of catalytic processes that account for roughly 90% of chemical maintained.
processing [1]. The roadmap built on earlier studies of Best Practice Aggregate studies of the chemicals industry on a global or region
Technologies (BPT) for improving energy efficiency in the sector [5] and scale have their limitations. Each country has, in reality, its own dis-
of carbon abatement innovations. The former used mainly a top-down tinctive historical background, structural characteristics (including ac-
approach to examine some 57 processes and indicated energy saving cess to resources), and potential for energy savings and decarbonisa-
potentials of 5–15% [5]. BPT are the most advanced, economically vi- tion. Therefore the present work seeks to draw out lessons from the
able technologies on an industrial scale. ICCA commissioned the chemicals sector and its development in Britain. Industry as a whole in
German Öko-Institut, with the support of McKinsey & Company, to criti- the United Kingdom (UK) accounts for some 21% of total delivered
cally review Carbon Life Cycle Assessment (cLCA) studies related to the energy and 29% of carbon emissions. There are large differences be-
chemicals industry [6]. This indicated the potential for GHG emissions tween industrial sectors in the end-use applications of energy, espe-
{carbon dioxide equivalent (CO2e)} abatement ‘from cradle to grave’, i.e., cially in terms of products manufactured, processes undertaken and
over a value chain incorporating the extraction of feedstock material technologies employed (see Fig. 1 [10]; where the final demand for
and fuels through to production, transport and distribution, product energy by broad UK industrial sectors is depicted against various energy
usage, and the ‘end of life’ (disposal or recycling) phases. More than one use categories). It is clear that the chemicals sector as seen in Fig. 1
hundred cLCA studies were submitted by ICCA member companies gives rise to the highest industrial energy consumption; mainly due to
from around the world to McKinsey for evaluation, and then the results low temperature heat processes (30%), electrical motors (19%), drying/
or data were reviewed by the Öko Institut. They concluded [6] that the separation processes (16%), and high temperature heat processes (11%)
best option for reducing global GHG emissions could be achieved by [10]. UK industry overall has been found to consist of some 350 sepa-
ensuring that each life-cycle stage of the value chain yielded its op- rate combinations of sub-sectors, devices and technologies [11,12].
timum contribution. Otherwise, a given stage might prevent larger CO2e Nevertheless, it is the only end-use energy demand sector in the UK that
reductions elsewhere along the chain, and consequently not elicit net has experienced a significant fall of roughly 40% in final energy con-
global reductions overall. sumption since the first oil price shock of 1973/74 [11,12]. This was in
Worldwide assessments of the chemicals industry have been sup- spite of a rise of over 40% in industrial output in value added terms.
plemented by regional ones. Europe, or the European Union (EU) However, the consequent aggregate reduction in energy intensity (MJ/
countries, in particular has instigated a number of studies of an energy- £ of gross value added) masks several different underlying causes: end-
efficient and low-carbon chemicals sector including, for example, those use efficiency {accounting for around 80% of the fall in industrial energy
sponsored by the European Commission (EC), via their Joint Research intensity; largely induced by the price mechanism); structural changes in
Centre (JRC) [7], the European Chemical Industry Council (Cefic) [8], and industry [a move away from energy-intensive (EI) industries towards non-
the European Climate Foundation (ECF) [9]. The JRC report [7] used a energy-intensive (NEI) ones, including services [11,12]}; and fuel
bottom-up model to evaluate the European chemical and petrochemical switching (from coal and oil to natural gas and electricity that are
sector. It assessed the current technological status of 26 basic chemical cleaner, more readily controllable, and arguably cheaper for the busi-
products (including fertilizers, organic and inorganic substances, nesses concerned).
polymers, etc.), as well as the associated sectoral energy use and GHG
emissions out to 2050. Over this period, it was found that the EU 1.2. The issues considered
chemicals industry would experience a 39% rise in energy consump-
tion, but a 15% fall in GHG emissions compared with the early 2010s The present study builds on work by Dyer et al. [11] commissioned
[7]. Around 50 Best Available Technologies (BAT), the most effective by the UK Government Office of Science (GOS) and on a recent ‘Advanced
innovations presently known, were examined. The importance of re- Review’ by Griffin et al. [13]. In each case, a variety of assessment
placing fossil fuel feedstocks by sustainable alternatives such as hy- techniques for determining potential energy use and GHG reductions
drogen (from electrolysis driven by renewables) and biomass was re- were discussed. Griffin et al. [13] then evaluated the wider UK in-
cognised. Two cross-cutting technologies [combined heat and power dustrial landscape with the aid of decomposition analysis [14] in order
(CHP), already widely used in the chemicals sector, and carbon capture to identify the factors that have led to energy and carbon savings over
and storage (CCS)] were recommended as having a potentially recent decades. They then assessed the improvement potential in two

2
P.W. Griffin et al. Applied Energy xxx (xxxx) xxx–xxx

Fig. 1. Final UK energy demand by industrial sub-sector and


end-use.
Source: Norman [10].

Fig. 2. Primary energy intensity, percentage of costs represented by energy


and water, and mean primary energy use per enterprise (reflected by the
area of the data points).
Source: adapted from Griffin et al. [13].

surfactants), and as sitting on the boundary between EI and NEI in-


dustries (see Fig. 2 [13]; where broad UK industrial sectors are split into
an EI and NEI categories). [A high value in any of the measures shown
in Fig. 2 suggests that the sub-sector is EI.] Nevertheless, it accounts for
some 19% of GHG emissions from UK industry – the second largest
sector after steel (see Fig. 3 [13]; where these emissions are split into
broad sectors in pie chart format {both those emanating from energy
use, including those indirectly emitted from electricity use, and process
emissions}). The opportunities and challenges to reducing industrial
energy demand and carbon dioxide (CO2) emissions (the principal GHG
[12]) in the UK Chemicals sector have therefore been evaluated, al-
though the lessons learned are applicable across much of the in-
dustrialised world. The data here has been largely extracted from an
industrial Usable Energy Database (UED) that was produced for the UK
Energy Research Centre (UKERC) by the present authors (see Griffin et al.
[13,15,16]). A set of industrial decarbonisation ‘technology roadmaps’
out to 2050 are finally reported, based on various alternative scenarios:
named Low Action (LA), Reasonable Action (RA), Reasonable Action in-
Fig. 3. Greenhouse gas (GHG) emissions from UK industry.
cluding CCS (RA-CCS), and Radical Transition (RT) respectively. Such
Source: adapted from Griffin et al. [13].
roadmaps represent future projections that match short-term (say out to
2035) and long-term (2050) targets with specific technological solu-
sectors: ‘Cement’ and ‘Food & Drink’, which represent the EI and NEI tions to help meet the key energy saving and decarbonisation goals.
industrial sectors respectively. Here the ‘Chemicals’ sector of UK in- Their contents were built up on the basis of the improvement potentials
dustry is examined in terms of energy use and GHG emissions from associated with various processes employed in the chemicals industry
product sub-sectors, as well as the improvement potential of its pro- and embedded in the UED [13,15,16]. They help identify the steps
cesses. This sector can be characterised as being heterogeneous (having needed to be made by industrialists, policy makers and other stake-
a diverse range of product outputs, including advanced materials, holders in order to ensure the decarbonisation of the UK chemicals
cleaning fluids, composites, dyes, paints, pharmaceuticals, plastics, and industry.

