Process System Innovation by Design
Process System Innovation by Design
Proefschrift
door
Samenstelling promotiecommissie:
G.P.J. Dijkema
p/a Delft University of Technology
P.O. Box 5015
2601 GA Delft
The Netherlands
e-mail: [email protected]
All rights reserved. No part of this publication may be reproduced, stored in a retrieval
system, or transmitted, in any form or by any means, electronic, mechanical, photocopying,
recording, or otherwise, without the prior permission in writing of the proprietor.
Gerard Dijkema
Voorburg, July 2004
Contents
PREFACE...................................................................................................................V
CONTENTS ...........................................................................................................VII
1 INTRODUCTION................................................................................................1
1.1 OVERVIEW: WHAT, WHY AND HOW?....................................................................1
1.2 READER'S GUIDE .................................................................................................1
1.3 A SYSTEM IMAGE OF THE PETROCHEMICAL INDUSTRY ..................................... 2
1.4 INDUSTRY DEVELOPMENT, INNOVATION AND SUSTAINABILITY ..................... 17
1.5 A NEED FOR PROCESS SYSTEM INNOVATION? ..................................................29
1.6 RESEARCH APPROACH AND STRUCTURE OF THESIS .........................................34
REFERENCES .......................................................................................................227
ABBREVIATIONS .................................................................................................245
APPENDICES.........................................................................................................249
SUMMARY ..............................................................................................................279
SAMENVATTING..................................................................................................283
CURRICULUM VITAE..........................................................................................289
1 Introduction
1 Where in process system engineering (PSE) the focus is on single plant process systems, the
definition of 'Process system innovation' is based on the notion of layered networked process
systems, which structure and content can be modified at each level.
2 Process System Innovation by Design
New Resources
Resource Improved Extraction
Novel Deposits
Products
Product New or improved
Services
RPMT-
Combination
availability of crude oil for the production of naphtha. Since system structure and
technology may change because of changing market demand and because of
changing resource availability, we prefer the label Pmt-R to emphasize the
relationship between product markets, technology and resource use for the system.
In the petrochemical industry, many Pmt-R's are linked, as it is a complex system of
interconnected processes and process routes. The entire petrochemical industry,
however, can be modelled as a black box that is part of and contributes to our
industrial society via the inputs it requires and the outputs it generates (Figure 1-2).
Inputs are required for day-to-day system operation, for the erection of new facilities
and for rejuvenation or end-of-life system abatement (Grievink 1994). Physical inputs
comprise feedstock and utilities, construction material and equipment. Other inputs
are market information, skilled labour, technological know-how, system design,
project organisation and management skills, investments and working capital. Physical
outputs include products, by-products, waste and emissions as well as industrial plants
and complexes. Other outputs are know-how and services rendered, salaries, fees, taxes
and dividends paid.
Non-Physical Surroundings
Operating Cost
Skilled Labour
Regulations
Investment
Know-how
Emissions
Feedstock
Products
Utilities
Waste
System: The Petrochemical Industry
(Dis)comfort
Experience
Interface
Revenue
Salaries
Surroundings
Non-Physical Surroundings
services affect system economics (e.g. Peters and Timmerhaus 1991). Inputs from
and outputs to the physical surroundings result in adverse ecological effects, such as
resource depletion, toxic waste and harmful emissions. These have resulted in
societal concern, economic effects, environmental regulation, scientific interest,
technology development and shifting design criteria and changing design space (e.g.
Carson and Darling 1962; World Commission on Economic Development 1987;
Smith 1995; van Breda and Dijkema 1998; Wood 2001).
The industry's present ecologic impact is largely determined by its system content
and characteristics, which results from past decisions on product and technology
development, process system design, management, economics and regulation. In
competitive markets, the economic health of individual companies depends on
information and knowledge on how and when to properly bring novel products and
innovative production systems into being (e.g. Wei et al. 1979; Pisano 1997; Linse
2002; van Klaveren 2003). Company prosperity requires effective use of current
assets -technology, systems, skills and organisation- and strategic capabilities to
anticipate market dynamics, shifting societal concerns and changing regulation to
ensure sustainability. The system image of the petrochemical industry therefore
includes a summary of present developments in its surroundings.
1.3.4 PRODUCTS
The chemical industry produces intermediates that are used in many industrial
sectors to cater for consumer demand (Figure 1-3). Most chemical products are not
sold to consumers, however, and the product spectrum of the industry remains
relatively unknown to the general public.
Production
Pharma
Oil Network
Winning
Refineries Fine Chemicals Food
Textile
Olefins Polym. Moulding Paper
Consumers /
Packaging Households
Steam- Polym. Shaping
Gasoil
Cracker Electronics
Production Monomers
Complexe network Automobiles
Naphtha
s
Building
Aromatics Polym. Process
Materials
Machinery
Petrochemical industry Other
Figure 1-3: 'Chemical industry', 'petrochemical industry', 'polymer Industry' and society.
7
8 Process System Innovation by Design
Many oil fields co-produce associated gas or 'condensate' -ethane, propane and
butanes-. Natural gas wells often co-produce natural gas liquids (NGL) -butanes to
heptanes-. Both condensate and NGL are premium feedstock for steam crackers.
The butanes are the preferred feedstock for MTBE.
Natural gas3 is important as a carbon and hydrogen source for industry, as it is the
dominant feedstock for methanol via the production of synthesis gas (Figure 1-3).
In the Netherlands the large-scale process industry - petroleum refining,
petrochemical industry, polymer industry, fertilizer production and base metal
production - together account for over 30% of the end-use of fossil energy resources
(Dijkema et al. 1995). Combined, these sectors represent no less than 67% of annual
Dutch industrial energy end-use, which includes feedstock use and net energy use
for production. Worldwide some 10-15% of annual fossil resource use is consumed
by these industries. These data indicate that the Dutch economy is characterised by a
relatively large process industry sector, which accounts for some 1020 PJ4 end-use
on a total Dutch energy account of some 2700 PJ in 1990 (3100 PJ in 2001) (CBS
1992; 2002)56.
With the exception of natural gas, some offshore oil production and a limited share
of renewables, the Netherlands is completely dependent on imported crude oil and
coal for its internal energy use and the associated industrial production.
3 Hydrogen and nitrogen are converted to ammonia (NH3) in the Haber-Bosch process (Anonymous
1991; Jennings 1991). This is the industrial route for nitrogen fixation from air, and the basis of the
world nitrogen fertiliser industry. The primary source for hydrogen is natural gas (CH4).
4 Pèta Joule = 1015 Joule; One PJ is the energy content of 24 Thousand Tons Oil Equivalent (TOE).
Thus the net end-use of fossil resources in the Dutch process industry equates to some 24 Million
TOE; Total Dutch fossil resource use equates to some 74 Million TOE.
5 Since for many products of the process industry, fewer than three companies operate facilities in the
Netherlands, the Dutch Statistics Office CBS cannot disclose a detailed break-down between
feedstock, energy-use and production. The industry associations (VNCI, FediChem, FOI) do report
total sector energy use, but these figures exclude feedstock consumption. Based on the case studies
done in 'de grondstoffenstudie' (Dijkema et al. 1994a; 1995) we estimate the total amount of fossil
resources that does not end up in petrochemical products in the Dutch petrochemical industry at
350-400PJ.
6 Other use of fossil fuels in the Netherlands concerns energy conversion for electric power
generation (some 480-500PJ for 300-320 PJe), mobility (380-400PJ), space-heating (400-420PJ) and
miscellaneous industry and services (600 PJ).
Introduction 9
marketed anymore for heavy transportation fuel (ships) or road construction. In the
future, it can be expected that additional conversion technology is required to
further upgrade these streams, and capture sulphur, heavy metals, and ashes in an
environmentally acceptable way. In other large-scale processes, notably base metal,
low-value or waste by-product generation remains the rule, as well as in coal
processing.
Today, in response to global treaties some classes of harmful products, such as
chlorinated hydrocarbons and CFC's, have largely been phased out. Inadvertent
emission losses of volatile organics (VOC) have been reduced. Design and
performance of fired-equipment has been improved to reduce VOC, NOx, CO and
soot emissions. Thus, in the petrochemical industry, the environmental “sanitation”
of the production systems is considered to be on track, and harmful emissions are
continuously monitored, whilst process performance is continuously improved.
As in waste management, three decades of development and corporate experience
with environmental management systems should enable the industry to respond
appropriately to whatever environmental demand or regulation on harmful
emissions. The reduction of global human CO2 emissions, however, presents a
relatively new and formidable challenge. Production and combustion of oil products,
natural gas, petrochemicals or polymers inevitably yields CO2. Thus its formation is
linked to the industry's feedstock and energy supply and to the use and final
degradation of its products. Hence, allowed CO2 emission can become an active
constraint, and CO2 emission control a key to the industry's license to grow.
1.3.7 ECONOMICS
The traditional and major chemical economies are the US, the EU and Japan. In
general, OECD countries “have rather similar chemical industries with only minor
variations between these industrialized nations” (Mol 1995: 127). Today, growth
regions are the Middle East, South-East Asia and to a lesser extent Eastern Europe.