3
P.W. Griffin et al. Applied Energy xxx (xxxx) xxx–xxx

2. The chemicals sector caustic soda (sodium hydroxide - NaOH), and artificial fertilizers [17].
The organic chemical industry produced artificial dyestuffs from coal
2.1. Historical development of the chemicals industry tar, principally located in the Manchester and Glasgow areas [18], and
high explosives from compounds of nitric acid and cellulose material.
The way in which the chemicals industry has been perceived by By the late 19th Century, it also started to develop plastics and artificial
humanity has changed over time. In the ancient world, crafts were fibres [17].
originally developed to meet the requirements of a domestic setting The 1914–1918 Great War had the effect of making the UK re-
[17,18], such as the baking of bread, brewing, pottery making and cognise that it was vulnerable to disruptions in the importation of
leather tanning. Arabic science from about 3500 BCE {before the chemicals from Germany and its allies, and from a blockage of com-
‘Common Era’ (CE)}, based largely in Egypt and the Near East, led to mercial shipping by German submarines [20]. That meant that Britain
what is now recognised as chemicals [19–22]: the early smelting of had to reorganise its chemicals sector in order to produce substitute
metals [especially copper, gold and mercury (or ‘quicksilver’ - Hg), as materials for explosives [from conventional nitric acid to high ex-
well as alloys like bronze], for example, gave rise to an understanding plosives, like nitro-glycerine and trinitrotoluene (TNT)] and fertilizers
of the properties of their chemical compounds. Although it was only (such as nitrates; traditionally imported from Chile). Together the two
around 2000 BCE that iron was common in Egypt. Before dyes could be world wars of the 20th Century led to much more local manufacture of
applied to textiles, their fibres needed to be treated with substances chemicals and a focus, for example, on the indigenous production of
known as ‘mordants’ [18]. The most common of these was alum (a salt high explosives, gasoline from coal and eventually synthetic polymers
deposit), but this required purifying via a process of recrystallization. [20,24]. The structure of the UK chemicals industry has remained much
Common salt (sodium chloride - NaCl), which was initially used for the the same from that time onwards, although companies have come and
preservation of fish and meat [18], along with sal ammoniac, were also gone through mergers and acquisitions [20]. New products, such as
widely known minerals from ancient times. Fuller’s earth, which can act polythene, polyvinyl chloride (PVC) and nylon, were produced in the
as a natural cleansing agent was found in the Near East, whilst nitre 1950s and 1960s as a result of high levels of research and development
(from natural deposits in Babylonia) or saltpetre (potassium nitrate – (R & D). This arose from R & D expenditure that was nearly double that
KNO3) may well have been employed to make nitric acid (HNO3) [17]. in the whole of UK manufacturing sector as a percentage of sales in-
Glass production on a significant scale in Egypt was in place by about come [24]. A dramatic growth in the chemicals industry followed in the
1370 BCE, using soda (or ‘natron’ - Na2CO310H2O) from natural de- 1960s and early 1970s, until it became the mature processing industry
posits [17,21]. Crude soap was made from oil and alkali, possibly by that it is today.
burning plant material to produce potash (or ‘vegetable alkali’); any
mined or manufactured salt that contains potassium (K) in a water- 2.2. Structure of the modern chemicals sector
soluble form [20]. Elementary yellow sulphur certainly existed in natural
deposits near the Red Sea, and may have been adopted to make sul- Chemicals are a complex collection of many diverse and interacting
phuric acid (H2SO4) or ‘oil of vitriol’ [20]. Fermentation processes for sub-sectors covering a wide range of feedstocks, processes and products.
the purpose of brewing to make alcohol was in widespread use from Physical outputs are moved around on an international scale within or
antiquity [17,18,21]. The founder of commercial chemistry was later between major companies that are truly multi-national [24]. The in-
viewed as being Jabir ibn Hayyan (c. 721–815 CE), known in Europe as dustry is also highly focused on private R & D and protective of in-
‘Geber the Alchemist’ [13], who experimented from a workshop in Kufa formation, meaning that data availability is particularly poor. This high
(Iraq) on a range of chemical processes and initiated “the classification technology sector takes full advantage to modern developments in
of materials into spirits, metals and filtration” [21]. electronics and information and communications technology (ICT), such as
The fruits of Arabic chemical processing progressively migrated for the automatic control of chemical process plants and automation in
across Europe. However, the development of chemicals as they are the use of analytical instruments [24,25]. The scale of operation of
known today really commenced with the so-called Industrial Revolution chemical firms range from quite small plants (of a few tonnes per year)
in the UK from about 1760 CE onwards [17,19,23]. It focussed on the in the fine chemicals area, where high purity is required, to giant ones in
discovery of ways of bulk-producing acids and alkalis. This came about the petrochemical sector [24,25]. Batch production is employed by
from a fusion of empirical ‘rules of thumb’ with the chemical sciences SMEs where small quantities of chemicals (up to around 100 tonnes per
[21]. Chemistry itself may be traced to the first categorization of che- annum) are required. In contrast, continuous plants are typically used
mical elements by Robert Boyle (1627–1691) in 1661 [23]. Many in cases where a single output, or related group of products, are de-
practical technologies for the transformation of raw materials (or manded with plants of several thousands to millions of tonnes per year
‘feedstocks’) into useful or aesthetically pleasing artefacts were far more [24,25]. They often produce intermediates which are converted via
advanced than the understanding of the fundamental principles of ap- downstream processing into a wide range of products, such as benzene,
plied science during that period [19]. Some of the drivers behind the toluene and xylenes (BTX), ethylene, phenol, and PVC from petro-
expansion of the UK chemical industry came from the prolific increase chemical refineries or via ammonia plants [1,24].
in textile production (particularly cotton manufacture) that necessi- Central to the production of organic chemicals is the steam cracking
tated ever more bleaching agents [18], the improvement in the pro- process which manufactures lower olefins, also referred to as high value
cessing of sulphuric acid, originally by Joshua Ward (1685–1761) and chemicals (ethylene, propylene, butadiene and aromatics), from feed-
subsequently by John Roebuck (1718–1794), and then its large-scale stocks deriving from oil or natural gas [22]. The feedstock is primarily
use for the production of soda in a lead-chamber process devised by the naphtha and ethane-based, and most process fuel demand is met by
French chemist Nicolas Leblanc (1742–1806) [17,23]. UK chemical fuel-grade process by-product gases. There were four steam crackers
processing became centred on Merseyside (within a triangle bordered operating in the UK at the end of 2010 and three remaining in operation
by St Helens, Warrington and Widnes), Clydeside in central belt of today. The physical unit of production for the sector is tonnes of high
Scotland, and Tyneside (in the North East of England between New- value chemical (thvc). Feedstock input in 2010 was about 4 Mt (ac-
castle and North Shields) [17]. Soap production was largely based again cording to the UK Government's Digest of UK Energy Statistics (DUKES)
on Merseyside, where William Lever (1851–1925) established the seeds [26]) leading to an estimated production of 2.8 Mthvc. Based on the
of the multi-national conglomerate Unilever [18]. The ready and cheap feedstock mix and an assessment of European cracker specific energy
availability of sulphuric acid led to the production of soda instead of consumption (SEC) (according to the IEA) [27], process fuel demand was
potash for soap and glass manufacture [23]. Other materials associated an estimated 53 PJ and electricity demand was 1 PJ [28]. Thus, direct
with the heavy chemicals industry included ammonia (NH3), nitric acid, GHG emissions were estimated to be 2.2 MtCO2e, with total emissions

4
P.W. Griffin et al. Applied Energy xxx (xxxx) xxx–xxx

was about 2.4 MtCO2e. Table 1


Another key activity of the chemicals sector is ammonia (NH3) Ultimate yields for steam crackers.
Source: adapted from Griffin [28]. European data collated by Neelis et al. [31].
production. Ammonia is synthesised from hydrogen which is primarily
manufactured by steam reforming of methane from natural gas. The Chemical feedstock (kg product per tonne)
process is exothermic and heat is used to raise steam for other ancillary
processes. It also releases CO2 as a by-product. Two UK sites produce Naphtha Gas oil Ethane Propane Butane Othera Total
ammonia by conventional steam reforming [28]. Fuel and feedstock
High-value 645 569 842 638 635 645 688
input in 2010 was 10 PJ and 19 PJ of natural gas respectively [29] for a chemicals
production of about 857 ktNH3 [30]. Accordingly, process emissions Ethylene 324 250 803 465 441 324 489
were about 1 MtCO2e, whilst direct combustion emissions were Propylene 168 144 16 125 151 168 117
0.5 MtCO2e. Assuming an electricity requirement of 0.8 GJ/NH3 [18], Butadiene 50 50 23 48 44 50 42
Aromatics 104 124 0 0 0 104 40
total emissions from the sector were an estimated 1.6 MtCO2e.
Fuel products/ 355 431 157 362 365 355 312
Baseline process efficiencies and plant load factors adopted in the backflows
present study were largely based on information covering the region of Hydrogen 11 8 60 15 14 11 24
Western Europe. Nevertheless, the focus was on divisions 20–21 of the Methane 139 114 61 267 204 139 157
Other C4 62 40 6 12 33 62 32
UK Standard Industrial Classification (SIC 2007) for economic activities,
compo-
i.e., ‘chemicals and chemical products’ (20) and ‘basic pharmaceutical nents
products and pharmaceutical preparations’ (21). The breakdown of the C5 and C6 40 21 26 63 108 40 51
structure of the chemicals sector (SIC 20, 21), plus energy for petroleum compo-
feedstock production and CHP plant at integrated refineries, includes nents
C7 and non- 12 21 0 0 0 12 5
[28]:-
aromatics
< 430 °C 52 26 0 0 0 52 19
• Petrochemical processes > 430 °C 34 196 0 0 0 34 18
– Steam cracking (of ethylene, propylene, etc.). Losses 5 5 5 5 5 5 5
Process energy 264 261 314 249 242 264 270
– Petrochemical feedstock production (of naphtha, ethane, etc.).

Backflows to 91 170 0 113 123 91 81
Other chemical processes and miscellaneous refineries
– Identified chemical process (steam reforming, polymerisation, Share of total: 26% 3% 25% 22% 15% 8% 100%
etc.).
a
– Other (unidentified or ancillary processes, boiler plant, energy Assumed identical to naphtha based steam cracking.
overheads, etc.).
• CHP and non-CHP power plant {including major power producers that account for the bulk of energy use in the sector: a Pareto-like ap-
proach. The rest of the improvement potential in the chemical industry
(MPP)}
– Chemicals (includes one petrochemicals site and ‘other genera- was assessed by determining the possible impact of key cross-cutting
tion’). technologies [13,15,16].
– Refineries (includes two integrated petrochemicals sites).
2.3. Steam cracking of (lower) olefins
The baseline was modelled as 27 different sub-sector processes, each
with its own physical or product output; as displayed in Fig. 4 [15,28]. This process is at the heart of the UK petrochemicals sub-sector and
Here direct net energy demands (fuel, steam, and electricity) are shown produces olefins, which form the basis of many other chemicals (in-
in terms of their current {or ‘baseline’} and BPT levels for chemical cluding plastics). Olefins such as ethylene, propylene and butadiene are
processes in the UK. However, due to data restrictions, only a few of the derived from gas and petroleum feedstocks like ethane, propane, bu-
sub-processes identified were analysed in terms of their improvement tane, naphtha and gas oil. It is by far the largest energy-demanding sub-
potential. It may be argued that this level of disaggregation represents a sector in terms of both fuel and feedstock within the chemicals sector.
useful split of the UK Chemicals industrial structure that others may find DUKES [26] provides data for the consumption of petrochemical feed-
beneficial for future assessments. The procedure adopted was therefore stock from which it is possible to model the UK steam cracking process
to focus principally on a limited set of five chemical products/processes [28]. The typical yield of products deriving from each feedstock, along

Fig. 4. Direct net energy demand (fuel, steam, and electricity) of the
baseline and best practice technology (BPT) levels for chemical processes in
the UK.
Source: Griffin [28], adapted from Griffin et al. [15].

5
P.W. Griffin et al. Applied Energy xxx (xxxx) xxx–xxx

Fig. 5. Sankey energy flow diagram for the UK steam cracking process.
Source: Griffin [28].

with the proportion of raw materials consumed in 2010 is shown in assumed to be met by natural gas. Fuel efficiency was taken as 17.4 GJ/
Table 1 [28,31]. Total feedstock consumption was 4070 kt; equating to thvc; determined by recalculating the European average of 16.9 GJ/
a production of about 2800 kthvc. A Sankey diagram depicting the thvc [28], but then correcting it for the UK feedstock mix. The calcu-
calculated baseline energy flows for steam cracking is shown in Fig. 5. lation of emissions via this method provides data within 2% of the
This diagram is based on endothermicity; the theoretical amount of verified emissions figure under the European Union’s Emissions Trading
energy required in an endothermic reaction [32]. In this case, it was Scheme (EU-ETS) [28]. This estimation was based on 2010 data. The
derived from the energy released from the combustion of process fuel. average UK fuel SEC for Best Available Technologies (BAT) and Best
Estimates of efficiency, energy and carbon dioxide emissions asso- Practice Technologies (BPT) was also calculated for the UK feedstock
ciated with steam crackers are shown in Table 2 [27,28,31]. Combus- mix. The European BPT fuel SEC was 13.1 GJ/thvc [27], and so the UK
tion emissions were calculated from an estimation of process fuel de- BPT SEC becomes 13.5 GJ/thvc. According to Worrell et al. [33]), the
mand. For simplicity, it was assumed that production of by-product fuel BAT fuel SEC is 11 GJ/thvc for naphtha steam cracking and 12.5 GJ/
is equal to the approximated share as indicated by Neelis et al. [31] for thvc for ethane cracking. Assuming the average of these SECs for other
process energy covered by feedstock. The remaining process energy is feedstock flows, BAT SEC for the UK was found to be 11.8 GJ/thvc. This

Table 2
Energy and emissions analysis for UK steam cracking.
Source: adapted from Griffin [28]. European data for SEC, by-product share and by-product emissions factors from Neelis et al. [31]; readjusted SEC from the IEA [27].