Around the globe, the petrochemical industry is characterised by
- the use of large-scale to very large-scale continuously operated facilities
- large sunk costs - a huge amount of capital invested
- links with both cyclical and largely non-cyclical markets
- oligopolistic market structure
- a high level of internalisation, both in trade and production sites.
- a dependence on abundant, low-priced resources
- being capital intensive, labour extensive and knowledge intensive
- a relatively low return on investment (ROI) or return on asset capital employed
(ROACE) (Linse 2002)
(e.g. Wei et al. 1979; Stobaugh 1988; Brennan 1998; Whitehead 2000).
The volumes of many undifferentiated commodities merit continuously operated
facilities over batch wise operation because of achievable process efficiencies, asset
utilisation and investment per unit product. Investment for new continuous chemical
plants generally is given by I = I0 * (C/C0)p, where I is total plant Investment, C is
plant capacity, and suffix 0 denotes some known investment I0 at a known capacity
Introduction 11
C0. Since p has a value of around 0.6, a chemical plant of twice a nominal capacity
will cost only 50% more (Chauvel 1981). Generally it will require the same number
of operating personnel, and have similar overhead costs.
Whilst in (pseudo)-commodities, investments are huge (100-1500 million US$ per
facility), the profit margin or value added per unit product over feedstock cost is
relatively low (e.g. Wei et al. 1979). This margin, however, must be sufficient to allow
recovery of all operating costs and the investment-related costs, i.e. average costs.
Since petrochemical plants often exhibit a long technical lifespan -30-50 years-
would-be competitors in an established petrochemical market do face ‘marginal
competition’. Established manufacturers ('incumbents') compete on the basis of the
marginal cost of their facilities only because the original investment is a 'sunk cost'
that either has been recovered or written off in previous operating years.
Investments in new facilities, however, are justified only when sufficient return over
total cost can be realised. As a consequence, incumbents may operate economically
at efficiencies considerably below state-of-the-art.
The gap between marginal revenue and average revenue often can only be closed by
combining innovative process technology and economies-of-scale. Existing facilities
are expanded and modernized or phased out; in newly built plants state-of-the-art
technology is used and capacity is at the technically feasible design limit. Thus,
knowledge intensity, scale-of-operation and investments continuously increase to
bring down cost per kilogram of product.
The typically low margins of petrochemical operations must be sufficient to realise
return-on-investment targets. These target levels are relatively low in the
petrochemical industry - around 10% -. The financial risk, however, is substantial
due to the cyclical nature of markets, the risk of overcapacity and the technically and
economically limited flexibility of petrochemical plants in capacity-utilisation
(Kuipers 1999).
A typical petrochemical plant has a physical turndown ratio of some 50%, whilst the
average economic turndown ratio of a new plant may be as small as only 10%. Thus,
usually utilisation-factors at or below 80% indicate that the owner is not recovering
average cost. In the early 1990's, for example, many companies invested in plants
for PET - the plastic used for sodas sport drinks bottles, 'single service disposable
containers' - and its precursor, purified terephtalic acid because of expected high
market-growth. A few years later, however, the resulting capacity build-up grossly
exceeded real market growth. As a consequence, PET/TPA producers suffered
from low capacity utilisation and depressed market prices around the turn of the
century.
The PET/TPA market is an example of cyclicality, where the supply-demand
balance can only be restored by continued market growth or by phase-out of existing
capacity. At times of limited market growth -or decline- closure of facilities may only
be expected when their marginal cost cannot be recovered anymore and revamping
cannot be justified. Since capacity build-up was realised in a few years, most capacity
is in modern plants with similar performance. The inevitable result is price erosion
from which all players suffer.
In order to reduce risk, over time many major petrochemical companies have stated
they prefer to be number one or two in their respective markets or else exit.
12 Process System Innovation by Design
continuous operations that have become the blueprint image of the petrochemical
industry in the industrialised countries.
Such large-scale continuous plants are complex installations in appearance. Indeed at
a detailed technical level, they consist of a variety of interconnected apparatus,
instruments, pipes, and construction. As stated, however, any petrochemical plant
can be represented as a system where a number of physical inputs are transformed
into a number of outputs. As far as the representation of system content or design is
concerned, each petrochemical plant can be completely characterised by a single
configuration of a limited number of process sections or functional units, such as
feed-preparation, reaction, separation, final product purification, process flows,
process recycle flows and utilities (Montfoort et al. 1989; Dijkema et al. 1998). In
Figure 1-4 this is illustrated by a simple general implementation of a chemical
process. The core of the process is some reaction subsystem, which often consists of
multiple reactors. Since most chemical reactions do not give 100% yield of the
desired product only, a separation subsystem is required to achieve product
specification. As a matter of course, intermediate separation in multi-reactor systems
is a possibility.
A particular characteristic of continuous processing is the recycle structure, which in
its simplest form is the recycling of unconverted feedstock back to the reactor. In
multi-reactor / multi- separation systems, the recycle structure is an important
degree-of-freedom in system design.
Feed T Product
Recycle Flow
Since in refining and the petrochemical industry distillation is the dominant choice
for product separation, the image of this industry is formed by the well-known arrays
of distillation towers. Especially at night petrochemical plants offer a spectacular
sight when the pipes, distillation tower, reactors, vessels and a whole range of other
equipment are illuminated. Second in dimension to the distillation towers are the
huge cracking reactors in the refining industry, the largest being the Exxon
Flexicoker and Shell Hycon reaction systems, followed by continuous cat crackers
14 Process System Innovation by Design
and hydro crackers. In the petrochemical industry, industrial furnaces together with
their chimneys have developed into installations of enormous physical dimensions.
These also are the prime exception to the relatively small physical size of reactors in
the petrochemical industry: steam cracker furnaces and reformer furnaces are fired
reactors for the production of crude ethylene and synthesis gas respectively. In
contrast, a plant for the production of styrene, the reactor only is of limited size.
Chemical plant
Category
System
Chemical plant Chemical plant Element
System
Surroundings The Petrochemical Industry Boundary
Table 1.2: Overview of aggregation levels in and around the petrochemical industry7
System Chemical Plant Petrochemical (Inter)National
Boundary Industry Industry
Surroundings Petrochemical National Industry World Material
(dominant) Industry Cycles
System element Unit Operation Chemical Plant Industry Sectors
Flows: Materials, Energy, (Exergy)
Inputs, Outputs: Resources, Products, Waste, Emissions
Exchange Flows: Intermediate Products, By-Products, Energy
Intensive properties of Flows: Composition, Pressure, Temperature
As with the designs employed for methanol or propylene oxide production, the
present configuration is only one instance of a very large set of possible system
structures (Table 1.3). Where single plants are characterised by a combination of
'once through' interconnectivity and intra-connectivity via recycling between reaction
and separation, the petrochemical complexes and the industry predominantly exhibit
'once-through' intra-connectivity from feedstock to product. The majority of chains
originate from the steam cracker and re-unite somewhere in the polymer industry,
e.g. where polyurethanes are produced and moulded into car seat-cushions (Ch.4).
In the petrochemical industry, the major chemical activities are the chemical
rearrangement and partial oxidation of hydrocarbons. Thus it has become an
extremely fast element of the world's fast carbon cycle8 that is driven primarily by
life-on-earth. Biomass grows and decays, which maintains a balance of carbon as
CO2 in the atmosphere. A limited amount of biomass is temporarily extracted from
the cycle by fossilisation. Fossil resource extraction for use in power generation,
industry, transportation and households has resulted in a rapid reintroduction of
carbon fossilized over millions of years and which has induced a steep increase in
atmospheric CO2 content (IPCC 2001).
7
Currently, a great many (petro-)chemical plants are part of an industrial cluster. Thus, the
petrochemical industry' system element may also be defined as a petrochemical complex, with the
possibility in some cases the element or 'complex' only comprises of a single petrochemical plant.
8 The slow carbon cycle concerns geochemical processes where over millions of years erosion releases
carbon. In this slow cycle invertebrae sequester carbon by the formation of shells composed of lime
(CaCO3). The calcination of lime for cement production thus has become part of this slow carbon-
cycle.
Introduction 17
Solar Radiation
Sources
Interconnected
World >> Economic System Material Cycles
Inputs
'Extracted'
Sources
Nations >> Industrial Complexes
Outputs
'Emitted or
Wasted'
Space Occupied
Process System Innovation by Design
Introduction 19
In addition, economic growth stagnated, and Dutch industry was still focused on the
maritime industry and trade. By some it was considered ‘backward’ in comparison
with the rest of Europe, notably Germany, France and Great Britain. Gas
production from coal, and the work-up of the coal-tar by-products were important
activities, with around 180 factories in operation around 1900. The products, coal
tars and its aromatic distillates, were largely exported because vested interests of
natural dye producers prevented the set up of a Dutch synthetics dye industry. The
only surviving coal-tar refinery today, Cindu in Uithoorn was established in 1922.
The 20th century - interbellum. The development towards the petrochemical industry
known today started after World War I, when a transformation from agriculture-
based products to ‘chemical-based products’ started. In the Netherlands chemical
operations of well-known companies such as DSM, Akzo, and Shell Chemicals were
started. Before World War II, the Netherlands remained a net importer of chemical
intermediates.