Naphtha Gas oil Ethane Propane Butane Othera Total

SC fuel SEC (GJ/thvc) 17.8 20.7 21 17.6 17.1 17.8 18.7


SC fuel SEC readjusted (GJ/thvc) 16.5 19.2 19.5 16.3 15.9 16.5 17.4
SC fuel SEC readjusted (GJ/t) ethylene 32.9 43.7 20.4 22.4 22.9 32.9 24.4
SC fuel (PJ) 11.3 1.6 16.5 9.4 6.3 3.5 48.6
Share of SC fuel SEC from feedstock 100% 95% 80% 100% 100% 100% –
By-product fuel CO2 emissions factor (tCO2/t) 48.7 48.7 43.3 43.3 43.3 48.7 –
Supplementary fuel CO2 emissions factor (tCO2/t) – 50.5 50.5 – – – –
Specific CO2 emission (tCO2/thvc) 0.80 0.94 0.40 0.99 1.02 1.29 0.79
Specific CO2 emission tCO2/t ethylene) 1.60 2.13 0.42 1.36 1.47 2.57 1.11
CO2 emission (MtCO2) 0.55 0.08 0.34 0.57 0.41 0.27 2.22

Abbreviations: SC-steam cracker; SEC-specific energy consumption.


a
Assumed identical with naphtha based steam cracking.

6
P.W. Griffin et al. Applied Energy xxx (xxxx) xxx–xxx

rudimentary assessment implies that the technical improvement po- comparison, technical improvement potential for UK ammonia pro-
tential was 22–38%. duction is about 10%.

2.4. Reformate fractionation of aromatics 2.6. Ostwald processing of nitric acid

The expression ‘aromatics’ (that mainly constitutes benzene, to- Nitric acid (HNO3) is a colourless liquid that is used in the proces-
luene, and xylenes – BTX), which are characterised by double-bonded sing of inorganic and organic nitrates and nitro compounds for dye
carbon molecules [5] that can be modified easily. Their name arises intermediates, explosives (TNT), fertilizers, pharmaceuticals, rocket
from their pungent smell that stems from their extraction via the pyr- fuel, and many organic chemicals. It is most commonly produced from
olysis of ‘gasoline’/petroleum (a by-product of steam cracking) or from ammonia (NH3) via the Ostwald process. The Latvian-born, German
reformate (a product of the catalytic reforming of naphtha at re- chemist Wilhelm Ostwald developed the process, which he patented in
fineries). These processes are estimated to use an average of 2 GJ final 1902. (He subsequently received the Nobel Prize for Chemistry in 1909
energy per tonne of BTX aromatics extracted [1]. Ultimate yield of such for his research on catalysis, chemical equilibria and reaction velo-
aromatics from steam crackers with various chemical feedstocks are cities.) In this process, ammonia is converted to nitric acid in two steps.
listed in Table 1 [28,31]. BTX aromatics are critical to petroleum re- In the first step ammonia is oxidized by heating with oxygen in the
fining and the petrochemical industries [5], and demand for all three presence of a catalyst, such as a precious metal gauze [36] (e.g., pla-
products has risen rapidly in recent years. Half of all toluene con- tinum with 10% rhodiumto form), nitric oxide (N2O) and nitrogen di-
sumption is utilised as raw material in other aromatics production, i.e., oxide (NO), together with water/steam. This is a strongly exothermic
for xylene production (using ‘hydroproportionation’) and for benzene reaction that provides a useful heat source following its initiation.
production (employing ‘dealkylation’) [1]. Thus, the full calorific value Then, in the second step, the N2O is absorbed in water, which then
(45 GJ/t) must be allocated between these chemicals in order to avoid forms nitric acid (albeit in a dilute form with concentrations of up to
double counting, and the feedstock value of toluene must therefore be 69% [36]). It also reduces a fraction back to NO. This NO is then re-
corrected by the share of its consumption (by 50%) that is processed cycled, and the acid is concentrated to the required strength by dis-
into the other aromatics. tillation. Nitric acid production with state-of-the-art technology [36]
Benzene (C6H6) is one of the largest-volume petrochemicals and the induces a significant amount of energy recovery. A modern dual pres-
largest of the BTX aromatics [1,7]. It has been used as a solvent and as a sure HNO3 plant has a net energy output of 11 GJ/t HNO3-N as high-
component of motor vehicle fuel for improving gasoline quality. pressure steam, corresponding to about 2.4 GJ/t HNO3 (100%) [36].
However, 70–75% of benzene is consumed globally for the production The process evidently results in emissions of a potent greenhouse gas
of ethyl benzene; primarily for the production of polystyrene and cu- - nitrogen dioxide (N2O) [28]. Unfortunately, BAT in this area mainly
mene for phenol and acetone [7]. Its use has decreased drastically in has the effect of reducing CO2e emissions [7], rather than N2O. The NOX
recent years, because of the post-2008 economic downturn and its high concentration after the absorption section of the Oswald process mainly
toxicity. BTX aromatics are also important in the production of poly- depends on the pressure applied at that section. Low NOX emission
mers, other chemicals and several consumer products (such as solvents, levels can be achieved by high absorption efficiencies (e.g., by the ap-
paints, polishes, and pharmaceuticals) [7]. Likewise, they are used in plication of high absorption pressures) and/or by end-of-pipe technol-
health and hygiene, food production and processing, transportation, ogies. Among the end-of-pipe measures is the selective catalytic reduction
information technology and other sectors. Like most petrochemicals, (SCR) process for reducing NOX emissions. Emission reduction rates of
the consumption of BTX aromatics is strongly linked with consumer up to 95% [34] can be achieved via the state-of-the-art SCR process for
demand for plastics. Benzene is expected to grow at a rate of 3% an- new and existing plants. In new plants a combination of the SCR process
nually until 2018, while the estimated growth in overall toluene con- and high-pressure absorption (< 8 bar) are likely to result in NOX
sumption was less than 3% per year [7]. Xylene (C8H10) is actually the emissions at the level of 100–200 mg NOX/Nm3 (as NO2) [36].
name of three isomeric forms; that depend on their relative place in the According to Wiesenberger [36], Austrian nitric acid plants cur-
methyl groups. According to the JRC report [7] the consumption of rently have emission levels between 1200 and 2750 mg N2O/Nm3, de-
xylenes are expected to grow at 4.5% annually until 2020. pending on process technology (particularly the pressure level during
catalytic ammonia oxidation). The adoption of a high efficiency cata-
2.5. Steam reforming of hydrogen and ammonia lytic ammonia oxidation process will reduce this N2O formation. A
commercialised and patented process for the homogeneous decom-
Steam reforming is a method of producing hydrogen from natural position of N2O in the NH3 combustion unit is available [36] for new
gas (methane) feedstock. The steam methane reformer (SMR) separates nitric acid plants. Several test projects for developing N2O abatement
methane into hydrogen (H2) and carbon dioxide (CO2) as a by-product. processes are underway in industry and at research institutes [36].
Most ammonia (NH3) in the UK is also manufactured by steam re- Some promising results have been reported [36] on catalytic decom-
forming with methane from natural gas. In that case, the H2 is put position processes of N2O in the ammonia combustion unit. This process
through a secondary reformer, the Haber process, in which it is com- could also be retrofitted onto existing nitric acid plants. Wiesenberger
bined with nitrogen from air to form ammonia. The hydrogen for NH3 [36] suggests that catalysts for the catalytic N2O decomposition in the
production in the UK is manufactured at two sites: the Ince plants at ammonia conversion unit are already being tested on a prototype scale,
Billingham and Saltend. Ince use conventional routes at Billingham, and that they might be available for industrial use in the near future.
while at Saltend natural gas is used as feedstock for acetic acid and
acetic anhydride production. Hydrogen by-product from this process is 2.7. Mercury cell process (chlor-alkali)
subsequently fed through a small adjacent ammonia plant [34].
Natural gas for feedstock and combustion in conventional steam This is a process of producing chlorine by electrolysis of a salt so-
reforming was estimated from the UK GHG Inventory [34]. For the 2010 lution, a co-product of which is hydrogen (H2) [15,16,28,37]. It is al-
baseline, these were 17.1 PJ and 9.1 PJ respectively [28,30,21]. Pro- ternatively known as the chloralkali (chlor-alkali or chlor alkali) pro-
duction related to this natural gas feedstock was 857 kt, giving an cess. In the mercury cell process that has been favoured in Europe [37],
overall fuel SEC of 30.5 GJ/tNH3. The SEC of BAT ammonia production sodium (Na) forms an amalgam with the mercury (Hg) at the cathode.
is as low as 27.6 GJ/tNH3 and could be met by the conventional process The two metals react with the water in a separate reactor (called a
route, together with heat exchange ‘autothermal reforming’ technology ‘decomposer’), where an H2 gas and a ‘caustic soda’, or sodium hy-
(according to the European Commission [35]). Based on this droxide (NaOH), solution at 50% are produced {two moles of NaOH per