World War II - second oil crisis. After the war, industry successfully developed because
of the Marshall plan, the national industrial policy adopted to rebuild the
Netherlands and the forced reduction of competing German coal tar refineries. In
order to understand the present structure of what is known today as the
petrochemical industry, the post World War II period is most important.
The war initiated a large number of product innovations demanded by the military.
As a consequence, both production facilities and new products became available for
the civil sector. A second major development was the rapid-transformation from
coal-based to crude-oil based operations.
According to Chauvel and Lefebvre, the transition from coal to crude oil was mainly
initiated by the continuously growing demand for ethylene, the precursor of three
important plastics, viz. polyethylene, polyvinyl chloride and polystyrene (cf. Ch. 4).
Initially, ethylene was produced from coal, mainly via the processing of coke-oven
gas, a by-product of large-scale steel production (Chauvel and Lefebvre 1989a).
Although the pyrolysis of light petroleum fractions9 was developed in the U.S. for
the production of ethylene as early as 1920, coal-based supply for ethylene and other
chemical products became rapidly insufficient after the World War II. In addition,
economics were favourable at the time and large newly discovered oil reserves led to
an abundant supply of cheap raw material. Industrial plants could be realised at
lower investment costs for the installations whilst delivering better quality products.
At the time, chemical science and engineering rapidly developed and provided novel
chemical synthesis routes enabled by new or improved catalysts. Industrial process
plant development benefited from advances in process technology that range from
reactor technology to separation and rotating equipment. As a consequence, for
most petrochemical products, by the end of the sixties this transition was complete.
Since the Netherlands did not have a large coal-based chemical industry, the growing
demand for petrochemical products opened an opportunity for competition with the
German industry, which historically was largely based on coal, and continued to be
based on coal-derived aromatic products for a long time (Molle and Wever 1984). In
the sixties and early seventies, double-digit growth continued in the chemical
there has been a strong focus on minimising investment and initial time-to-market,
in order to obtain revenues as quickly as possible. Thus there has been a strong
emphasis on efficient organisation of the project execution from conceptual design
to start-up. Concepts used include, amongst others, 'concurrent engineering' to
shorten project execution elapsed time, 'front-end-loading, which is a strategy to
amass all crucial information and elucidate white-spots as early as possible and not to
allow any changes after an early date in the project and 'value engineering', where
project items are scrutinised for value-creation potential versus cost (Herder 1999).
R& D
Invention
Feedback
Feedforward
Innovation
Market
Public <development> Market Orientation
<commercialisation> Orientation
Orientation <Business-to-Business >
<Consumers>
Diffusion
<customer acceptance>
<societal acceptance>
Feedback Feedback
Performance
Figure 1-7: Framework for the innovation process in the petrochemical industry; adapted
from (Grunert et al. 1997: 3).
development. Its products remain largely invisible and unknown to the general
public.
Where consumers thus interact only indirectly with the industry, a variety of other
stakeholders directly influence its change and innovation. Apart from shareholders,
competitors and would-be competitors include the public at large, governments,
non-governmental organisations, pressure groups and universities. Necessitating a
'Public orientation' (Figure 1-7), petrochemical companies have realised they must
ensure their license-to-operate and respond to environmental, safety and
sustainability concerns. Together with sustained value creation, these factors
determine long-term competitiveness and continuity and thereby have become
drivers for innovation by the industry and its knowledge and service-providers such
as universities, technology licensors and engineering contractors.
The changes or innovations in the industry vary in level of detail, scope and time-
scale. Production volume change, for example involves day-to-day utilisation
adjustment, small plant changes, incremental capacity increase by revamping, and
closure or erection of facilities, while for many products the market stimulates
sustained capacity increases. Short-term competition drives day-to-day operation and
organisation; long-term competition involves process innovation, which leads to
phase-out of currently employed facilities. Increasing public awareness of impact of
suspect chemicals combined with cumulating scientific knowledge and evidence may
call for the phase-out of particular products, production processes or production
routes. In other cases, product innovation leads to successful substitute competitive
products, and total phase-out of current facilities.
1.4.5 SUSTAINABILITY
With the 1989 publication of ‘Our common future’, the worries about the status of
our planet and worldwide society were coined by the caption of a need for
‘sustainable development’, which in Brundtlandt’s definition is (economic)
development and consumption that does not hamper the needs of future
generations (World Commission on Economic Development 1987). Welford argues
that ever since
There exists a strange and fruitless search for a single definition of
sustainable development amongst people who do not fully understand
that we are really talking here of a process rather than a tangible
outcome. This search is most apparent amongst positivist researchers
who grope for a hard core of definitions and data that they can
manipulate to produce simple solutions and singular answers to very
complex concepts. Such simplifications cannot exist in the post modern
world and they simply hide a scientific research bias, which is not
appropriate to a highly political issue such as sustainable development
(Welford 1997).