7
P.W. Griffin et al. Applied Energy xxx (xxxx) xxx–xxx

mole of chlorine (Cl)}. Solid salt is required to maintain the saturation


of the salt water, and it is usually re-circulated. The brine is first de-
chlorinated, and then purified by a precipitation-filtration process. It
yields chemical products that are extremely pure, so that the chlorine
(along with a little oxygen) can normally be used without further
purification. The mercury process uses the most electricity and also
requires measures to prevent serious environmental contamination.
However, no steam is required to concentrate the caustic solution. Na
must finally be removed from the hydrogen gas and caustic soda so-
lution.
Mercury losses have been considerably reduced over the years, as
chlorine producers have increasingly moved towards membrane tech-
nology, which has much less impact on the environment [38]. Emis-
sions for all mercury cells across Western Europe reached 0.68 g/t of
chlorine capacity in 2015. Some 20 mercury-based chlorine plants
currently remain to be phased out or converted to non-mercury tech-
nology at a cost of more than 3000 million €. These plants account for
an ever decreasing part (less than 20% in 2015) of European chlorine
capacity. Ayres and his co-workers [39,40] argued that “industrial se-
crecy” constrained the availability of the data needed to conduct proper
LCA studies of the mercury cell chlor-alkali process in Europe. They
therefore constructed a plausible emissions profile from process data
and top-down mass balances [40].
There are two other basic processes for the production of chlorine
and caustic soda from brine: the diaphragm cell (which dominated US
chlor-alkali production in the 20th Century [37]) and the membrane Fig. 6. Schematic representation of an integrated top-down and bottom-up modelling
approach for the UK industrial sector.
cell (favoured in Japan [37]). The performance of all such devices is
Source: Griffin et al. [13], adapted from by Dyer et al. [42].
governed by the requirements of electrochemical technology [37].
Membrane cells constitute the most modern process, which yields both
economic and environmental advantages. Chlorine (at the anode) and Energy Database (UED) [15,16], produced by the present authors for the
the H2 (at the cathode) are kept apart by a selective polymer membrane UK industrial sector as part of the research programme of the UK Energy
[38] that allows the sodium ions to pass into a cathodic compartment Research Centre (UKERC). Aspects of both top-down and bottom-up
and react with the hydroxyl (OH) ions to form caustic soda. The de- models were adopted, with detailed bottom-up studies set within a top-
pleted brine is dechlorinated and recycled back to the input stage. down framework. Using this approach would normally entail focusing
Overall, the energy consumption in a membrane cell process is of the on a number of sub-sectors for the bottom-up study [13], with the re-
order of 2200–2500 kWh/t, as against 2400–2700 kWh/t of chlorine for mainder of the sector being treated in a generic manner. Sub-sectors
a diaphragm cell process [38]. Moussallem et al. [41], following a re- that use a large amount of energy are obviously prioritised for bottom-
view of the development of chlor-alkali electrolysis, suggested that the up studies. In additional, sub-sectors that use energy in a relatively
replacement of the H2 evolving cathodes in classical membrane cells by homogeneous manner are easier to analyse and this may also be con-
improved cathodes would require novel electrolysis cell designs, but sidered when selecting appropriate sub-sectors. Sub-sectors that are not
might lead to the reduction of the cell voltage, and correspondingly the the subject of detailed bottom-up modelling require a focus on the
energy consumption, by up to 30%. potential reduction in emissions through widely used, ‘cross-cutting’
technologies [13,15,16].
3. Methods and materials
3.2. Modelling chemical processes
3.1. A hybrid top-down/bottom-up approach
The full range of 27 chemical processes modelled and their con-
There are two broad ways to modelling the industrial sector [13]: tribution to the sector’s overall energy demand are depicted in Fig. 4
top-down and bottom-up approaches; as illustrated in Fig. 6 (adapted [28]. The method and assumptions for building the baseline structure
from Dyer et al. [42] and Griffin et al. [13]). A top-down approach are set out in greater detail in the UED [15] and associated spreadsheets
splits industry into sub-sectors, usually based on available statistical [16], but they are summarised here. Chemical process plant capacity
data, and uses this data to determine energy use, output, energy in- data was sourced from chemical profile articles of the industry in-
tensity and other measures for which data is available. This approach telligence provider (ICIS – see < http://www.icis.com/ > and again
has the advantage of covering a large proportion of energy demand, but Griffin et al. [15,16]). Plant production levels were calculated by ap-
it is limited by the level of disaggregation available from industry-wide plying estimated load factors calculated, for the petrochemical industry,
statistical sources. Thus, the conclusions that can be drawn from such from information provided by the Association of Petrochemicals Produ-
top-down studies are often only indicative in nature. In contrast, a cers in Europe (APPE) [43]. This body publishes production and capacity
bottom-up approach would typically focus on a single industrial sub- figures of key products and derivatives in Western Europe. The average
sector. Energy use can then be separated into lower order sub-sectors, capacity of the products listed was used for other organic chemicals.
processes or manufacturing plants. The data used for this type of Inorganic chemical plant load factors were informed in part by further
bottom-up study typically comes from more specific information articles sourced from ICIS (Griffin et al. [15]). In cases where in-
sources, such as trade associations, company reports, and case studies. formation was not available, a default value of 70% was adopted. The
Such a bottom-up study can therefore be useful in terms of presenting exceptions to these methods are: (i) olefins, which were calculated by
more accurate findings [42], although it will be limited in the breadth analysing petrochemical feedstocks (extracted from the annual edition
of its application. of DUKES [26]) with data on their typical product yields from steam
A hybrid approach was employed to develop the industrial Usable cracking by Meyers [44]; and (ii) NH3, for which an estimate made by

8
P.W. Griffin et al. Applied Energy xxx (xxxx) xxx–xxx

Fig. 7. Sankey energy flow diagram of the UK Chemicals sector as modelled here; baseline data in 2010.
Source: Griffin et al. [15].

the US Geological Survey (USGS) [45] was employed. Baseline process 4. Improvement potential
efficiencies were taken from a report by the IEA [27], which relates to
Western European production. The exception to this was ammonia, for 4.1. Process improvement
which the data for combustion and feedstock use of natural gas was
sourced from UK Office of National Statistics (ONS) [46], oxygen was Identified technologies and measures generally fit into three broad
taken from the International Iron and Steel Institute (IISI) [47], and hy- categories. Fuel switching was not identified separately, but occurs via
drogen from the European Commission [48]. The sub-sector final de- many of the process substitution options. Natural gas, the least CO2
mand fuel split [32,49], which is primarily natural gas, was applied to -intensive fossil fuel, is by far the most widely burnt fuel in the sub-
the processes as an average. The sub-sector was split for the purposes of sector; much of it being used in CHP plants. In a report for the IEA [27],
determining steam and electricity demand between self-produced and a broad range of potential technological improvements for the global
imported products applied across the various processes. This necessi- chemicals and petrochemicals sector was evaluated. This study aimed
tated that an average emissions factor was employed at the process to determine which improvement option constituted the Best Practice
level. Technology (BPT) out of a portfolio of 66 of the most common processes.
The energy flows through the Chemicals sector are illustrated in BPTs represent the ‘best’ technology that is currently in use, and
Sankey diagram displayed as Fig. 7 [15]; for a 2010 baseline. The data therefore economically viable. This can be distinguished from a Best
underlying this sector model was primarily based on DUKES [26]. Available Technology (BAT), which includes proven technologies that
Special attention was given to combined heat and power (CHP) within may not yet be economically viable.
the Chemicals sector. Some CHP plants were classified as major power Although the IEA report [27] had a global remit, most of the BPTs
producers (MPP), and so would normally be outside the scope of the and the average SEC data was European-sourced, and thus provides a
final energy demand adopted by the UK Government’s statistical service reasonable representation for the UK setting. By taking the SEC data
for the purposes of compiling DUKES [26]. Here the fuel use for the and applying it to UK process outputs, it was possible to build a simple
Chemicals sector CHP and autogeneration were specified following bottom-up estimation of process energy demand and GHG emissions for
discussions with industrial representatives and personal correspon- the sector. Production from most processes was estimated by collecting
dence with the analysts at the UK Government’s former Department of plant capacity data from published sources (such as the ICIS [15,47]),
Energy and Climate Change [DECC] (see again Griffin et al. [15]). Power and then applying estimated load factors. The weighted average load
consumption was split between grid and autogenerated electricity. factor for the sector in 2010 was estimated to be 75% [28]. Exceptions
Imported heat, exported heat and exported electricity were all ac- to this approach were for the production of lower olefins and ammonia.
counted for. Approximately 60% of the electricity produced by CHP in The energy demand accumulated from the 27 identified processes
the Chemicals sector is exported; 50% to the grid and 10% to other applicable to the UK chemicals sector that have been examined in some
industrial sub-sectors. The remaining fuel demand not used in CHP or detail in the present study are again represented in Fig. 4 [28]. It can be
autogeneration is used by processes, in addition to auxiliary boilers and observed that production of high value chemicals from steam cracking
other miscellaneous equipment. The processes modelled in a bottom-up demands by far the greatest amount of energy. The curve then rises
manner account for 92% of the non-CHP fuel demand. The steam and steeply for titanium dioxide, chlorine, soda-ash and ammonia, before
electricity demand represented by a summation of the bottom-up reducing in gradient. The curve peaks before the last process as some
modelled processes accounted for 50% and 25% of the total demand processes are exothermic, and therefore net energy producers. In ag-
respectively. These may appear low as the energy carriers have higher gregate, BPTs have the potential to reduce direct energy demand from
demand for energy overheads (such as the lighting and heating of 98 PJ to 71 PJ, or save 28 PJ (28%) [28]. Primary energy reduction
buildings) and auxiliary processes associated with, but not included in, may be simply estimated on the basis of using the sector efficiency of
the chemical process demands listed by the IEA [27]. steam generation (95%) and weighted average efficiency of owned and
imported electricity. This yields figures of from 121 PJ to 88 PJ, or a
saving of 33 PJ (∼27%) [28]. Given the dominance of lower olefin
production, applying the BAT standard for steam cracking provides a

9
P.W. Griffin et al. Applied Energy xxx (xxxx) xxx–xxx

reasonable approximation for sector BAT potential. Improvement po- specifically depicted. It can be seen that these options generally perform
tential of direct and primary energy becomes 41 PJ and 36 PJ respec- well compared with the alternative process routes. Abatement from CCS
tively, or just over a third of demand. The associated reduction in GHG is more advantageous with the biomass-based process routes, because it
emissions from BPT and BAT options is consequently 1.8 MtCO2e (22%) gives rise to so-called ‘negative emissions’ (provided that carbon in the
and 2.3 MtCO2e (27%) respectively [28]. Due to the uncertainties as- product is not released at a later stage). On the other hand, these op-
sociated with using European efficiencies and in the estimation of load tions are more resource efficient and rely on established supply systems.
factors, these improvement potentials should be treated as indicative. The prospect of applying CCS is reasonable compared with other
industrial sectors, such as iron and steel [28], or cement [13,28]. All
4.2. Process substitution steam crackers and ammonia plants are situated within CCS cluster
regions (see Fig. 9 [13]). Here the distribution of CO2 point sources and
These include retrofit technologies, such as dividing wall columns potential UK CCS cluster regions are illustrated. The cost of capture
and membranes for chemical separation. The options assessed in the from ammonia is also very low because of the purity of its process CO2
present study embraced all those reviewed by Ren and his co-workers emissions. Retrofitting post-combustion capture at steam cracker fur-
[32,49] in their thorough account of long-term opportunities for the naces has been estimated to cost in the range of £28–52/tCO2 [52]. This
petrochemicals sector. They range from the replacement of steam range covers the options identified in this study with the lower limit
cracker plants with the existing state-of-the-art equipment to radically being comparable to oxyfuel capture technology and attained through
different biotechnologies presently in the R & D stage, such as bioe- the utilisation of process waste heat. Unlike oxyfuel capture, however,
thanol-to-ethylene. Many of the technologies would entail a large post-combustion treatment would not require major process modifica-
structural shift in the way petrochemicals are produced in the UK. Most tion. Irrespective of capture costs, the cost and availability of CO2
of the replacements entail a substitution in feedstock as well as fuel, and transport and storage presents an even greater challenge [53,54].
therefore have significant energy and GHG emissions ramifications for Cluster regions of industrial activities have been identified for storage
upstream fuel processing industries. These processes were also analysed under both the North Sea [54] and the Irish Sea [57] are shown in Fig. 9
by Ren and his co-workers [32,49], but were not included in this as- [13]. It is interesting to note that the main industrial regions of the UK
sessment as they are outside the scope of the Chemicals baseline. Im- remain very much as they were at the time of the Industrial Revolution
provement potential of these replacements should not be considered in (see Section 2.1 above). Thus, UK chemical processing is still largely
isolation, and to properly understand their potential would require an centred on Merseyside (within a triangle bordered by St Helens, War-
extension of the definition of industry to include the refineries sector. rington and Widnes), Clydeside in central belt of Scotland, and Tyne-
side (in the North East of England between Newcastle and North
4.3. Fuel switching Shields) [17].

The current modelling approach [15,16,28] separates a proportion 5. Petrochemicals production


of generating plant allocated to petrochemical production. This was
estimated to account for a quarter of sector heat and power generation, 5.1. The petrochemical context
as well as 30% of associated GHG emissions, based on three major sites
[28]. The high proportion of emissions was caused by the necessary use The petrochemical technologies assessed here represent options for
of refinery by-product gases. Thus, a limit is set on the degree to which either substituting existing facilities with radically different process
fuel may be switched to biomass. Steam crackers are likely to improve routes, or maintaining existing facilities and retrofitting carbon capture
in efficiency going forward, and so it was assumed that additional equipment. The representation of alternative processes in the present
surplus cracker fuel gas will be made available to CHP plant, thereby study (see Figs. 4 and 7) was largely informed from an assessment by
lowering the limit or restriction on fuel switching [28]. This particular Ren and his co-workers [32,49], whilst the application of CCS was
conflict between improvement options may be circumvented by mainly underpinned by analysis of Johansson et al. [52]. Both of these
switching to alternative petrochemical processes. However, upstream studies provide detailed techno-economic comparisons within a wide
refinery gases were assumed to remain unchanged as they will continue portfolio of improvement potential options. Many of these possibilities
to produce such gases regardless [28]. Fuel switching is also an option were modelled here with ancillary heat and power systems to meet the
for conventional generating plant, and is far less restricted in that ap- envisaged new demand [13,15,16]. The systems were treated as addi-
plication. Biomass fuel switching generally applies to coal and natural tional to existing onsite heat and power capacity in the sector, which
gas only. In the case of coal, the present hybrid model automatically are kept separate to avoid risk of double counting opportunities in the
prioritises biomass substitution before switching away from natural gas energy sector. However, the present hybrid model (see Fig. 6) allowed
[28]. for downsizing of existing facilities to compensate for new generating
capacity or, in the case of alternative processes, the option to exclude
4.4. Carbon capture and storage the additional generation capacity. Moreover, for each option ex-
amined, there was a choice as to whether the ancillary system should be
Opportunities for retrofitting post-combustion carbon capture and targeted for retrofitting capture facilities to the petrochemical process.
storage (CCS) facilities [50,51] onto existing steam crackers were as- Although these features add flexibility to the model, it is not possible to
sessed by Johansson et al. [52]. There is also significant opportunity to accurately integrate new and existing systems without more detailed
apply CCS to steam reformers [53,54]. This option is relatively cheap as and reliable data on the energy balances of individual petrochemicals
the process already produces a pure stream of CO2. It was assumed that sites [28]. This assessment also excludes the processes for manu-
the application of CCS would have a 20% energy penalty and a direct facturing alternative plastics, such as starch plastics and polylactic acid
capture efficiency of 85% [50]. The energy penalty and capture effi- (PLA), that could partially replace conventional plastics applications
ciency values were informed by the literature [50,51,55,56], and ap- (and therefore petrochemical steam cracking); again due principally to
plied conservatively owing to the mix of fuels considered in the present a lack of available technical and economic information sources [28].
UK study [28]. CCS was assumed to be applied at a generation capacity
that represents petrochemical production; reflecting the application of 5.2. Waste plastics utilisation
CCS to steam crackers. A comparison of retrofit CCS options for existing
steam crackers is illustrated in Fig. 8 [28]. Here energy and GHG It is possible to convert waste plastics, such as the polypropylene in
emissions from petrochemical steam cracker retrofitted CCS are plastic carrier bags, into naphtha and other oils. The method uses

10
P.W. Griffin et al. Applied Energy xxx (xxxx) xxx–xxx

Fig. 8. Energy and GHG emissions from petrochemical steam cracker


(SC) retrofit CCS. [A –specific energy consumption (SEC) by energy type;
B – corresponding GHG emissions, less those avoided by auto gen-
eration.]
Source: Griffin [28].

hydrogen, steam and other catalysts to produce the petrochemical route (which has been deployed in South Africa by Mossgas and Ma-
feedstocks through a stepped process of liquefaction, pyrolysis and se- laysia by Shell [49]). Nevertheless, the biomass and biomass/coal blend
paration. Originally developed by BASF in Germany, the technology options via FT processes were modelled in the present study [15,16,28]
was abandoned before reaching commercial scale [49]. According to an as they exhibit the greatest reduction in GHG emissions.
estimation by WRAP [58] - a registered UK charity and company lim-
ited by guarantee delivering waste and recycling assessments for gov- 5.4. Methanol routes
ernments and international bodies - plastic waste arisings in the UK
were about 3 Mt in 2010. Of this, around 2 Mt was packaging waste that Methanol provides an alternative route for converting methane, coal
is not collected for recycling. However, not all of this waste would or biomass into olefins. The methanol-to-olefins (MTO) route comprises
potentially be available as feedstock. The economics of the processed of three steps: methanol production, methanol conversion to olefins and
waste and feedstock together, relative to oil prices, would dictate the gasoline, and product recovery and separation [28,49]. Methanol is
likely level of substitution. The present hybrid model [15,16,28] en- produced from methane, coal and biomass by gasification at elevated
ables any substitution rate, though a diversion of 10% of the available temperatures into a syngas, and subsequently converted to methanol
waste stream (200 kt) was assumed. Naphtha consumption for steam via synthesis processes. Heat from methanol synthesis may be utilised
cracking in 2010 was about 1 Mt, which equates to a substitution rate of to convert some of the methanol into dimethyl-ether (DME) and water;
20% [28]. the former being used in the synthesis of olefins using a fluidised or
fixed-bed reactor. Recovery, separation and cooling processes are es-
5.3. Naphtha routes sentially unchanged from those used for steam cracking with the exact
yield depending on the severity, catalysts used, and reactor configura-
Naphtha can be produced from methane (natural gas), coal or bio- tion [49]. Research into MTO techniques began 20–30 years ago and
mass in various processes, including the Fischer-Tropsch (FT) process. FT two pilot plants currently operate in Norway [49]. The MTO process
naphtha is converted from methane via natural ‘gas-to-liquids’ pro- was modelled in the present study [15,16,28] via biomass gasification,
cesses or from coal via ‘indirect liquefaction’ [28,49]. Biomass may plus coal gasification installed with CCS capture equipment.
similarly be converted to naphtha via Fischer-Tropsch processes, al-
though its efficiency is influenced by the relatively high moisture 5.5. Oxidative coupling
content. Steam cracking with FT naphtha typically produces a 40%
higher ethylene yield than conventional naphtha. But, owing to an Oxidative coupling is a process by which ethylene can be converted
absence of aromatics, this produces only a 5% greater high value che- directly from methane without the intermediate steps of methanol,
mical (hvc) yield. In addition, coal and lignocellulosic biomass may be naphtha, or ethane production. Oxidative coupling of methane (OCM) is
processed together for added flexibility. Alternatively, coal may instead also referred to as ‘catalytic oxidative dimerization’ of methane or
be used in the process of ‘direct liquefaction’, although this produces partial oxidation of methane to ethylene. A fluidised bed reactor (FBR) is
aromatic-rich naphtha with a 15% lower hvc yield [28,49]. All these a common reactor design used for this process. Oxygen reacts with
routes have yet to reach full commercialisation, except the methane methane inside the reactor in the presence of a catalyst to form a methyl

11
P.W. Griffin et al. Applied Energy xxx (xxxx) xxx–xxx

Fig. 9. Distribution of CO2 point sources and CCS cluster regions in the UK.
Source: adapted from Griffin et al. [13].