Although the above is valid for the economy at large, we do believe that for the
petrochemical industry some criteria can be formulated. As we have argued, today
harmful emissions and the wastes produced by the petrochemical industry have
become 'managed problems', with the exception maybe of SO2 and NOx emissions
associated with burning fuel oil and natural gas. It is the large quantities of fossil
resources processed that will become the future problem of the industry. At the
output side, large quantities of CO2 are emitted to atmosphere. At the input side the
industry depends on crude oil and natural gas. In the near term future, we expect
society to require continued availability of petrochemical products and derivatives. It
may not before long, however, before these will have to be produced from much
more expensive fossil resources, under more stringent arrangements with respect to
the fate of CO2 generated, or from alternative feedstock altogether.
But how has the petrochemical industry responded to the demand for sustainability?
In retrospect, the petrochemical industry has done a good job in improving its day-
to-day operations as well as technology and systems portfolio. The benefits from
increased resource utilisation provided sufficient incentive for R&D efforts, which
indeed have led to a substantial reduction in the consumption of fossil resources per
unit of product. Meanwhile, however, the industry has demonstrated almost
26 Process System Innovation by Design
continuous growth since 1950, and as a consequence net fossil resource use
increased dramatically.
Although we labelled the petrochemical industry as being furthest in embarking on
strategies for sustainable development(Mayer and Dijkema 1999), there is no
evidence that suggests that these strategies involve drastic process innovation. On
the contrary, whilst from 1989 onwards there has been a renewed growth in the
world-wide petrochemical industry and its consumption of fossil resources, the
number of successful petrochemical process innovations expressed as novel
industrial processes that were commercialised has declined for three decades
(Satterfield 1991). Despite the continued interest in the use of biomass and coal, the
increased demand for the building-blocks of commodity-plastics has largely been
met by the construction of an increasing number of steam crackers that use naphtha
or gas oil for feedstock. Exceptions are the first methanol-to-olefins cracker project
(Dijkema and Kuipers 2001) and Fischer-Tropsch synthesis that uses coal- or natural
gas-derived synthesis gas. Since their economic life-time is long, the fossil resource
demand and the products that result from these installations will dominate the
industry for decades. Despite process and product improvements, e.g. by the use of
new catalysts or separation technology, the net industries' share of fossil resource
consumption will remain at a high level.
Presently, in the commodity markets of the petrochemical industry, it is hard to
foresee that innovative products will be developed short-term that will completely
replace current products. Quite the contrary, many long-established products still
demonstrate considerable growth rates. As a consequence, improvement of the
sector largely has to be fuelled by process innovation. But according to Abernathy
and Utterback, the rate of such innovations eventually will fall off as prospective
incomes fall (Abernathy and Utterback 1978). Indeed, in the last decade the sector
appears to increasingly emphasize organisational and even inter-organisational
innovation. Product-chain management and sustainability are elements of the
Responsible Care (RC) programme, a strategic communication programme that first
has been developed by the petrochemical industry of Canada. The concept of RC
expresses the notion that a chemical company is part of its environment.
Consequently it must act responsibly and precautionary towards the environment
and society. The RC-programme encompasses the aspects of health, safety,
environment, welfare and responsible entrepeneuring. As stated, however,
Responsible Care is primarily a strategy for adequate communication with all the
stakeholders of a chemical company, in society, politics, and business. It also helps,
however, to provide arguments such as that “we have done enough, others outside
our company must now take action first“, or as one respondent remarked, a
mechanism to buy time, so that investments required for environmental
improvement can be matched with the right time for business (Mayer and Dijkema
1999). Thus, Responsible Care and similar programmes can create both incentives
and disincentive for the much needed process innovations.
In the Netherlands, the petrochemical industry has negotiated voluntary agreements
on energy-efficiency improvement, and it appears that the petrochemical industry is
on-target with respect to the objective of 20% energy reduction (VNCI 1998). These
agreements largely focus on the net ‘utility’ energy efficiency. As stated, however, in
Introduction 27
significantly, the OECD predicts that around 2015-2020 the time horizons for
availability of oil and gas resources will start to shrink (OECD 1999).