radical (CH3) and water [49]. By combining together, the methyl ra- waste, and subsequent fermentation to ethanol); and lignocellulosic
dicals form ethane, which subsequently dehydrogenates into ethylene. biomass gasification with microbial fermentation or chemical conver-
The catalysts originally used were oxides of alkali, alkaline earth ma- sion with a catalyst. Detailed analysis on bioethanol production pro-
terials, or other precious earth metals, which are needed in part to cesses as part of a portfolio of ‘white’ biotechnology opportunities was
control the oxygen-ions and to maintain the reaction [49]. However, undertaken by Patel et al. [62]. Bioethanol production via lig-
novel catalysts based on research into genetically modified (GM) bac- nocellulosic biomass is still at the development stage, and has yet to
teriophages developed at MIT are presently being advanced by a San have been demonstrated at an industrial scale. The production of
Francisco based start-up [59]. One concern about this process is the bioethanol from sugar and maize already exists in production facilities
difficulty in preventing the reaction from progressing and converting worldwide, and a 0.2 Mt/yr ethanol-to-ethylene (ETE) production plant
ethylene into CO2 and water. Nevertheless, a recent breakthrough in based on sugarcane began operation in Brazil in 2010 [28]. Installed
catalyst design [28] has yielded promising results in two pilot plants and planned annual bioethanol capacity in the UK is nearly 1 Mt with
built in 2012, and a larger scale demonstration plant is likely to begin 90% derived from wheat, in contrast to 7% from sugar beet, and 3%
operation in the near future. The company has also partnered with from municipal solid waste [28]. The UK does not presently have ETE
German cracker equipment supplier in order to scale up the plant for capacity, and predominantly manufactures bioethanol from wheat
commercialisation within the next few years [60]. The reaction takes (together with some sugar beet). The feedstocks reported in the lit-
place at two thirds the temperature of steam cracking and is exo- erature consulted for the present work were principally maize and su-
thermic. However, detailed recent performance data was not publically garcane [49,62]. As the choice of feedstock can influence energy and
available for the present study, and therefore the technology modelled emissions intensity [28], these representations should be treated as only
was based on the ‘OCM I’ design developed by Ren et al. [49] from the indicative for the case of the UK. ETE based on bioethanol from starch,
work of Swanenberg [61]. sugar, and lignocellulosic biomass was modelled in the present study
[15,16,28]. In the case of lignocellulosic biomass, the option of utilising
5.6. Bioethanol routes more of the resource in CHP was also considered.

It is possible to convert biomass into ethanol for subsequent dehy- 6. Future prospects: towards a bio-economy
dration to form ethylene. There are three well-known methods for
producing ethanol from renewable sources [49]: direct fermentation of Bioenergy can be produced from either biomass (any purpose-grown
starch/sugar rich biomass (e.g., maize starch, sugar beet or sugar cane); material, such as crops forestry or algae) or biogenic waste (including
hydrolysis of lignocellulosic biomass (e.g., wheat, wood or agricultural household, food and commercial waste, agricultural or forestry waste,

12
P.W. Griffin et al. Applied Energy xxx (xxxx) xxx–xxx

and sewage sludge). Sustainable bioenergy is a renewable resource, technological development.


used together with flexible technologies, that is often low-carbon and
potentially generates ‘negative emissions’ when coupled to CCS facil- 7. UK chemical ‘technology roadmaps’ to a low-carbon future by
ities. However, virtually no bioenergy is currently used in the chemicals 2050
sector, except for the production of bio-hydrogen. The UK
Government’s UK and Global Bioenergy Resource Model (an updated 7.1. Background
feedstock availability model; awaiting open publication at the time of
writing) suggests that there is substantial quantities of indigenous A set of technology roadmaps have been developed in order to
biomass and biogenic waste available even accounting for the appli- evaluate for the potential deployment of the identified chemical tech-
cation of more stringent sustainability and land use criteria; perhaps nologies out to 2050. The extent of resource demand and GHG emis-
equivalent to some 1200 PJ. But many industrial sectors will be com- sions reduction was therefore estimated and projected forward. Such
peting for this resource alongside, for example, power generation. roadmaps represent future projections that match short-term (say out to
Indeed, in their stakeholder engagement with the UK Government, re- 2035) and long-term (2050) targets with specific technological solu-
presentatives of the chemicals industry identified what they regarded tions to help meet key energy saving and decarbonisation goals. A
as a high level of uncertainty over the supply of low-carbon bioenergy. bottom-up technology roadmap approach has been adopted, based on
A number of the chemical processes (e.g., ammonia production) pre- that were previously used by Griffin et al. [13,66] to examine the im-
sently utilise very large quantities of natural gas feedstock. In order to pact of UK cement decarbonisation (for further details see Griffin [28]).
replace it, this would need the replacement by major gasification ca- Thus, their contents were built up on the basis of the improvement
pacity, together with associated bioenergy supply. The sector believes potentials associated with various processes employed in the chemicals
that this is unrealistic in view of the cost and possibly restricted industry and embedded in the UED [13,15,16].
availability of feedstock. The focus of biomass and biogenic waste ga-
sification is therefore likely to be on burning biogas, hydrogen (repla- 7.2. Baseline UK chemical technology projections
cing steam reforming of natural gas) and syngas going forward.
However, burners designed for natural gas would need to be adjusted as The projected baseline is affected by sector output, grid dec-
gaseous bioenergy has a lower calorific value. The large number of arbonisation, and deployment of BPT/BAT. It is assumed that the grid
smaller fossil-fuelled heaters scattered around a typical chemical fac- will decarbonise by around 85% over the period 2010–2050. Owing to
tory would also make it difficult to distribute the correct fuel, or fuel the high trade intensity of the sector and high dependency on fuel
mixture, to given burners. Pulverised biomass or biogenic waste could prices, it is impossible to project forward with high confidence the fu-
be adopted for high temperature process heat, but that would give rise ture trend in petrochemicals production. Investment is going into shale
to dust deposition on heater tubes and an increase in air pollution gas imports into Britain, as well as potential local shale gas extraction,
(possible beyond legal limits). that will be sent to the Grangemouth refinery complex in Scotland; in
Lower olefins primarily serve as the building blocks to polymers for large part to produce high value chemicals (hvc) [4,63,67]. However, the
the production of plastics, rubbers and fibres. For example, more than extent to which shale gas will affect the market over the short-medium-
half of ethylene production is polymerised into polyethylene with most term is uncertain [4]. Based on planned capacity closures, and the 2010
of the remainder forming precursors of other plastics, such as polyvinyl load factor, lower olefin production has been an estimated 2.4 Mthvc in
chloride (PVC), polystyrene, polyester, and so [28]. A lot of research 2015 [28]. For simplicity, and in the absence of more detailed in-
attention has been given to the potential for substituting these products formation, hvc production levels were assumed to remain at roughly
with bio-based polymers. There are three main approaches to producing this level into the future. Similarly, the production from other chemical
bio-based polymers [28]: (i) modifying natural polymers (e.g., starch products modelled in the present study [15,16,28] were assumed to
plastics); (ii) the manufacture biomass-derived monomers by fermen- remain at their 2010 levels. It has been presumed that progress towards
tation or conventional methods and polymerise), or directly synthesise the take-up of BAT will be gradual, and reach the 2010 BPT level by
from micro-organisms; and (iii) via genetically modified crops. The UK about 2050 [28]. Reductions in steam and electricity demand are
appears to have the technical expertise and feedstock capabilities [63], thought not to lead to significantly reduced generation of heat and
as well as the downstream industrial demand, to develop a bio-based power in the chemicals sector, but rather to increase exports to the
polymers manufacturing industry. public distribution system. Efficiency improvements via CHP plant were
The UK produces some 2.5 Mt of primary plastics, or resins, per year not directly assessed here, due largely to uncertainty about the impact
and nearly 5 Mt of converted plastics, such as packaging, downstream of fuel switching. General energy saving methods, such as improved
[28]. It is an importer of primary plastics and consumption to the tune motors, better boiler efficiency, and so on, were assumed to steadily
of 5 Mt [28]. Domestic production of bio-based plastics is presently very reduce sector energy demand by 10% in 2050. This reduction was taken
low with no large-scale production facilities, although processors have as being similar across fuel, steam, and electricity demand.
indicated a willingness to expand their use of bio-based plastics and
already import them [29]. A feasibility study commissioned by the 7.3. Scenario definition
National Non‐Food Crops Centre (NNFCC) suggests that a wheat-based
200 kt/yr polylactic acid plant could be commercially viable in the UK The identified improvement technologies for the UK were in-
market. Britain also produces the feedstock for ethanol-to-ethylene corporated into a chemical technology roadmap framework through a
(ETE), as indicated in Section 5.6 above, or bio-based polypropylene via series of scenarios. The baseline year for the framework was taken as
its growing bioethanol sector, and is well placed to produce wheat 2010. Full details of the both the 2010 baseline and the BAT/BPT im-
starch for starch-based plastics [28,64]. However, the industry is pre- provements can be found in the UKERC industrial UED [15,16]. Four
sently in its infancy, although wheat supply could be increased through future scenarios were devised in order to demonstrate this approach.
the utilisation of set-aside land that might yield an additional 2.1 Mt/yr The chemical industry has been active in the area of technology road-
[64]. In the long-term, a further 0.5 Mt/yr could be made available mapping, particularly at the global level via the IEA [1], and ideas from
from the conversion of temporary grassland, whilst another 2 Mt/yr such roadmaps were drawn on in constructing some of the scenarios
could yield higher intensity crop rotation [64]. Global production ca- detailed below [13,28,66]:
pacity for bio-based polymers is expected to grow substantially from
800 kt in 2012 and to over 5 Mt in 2017 [65]. Appropriate public • Low Action (LA). This scenario describes a path of only slight im-
support and incentives would be required in order to push forward provements. No further investment is made in additional process

13
P.W. Griffin et al. Applied Energy xxx (xxxx) xxx–xxx

Fig. 10. GHG emission pathways associated with the technology roadmaps
for the UK chemicals sector (on a credit basis).
Source: Griffin [28].

technology improvements, and efficiency is only improved in- alternative lower olefin production routes, as modelled for example by
cidentally through the replacement of retired plants. Ren et al. [32], are excluded in order to avoid conflict with existing
• Reasonable Action (RA). All identified efficiency technologies are generation plant. That also provides a simpler basis for comparison. As
installed by 2025, and retired plants are replaced with best practice electricity and heat are exported from the sector, it is useful to analyse
ones by 2030. the associated emissions by what would otherwise be emitted from
• Reasonable Action including CCS (RA-CCS). This scenario is based public distribution systems. Due to commercial restrictions on data
on RA, but includes CCS. Biomass co-firing with CCS may, of course, availability it was not possible to accurately determine the extent of
mitigate upstream emissions on a full life-cycle basis, due to po- sector autogeneration and heat trading over the earlier years. Instead
tential ‘negative emissions’ [68]; something that will need careful these graphs are used as a way of comparing the future potential of the
study in future studies. constructed roadmaps and their relationship with the wider UK system.
• Radical Transition (RT). This scenario explores a boosted or radical Indirect emissions presented in Fig. 10 were estimated on a ‘credit basis’
version of the reasonable action (without CCS) scenario [28]. [28].
Graphs B and C in Fig. 10 indicate that heat and electricity is ex-
7.4. Alternative UK chemical technology roadmaps ported from all pathways with emission factors calculated from their
generation. Roadmaps with lower GHG emissions intensity of genera-
The chemicals sector GHG emission pathways are illustrated in tion therefore have a lower emissions trajectory. The baseline only re-
Fig. 10 (on a ‘credit basis’ [28]) for the assessed roadmaps and the quires a 10% biomass demand for CHP, and so it benefits by exporting
projected baseline. Graph A includes direct emissions (scope 1), this heat and electricity to other users to whom the higher GHG emis-
whereas Graph B includes direct and indirect emissions (scope 1–2/3). sions factor would elseways be attributed. Conversely, the 80% biomass
Graph C includes GHG emissions from the combustion of lower olefins CHP roadmaps show a narrower improvement on their counterparts
(on top of Graph B emissions). Additional autogeneration plant for and all CCS roadmaps, for which emission is captured from auxiliary

14
P.W. Griffin et al. Applied Energy xxx (xxxx) xxx–xxx

generation plant producing surplus electricity, appear less attractive. the emissions avoided through switching to biomass or biogenic waste
An indication of the relative degree of decarbonisation of the grid and derived feedstock. Depending on the scope of analysis, carbon fixation
district heat systems may also be deduced. Biomass demand highlights from biomass growth may be subtracted, or GHG emissions from end-
the greater levels necessary for radical abatement under the Radical product incineration may be assumed zero. In order to reduce emissions
Transition (RT) set of roadmaps as shown in Fig. 10. RT3 gives a similar via waste recycling, the waste stream must be diverted from degrada-
abatement to RT5 [bio-CHP], but requires much larger amounts of tion in landfill or incineration.
biomass. RT5 would be a far more effective way to utilise this level of The prospect of applying CCS [50–57] is reasonable compared with
consumption. Thus, Graphs A and B show that the RT roadmaps abate other industrial sectors, such as iron and steel [28], and cement
less than the Reasonable Action including CCS (RA-CCS) roadmaps; [13,28,66]. All steam crackers and ammonia plant are situated within
especially so on a credit basis [28]. However in Graph C, which includes CCS cluster regions (see again Fig. 6) [13,28]. Pipeline technology for
GHG emissions from the combustion of hvc, the RT roadmaps which use building a CO2 transport network is ready to be rolled out, and the UK
biomass as a feedstock display significantly improved, relative perfor- already has preliminary plans for at least two large transport ‘hubs’ that
mance. In particular, RT3 (lingo FT naphtha) and RT5 (bio-ETE) abate will eventually be centrally located amongst a cluster of CCS power
more than the CCS roadmaps in Fig. 10. The use of biogenic waste- stations and industrial sites. The cost of capture from ammonia is also
derived naphtha, which is limited to replacing 20% of naphtha feed- very low, because of the purity of its process CO2 emission. Retrofitting
stock, shows only a marginal net improvement in Graph C. post-combustion capture at steam cracker furnaces has been estimated
The trajectory of the petrochemicals sector were determined against to cost in the range of £28–52/tCO2 [52]. This range covers the options
the higher and lower carbon caps of the EU-ETS (although they are not identified in the study with the lower limit being comparable to oxyfuel
shown here – the interested reader should consult Griffin [28]). Because capture technology and attained through the utilisation of process
the scope of the petrochemicals sector must include feedstock produc- waste heat. Unlike oxyfuel capture, however, post-combustion would
tion to enable comparison between lower olefin production routes, caps not require major process modification [53].
are proportionally adjusted up from the GHG emissions verified for
steam cracking. No constraint specific to this scope of emissions exists Acknowledgements
under the EU-ETS as the allocations are necessarily site-based. How-
ever, this study has provided an indication of the potential for meeting This is an extended and updated version of a paper originally pre-
the cap. Direct emissions, as indicated in Fig. 10, do not fully account sented at the 8th International Conference on Applied Energy (ICAE2016)
for the whole picture. It can therefore be deduced that the EU-ETS held in Beijing, China over the period 8–11 October 2016 (denoted then
unfairly incentivises the use of retrofit CCS over bio-based process as paper ICAE2016-560). The work reported forms part of a programme
routes [28]. This is because CCS specifically addresses emissions at the of research at the University of Bath on the technology assessment of
point of installation. energy systems and transition pathways towards a low-carbon future
that has been supported by a series of UK research grants and contracts
8. Concluding remarks awarded by various bodies associated with the Research Councils UK
(RCUK) Energy Programme for which the second author (GPH) was the
The opportunities and challenges to reducing industrial energy de- holder. That associated with industrial energy demand and carbon
mand and carbon dioxide (CO2) emissions in the Chemicals sector have emissions reduction originally formed a part of the ‘core’ research
been evaluated with a focus is on the situation in the United Kingdom programme of the UK Energy Research Centre (UKERC); Phase 2,
(UK), although the lessons learned are applicable across much of the 2009–2014 [under Grant NE/G007748/1]. The first author (PWG) and
industrialised world. This sector can be characterised as being quite third author (JBN) undertook their contributions to the present work as
heterogeneous, and as sitting on the boundary between energy-in- part of a UKERC flexible funding project entitled ‘Industrial Energy Use
tensive (EI) and non-energy-intensive (NEI) industrial sectors (see again from a Bottom-up Perspective’ [for which the second author (GPH) was
Fig. 2). Currently-available technologies will lead to further, short-term the Principal Investigator]. The second author (GPH) was also a Co-
energy and CO2 emissions savings in chemicals processing, but the Investigator of the UK Biotechnology and Biological Sciences Research
prospects for the commercial exploitation of innovative technologies by Council’s (BBSRC) Sustainable Bioenergy Centre (BSBEC) during
mid-21st century are far more speculative. There are a number of non- 2009–2013, as part of the ‘Lignocellulosic Conversion to Ethanol’
technological barriers to the take-up of such technologies going for- (LACE) Programme [under Grant Ref: BB/G01616X/1]. During the
ward. Consequently, the transition pathways to a low-carbon future in preparation of this paper, the second (GPH) and third (JBN) authors
the UK chemicals industry by 2050 will exhibit large uncertainties. The continued to work in the field of industrial energy use and carbon
attainment of significant falls in carbon emissions over this period de- emissions reduction; supported by the UK Engineering and Physical
pends critically on the adoption of a limited number of key technologies Sciences Research Council (EPSRC) ‘End Use Energy Demand’ (EUED)
[e.g., carbon capture and storage (CCS), energy efficiency {including Programme, as part of the Centre for Industrial Energy, Materials and
combined heat and power (CHP)} techniques, and bioenergy], alongside Products (CIE-MAP) [under Grant EP/N022645/1], as a Co-Director
a decarbonisation of the electricity supply [13]. Thus, this technology and Research Fellow respectively.
assessment and associated roadmaps help identify the steps needed to The authors' names are listed alphabetically.
be made by industrialists, policy makers and other stakeholders in order
to ensure the decarbonisation of the UK chemicals sector. References
The chemicals sector has long been the largest owner of generating
plant in UK industry. Most generation is from CHP plant with sig- [1] International Energy Agency (IEA), International Council of Chemical Associations,
nificant amounts of excess electricity exported to the grid or other in- DECHEMA. Technology roadmap: energy and GHG reductions in the chemical in-
dustry via catalytic processes. Paris (France): IEA Publications; 2013.
dustrial sectors. Special care should be taken not to ‘double count’ auto- [2] Jacquelin L-M, de Bucy J, Caujolle A. Facts & figures: energy and carbon for green
generation and grid decarbonisation, and the relative contributions to chemistry. Paris, France: ENEA Consulting; 2015.
decarbonisations of each should be accounted for separately. [3] International Energy Agency [IEA]. World energy outlook: are we entering a golden
age of gas? Special report. Paris (France): IEA; 2011.
Replacement process technologies (e.g., bio-processing and methanol- [4] Hammond GP, O’Grady A. Indicative energy technology assessment of UK shale gas
to-olefins) cannot be properly compared without consideration of up- extraction. Appl Energy 2017;185(2):1907–18.
stream feedstock processing. This is because they replace both the [5] Saygin D, Patel MK, Tam C, Gielen DJ. Chemical and petrochemical sector: potential
of best practice technology and other measures for improving energy efficiency.
stream cracking process and production of petroleum feedstocks, e.g., Paris (France): OECD/IEA; 2009.
naphtha and ethane. Special care should also be taken when modelling