After the Gulf War (1991), abundant oil-supply by eager producing countries
combined with free market conditions drove down oil prices to historically low
levels. This development appears to support Grubbard's scenario where oil-price
remains relatively low, until the moment where the time-horizon of perceived
shortage becomes comprehensible (Grubbard 1992). Then oil-prices rise to levels
never seen before, largely because the perceived abundance of oil not only results in
low energy prices and limited exploration and development of oil fields, but also to a
lack of economic incentive for R&D and innovation to reduce and change our
energy use. Indeed, currently OPEC and other producing countries control their oil
production to manage the supply-demand balance to maintain oil-prices and
optimise revenues. Other scenarios have been developed, however, wherein such
managed oil prices rise slowly but steadily in response to awareness of a definite
dwindling of known economic oil reserves. These scenarios recognise that price
increases create additional incentive for exploration and for the development of
previously uneconomic resources. Managed crude oil availability and managed price-
levels, however, still adversely affect investments in and R&D into improved energy
resource utilisation and alternative energy sources. Long-term, however, it is certain
we will run out of relatively cheap oil and gas and that alternative energy sources
must be developed.
As a matter of fact, crude oil already is being replaced by natural gas as the world's
most important energy source. Researchers of IIASA11 investigated and modelled
the evolution of the use of energy carriers by humankind. They modelled the use of
a particular energy carrier by it's share f as a function of time on a logarithmic scale,
where (f/(1-f)) exhibits linear growth for a number of decades, peaks, and
subsequently declines linearly (Marchetti and Nakicenovic 1979; cited in Spiro and
Stigliani 1996: 54). Historic data for the market share of wood, coal, crude oil and
natural gas were found to be in fair accordance with this model.
During the industrial revolution coal replaced wood as major energy carrier. In the
20th century oil replaced coal. Currently, the share of natural gas is growing to replace
crude oil as the major energy carrier. History indicates that energy carriers start being
replaced when the most economic sources have been depleted. Thus, due to the
availability of fossil resources and ongoing biomass depletion, the share of biomass
in worldwide energy supply has dwindled. Initially, during the industrial revolution it
was replaced by coal. Since World-war II, the use of biomass and coal declined due
to substitution by oil products and natural gas. Nuclear fission for electric power
generation started to develop in the sixties and seventies, and at the time nuclear
fission and fusion were expected to become the next major energy source for
humankind.
The inevitable changes in fossil resource availability and price represent major
incentives for innovation in the petrochemical industry. To ensure its continuity as
the industry's infrastructure, current feedstock must be used even more efficiently;
Whilst the depletion of fossil resources and the threat of global-warming have led
the public, governments and NGOs to believe that dramatic change is required in
industry to achieve sustainability, only few companies have acknowledged that
sustainability is crucial for their long-term prosperity (Dijkema and Mayer 2001a).
The Dutch Scientific Council for Government Policy recently has categorized CO2
emission, climate change, biodiversity and the continuously growing energy-use as
wicked problems (Wetenschappelijke Raad voor het Regeringsbeleid 2003). These are
labelled wicked because to date these have largely escaped (inter)national policy and
regulation. Upon a series of wake-up calls on environmental pollution (Carson and
Darling 1962), in the past 30 years many tough environmental problems have been
turned into manageable problems. The interplay of end-of-pipe measures, technology
development, process improvements, investments, environmental legislation and
covenants has fuelled a continuous process of improvement. Wicked problems also
must be addressed effectively to enable sustainable development. Effective
communication on wicked problems is difficult, however because there is no direct
exposure to the associated adverse effects that develop gradually over decades in a
global context. Moreover, vested interests are huge, access to cheap energy being
pivotal to the prosperity of many an economy or private company.
Drastic reduction of resource consumption and CO2 emission is a wicked problem
for the petrochemical industry because the core business of the industry is the
conversion of petroleum products, its CO2 production is linked to its present use of
these fossil feedstock, and most companies compete in global markets where
opinions on CO2 and global-warming vary. Thus, despite efficiency gains, the
industry's CO2 emission has remained high largely because of the continuous growth
of production volume, which is illustrated by the doubling of ethylene production in
the past 25 years (see Ch. 4).
A sustainable petrochemical industry is characterised by appropriate resource
selection, effective resource utilisation, the avoidance of waste and emissions and a
product slate that fosters a sustainable society.
While “innovative design and clever process modifications are essential to the
economic health of the producer of commodity chemicals” (Wei et al. 1979: 255),
companies or sectors differ in their degree and development with regard to
innovation strategies for sustainability. Amongst others via environmental
management and technological development the petrochemical industry has reduced
its harmful emissions. In addition, the industry has realised that it must continuously
renew its license-to-operate obtained from local communities and other
stakeholders. Through the Responsible Care programme, therefore, the industry has
been developing its communication and image with local communities and the
general public (VNCI 1998). There are a few 'early adopters' or trendsetters that
have realised and acknowledged that sustainability is crucial for their long-term
continuation.