15
P.W. Griffin et al. Applied Energy xxx (xxxx) xxx–xxx

[6] International Council of Chemical Associations (ICCA). Innovating for greenhouse available techniques (BAT) reference document for the production of chlor-alkali.
gas emission reductions: a life cycle quantification of carbon abatement solutions JRC Science for Policy Report (JRC91156/ EUR 26844 EN). Luxembourg:
enabled by the chemical industry. Brussels (Belgium): ICCA; 2009. Publications Office of the European Union; 2014.
[7] Boulamanti A, Moya JA. Energy efficiency and GHG emissions: prospective sce- [38] World Bank Group. Chlor-alkali plants. Pollution prevention abatement handbook.
narios for the chemical and petrochemical industry. JRC Science for Policy Report Washington (DC, USA): World Bank Group; 1998. p. 279–328.
(JRC105767/EUR 28471 EN). Luxembourg: Publications Office of the European [39] Ayers R. The life-cycle of chlorine, part I: chlorine production and the chlorine-
Union; 2017. mercury connection. J Indust Ecol 1997;1(1):81–94.
[8] The European Chemical Industry Council [Cefic]. European chemistry for growth: [40] Ayres RU, Ayres LW. The life cycle of chlorine, part II: conversion processes and use
unlocking a competitive, low carbon and energy efficient future. Brussels (Belgium): in the European chemical industry. J Indust Ecol 1997;1(2):65–89.
Cefic; 2013. [41] Moussallem I, Jörissen J, Kunz U, Pinnow S, Turek T. Chlor-alkali electrolysis with
[9] European Climate Foundation [ECF]. Europe’s Low-carbon Transition: oxygen depolarized cathodes: history, present status and future prospects. J Appl
Understanding the challenges and opportunities for the chemical sector. Brussels Electrochem 2008;38(9):1177–94.
(Belgium): ECF; 2014. [42] Dyer CH, Hammond GP, McKenna RC. Engineering sustainability: energy efficiency,
[10] Norman JB. Industrial energy use and improvement potential PhD Thesis Bath (UK): thermodynamic analysis and the industrial sector. Environ Eng (J Sustain Environ
University of Bath; 2013 Eng, Inst Engin, Australia) 2008;9(2):17–22.
[11] Dyer CH, Hammond GP. C.I. Jones CI, McKenna RC. Enabling technologies for in- [43] Association of Petrochemicals Producers in Europe [APPE]. Capacity and produc-
dustrial energy demand management. Energy Policy 2008;36(12):4434–43. tion data (online) < http://www.petrochemistry.eu/about-petrochemistry/facts-
[12] Hammond GP. Industrial energy analysis, thermodynamics and sustainability (In and-figures/capacity-and-production-data.html > [accessed 5 August 2013].
memoriam: Willem van Gool). Appl Energy 2007;84(7–8):675–700. [44] Meyers R. Handbook of petrochemicals production processes. New York (USA):
[13] Griffin PW, Hammond GP, Norman JB. Industrial energy use and carbon emissions McGraw-Hill Education; 2004.
reduction: a UK perspective. WIREs Energy Environ 2016;5(6):684–714. [45] US Geological Survey [USGS]. Mineral commodities summaries: nitrogen (fixed) -
[14] Hammond GP, Norman JB. Decomposition analysis of energy-related carbon ammonia. Reston (VA, USA): USGS; 2012.
emissions from UK manufacturing. Energy 2012;41(1):220–7. [46] Office of National Statistics [ONS]. Annual business survey: section C – manu-
[15] Griffin P, Hammond G, Norman J. Industrial energy use from a bottom-up per- facturing. Newport (UK): ONS; 2012.
spective: developing the usable energy database (Beta Version). Report UKERC/ [47] International Iron and Steel Institute [IISI]. Energy use in the steel industry.
WP/ED/2013/002, London (UK): UK Energy Research Centre; 2013 < http://data. Brussels (Belgium): IISI; 1998.
ukedc.rl.ac.uk/cgi-bin/dataset_catalogue//view.cgi.py?id=15 > [accessed 20 [48] European Commission [EC]. Integrated pollution prevention and control (IPPC):
October 2014]. best available techniques (BAT) reference document for iron and steel production.
[16] Griffin P, Hammond G, Norman J. Industrial energy use from a bottom-up per- Seville (Spain): EC; 2012.
spective: usable energy database (spreadsheet - beta version). London (UK): UK [49] Ren T, Patel MK. Basic petrochemicals from natural gas, coal and biomass: energy
Energy Research Centre; 2013 < http://data.ukedc.rl.ac.uk/cgi-bin/dataset_ use and CO2 emissions. Resour, Conserv Recycl 2009;53(9):513–28.
catalogue//view.cgi.py?id=15 > [accessed 20 October 2014]. [50] Hammond GP, Ondo Akwe SS, Williams S. Techno-economic appraisal of fossil-
[17] Buchanan RA. Industrial archaeology of Britain. Harmondsworth, UK: Penguin fuelled power generation systems with carbon dioxide capture and storage. Energy
Books; 1972. 2011;36(2):975–84.
[18] Butt J, Donnachie I. Industrial archaeology of the British Isles. London, UK: Paul [51] Hammond GP, Spargo J. The prospects for coal-fired power plants with carbon
Elek; 1979. capture and storage: a UK perspective. Energy Convers Manage 2014;86:476–89.
[19] Derry TA, Williams TI. A short history of technology: from earliest times to A.D. [52] Johansson D, Sjöblom J, Berntsson T. Heat supply alternatives for CO2 capture in
1900. London: Oxford University Press; 1960. the process industry. Int J Greenh Gas Control 2012;8:217–32.
[20] Williams TI. The chemical industry: past and present. 2nd ed. Wakefield: Ep [53] Collodi G, Wheeler F. Hydrogen production via steam reforming with CO2 capture.
Publishing; 1972. Chem Eng Trans 2010;19:37–42.
[21] Al-Hassani STS, Woodcock E, Saoud R, editors. 1001 inventions: muslim heritage in [54] Element Energy. Potential for the application of CCS to UK industry and natural gas
our world. 2nd ed.Manchester (UK): Foundation for Science Technology and power generation [Final Report for the UK Committee on Climate Change].
Civilisation; 2005. Cambridge (UK): Element Energy; 2010.
[22] Al-Khalli J. Pathfinders: the golden age of Arabic science. London, UK: Allen Lane; [55] Intergovernmental Panel on Climate Change [IPCC]. IPCC special report on carbon
2010. dioxide and storage. Cambridge: Cambridge University Press; 2005 [Prepared by
[23] Cossons NBP. The book of industrial archaeology. 2nd ed. Newton Abbot (UK): Working Group III of the IPCC].
David & Charles; 1987. [56] International Energy Agency [IEA]. Cost and performance of carbon dioxide capture
[24] Heaton CA, editor. An introduction to industrial chemistry. 2nd ed.London (UK): from power generation. Paris (France): IEA; 2011.
Blackie Academic & Professional; 1991. [57] Department of Energy and Climate Change [DECC]. 2012. CCS roadmap: sup-
[25] Royal Society of Chemistry [RSC]. Industrial chemistry case studies. London (UK): porting deployment of carbon capture and storage in the UK. London (UK): DECC;
RSC; 1998. 2012.
[26] Department of Business, Energy and Industrial Strategy [BEIS]. Digest of United [58] Waste & Resources Action Programme [WRAP]. UK plastics waste – a review of
Kingdom energy statistics. London (UK): BEIS; 2016 [annual]. supplies for recycling, global market demand, future trends and associated risks.
[27] International Energy Agency [IEA]. Chemical and petrochemical sector: potential of Banbury (UK): WRAP; 2006.
best practice technology and other measures for improving energy efficiency. Paris [59] Bullis K. Chasing the dream of half-price gasoline from natural gas. MIT Technol
(France): IEA; 2009. Rev (online) 15 January 2014 < https://www.technologyreview.com/s/523146/
[28] Griffin PW. Radical change in energy intensive UK industry PhD Thesis Bath (UK): chasing-the-dream-of-half-price-gasoline-from-natural-gas/ > [accessed 15
University of Bath; 2013 January 2017].
[29] Office for National Statistics [ONS]. Energy use by industry - source and fuel [60] Tullo AH. Siluria’s oxidative coupling nears reality. Chem Eng News
1990–2012. London (UK): ONS; 2014 [Spreadsheet]. 2014;92(27):20–1.
[30] Webb N, Broomfield M, Cardenas L, MacCarthy J, Murrells T, Pang Y, et al. UK [61] Swanenberg GMJM. Cogeneration of ethylene and electricity with oxidative me-
greenhouse gas inventory 1990 to 2011: annual report for submission under the thane coupling: study of technical feasibility and economic merit. Delft (The
framework convention on climate change. Report Ricardo-AEA/R/3355. Didcot Netherlands): Stan Ackermans Instituut; 1998.
(UK): Ricardo-AEA; 2012. [62] Patel M, Crank M, Dornburg V, Hermann B, Roes L, Husing B, et al. Medium and
[31] Neelis M, Patel M, De Feber M. Improvement of CO2 emission estimates from the long-term opportunities and risks of the biotechnological production of bulk che-
non-energy use of fossil fuels in the Netherlands. Utrecht (The Netherlands): micals from renewable resources - the potential of white biotechnology The BREW
Department of Science, Technology and Society; Universiteit Utrecht; 2003. Project Utrecht (Netherlands): Utrecht University; 2006
[32] Ren T, Patel M, Blok K. Olefins from conventional and heavy feedstocks: energy use [63] The Knowledge Transfer Network [KTN]. From shale gas to biomass: the future of
in steam cracking and alternative processes. Energy 2006;31(4):425–51. chemical feedstocks. Horsham (UK): KTN; 2016.
[33] Worrell E, Price L, Neelis M, Galitsky C, Zhou N. World best practice energy in- [64] Barker M, Safford R. Industrial uses for crops: markets for bioplastics. London (UK):
tensity values for selected industrial sectors Report LBNL-62806 Berkeley (CA), UK: HGCA; 2009.
Lawrence Berkeley Laboratory; 2008 [65] European Bioplastics. Bioplastics: facts and figures. Berlin (Germany): European
[34] Department for Energy and Climate Change [DECC]. GHG inventory research: non- Bioplastics; 2013.
energy use of fuels. London (UK): HMSO; 2007. [66] Griffin PW, Hammond GP, Norman JB. Prospects for emissions reduction in the UK
[35] European Commission [EC]. Integrated pollution prevention and control (IPPC): cement sector. Proc Inst Civil Eng: Energy 2014;167(3):152–61.
reference document on best available techniques for the manufacture of large vo- [67] Hammond GP, O’Grady A. The life cycle greenhouse gas implications of a UK gas
lume inorganic chemicals - ammonia, acids fertilisers. Seville (Spain): EC; 2007. supply transformation on a future low carbon electricity sector. Energy
[36] Wiesenberger H. State-of-the-art for the production of nitric acid with regard to the 2017;118:937–49.
IPPC directive. Monograph 150 (M-150). Vienna (Austria): Umweltbundesamt [68] Kruger T, Darton R. Negative emissions technologies could become the world’s
GmbH (Federal Environment Agency Ltd); 2001. largest industry. Proc Inst Civil Eng: Civil Eng 2013;166(2):51.
[37] Brinkmann T, Giner Santonja G, Schorcht F, Roudier S, Delgado Sancho L. Best

16

You might also like