Some companies have formulated a sustainability strategy in interaction with external
stakeholders, others perceive sustainability as a continuous development process, in
which the specific goals repeatedly need redefinition and reformulation (Mayer and
Dijkema 1999). In the petrochemical industry, notably substantial CO2 emission
reduction, however, cannot be realised by environmental management only. Good-
Introduction 31
We argued that while end-user markets are catered for by newly innovated products,
the products of the petrochemical industry are probably here to stay for a prolonged
period. To meet sustainability demands, therefore, and to produce alternative
products, innovation of the 'T', the process systems, is required. Presently, however,
economically, the industry is in a mature phase. In a response to perceived
shareholder pressure, shareholder-driven CEOs lead the industry on an ever-
accelerating pace of mergers and acquisitions to capture the benefits of economy-of-
scale. At the same time R&D budgets are being cut or activities in
(pseudo)commodity markets are divested.
An image of current business climate emerges that is unfavourable for process
innovation initiated from within the industry. This dichotomy between industry
32 Process System Innovation by Design
climate and need for global sustainability demonstrates the relevance of our central
research theme:
The systemic, top-down specification of the desired technological content
for innovation in the petrochemical industry in relation with business
and societal needs.
Preferably, an approach results that includes and links system analysis and
subsequent synthesis and conceptual specification of options for R&D: process
system innovation by design.
12 Each chemical plant in a particular cluster may be seen as the result of an individual company's
strategy. Each company innovates its slate of technologies employed and its organisation for efficient
realisation of large-scale facilities. Once completed, these represent huge sunk-costs, have very long
technical lifespan, and require large number of highly trained and skilled personnel. The facilities are
'sticky', and while expansions or additions to the cluster may cease sometime, existing facilities remain
in operation ofter for decades before closure. Thus, the dynamics-of-change for the cluster are slow.
Introduction 33
We conjecture that in order to prepare for the future of the petrochemical industry,
integrated innovation at all system levels (Table 1.3) is required, from material to
apparatus to plant to petrochemical complex and industry configuration.
Furthermore, the industry system structure must be considered a degree-of-freedom
in the search and specification of yet unknown technological developments, as well
as in the adoption of extra-sector innovations.
Our hypothesis is that there exists scope for improvement of the petrochemical
industry by process innovation, not withstanding the dramatic improvement of the
chemical industry’s performance with respect to economic and ecological efficiency.
Secondly, our proposition is that this scope for improvement cannot be reaped by
progress in the existing disciplines of chemistry, physics and engineering alone, but
rather that a large source of innovation remains unexplored: process system
innovations.
The research by Pavitt (1984) demonstrated that traditionally innovation has come
from within the sector, whilst presently the capacity for R&D within the sector is
reduced. This dilemma of stimulation of process innovation in a mature industry has
been explored from a technological perspective. This has been addressed under the
third research question:
(3) what is the usefulness and scope of application of the methods developed?
Do the methods result in innovative concepts at industry, complex or plant
level? Is the application of outside sector innovations supported or
facilitated?
1.5.5 RELEVANCE
In the past, systems in the industry , such as entire chemical plants or steam cracker
complexes (see Table 1.3, p.17), were dramatically improved. Presently, however, the
34 Process System Innovation by Design
5.2
1.1 Straightforward
Model
1.2 Modeling &
5.1 Decomposition
Strategy
1. Select / Develop
Model
4.1 1.2
5.4
3.4 2.3
4.1. Systematic
Search R&D
options
3. Design / Compile 2.1 Losses,
4.2. Specification
of Innovation Efficiency
Assessment
Content
Procedure
x.y
Legenda: Activity Objective Result Info-label Information
from x to y
Chapter 4 is a case study into the worldwide system around olefins, where ethylene
ranks first, with global production capacity exceeding 100 Million Tons per annum.
Use of the methods developed enables the specification of process system
innovations.
Chapter 5 demonstrates the use of the method in investigating innovative use of fuel
cells in the chemical industry. Notably, the early evaluation of system concepts
developed is addressed. This is expanded into a treatise on the design of
trigeneration systems.
Chapter 6 is a case study into the development of an industrial cluster beyond the
traditional petrochemical industry. The advent of novel technology, again fuel cells,
is used to assess the impact on petrochemicals and the global platinum material
cycle. A number of process system innovations are specified.
Chapter 7 contains final conclusions and recommendations.
2 Process System Innovation Sources
2.1 Introduction
The objective of our research was the development of a systemic approach to
specify the desired content of process system innovations, which “… are defined as
changes in the system structure or system design of the petrochemical industry, its
industrial complexes, or individual plants. These can be enabled by technological
inventions or vice versa” (Dijkema et al. 2003). In this chapter, the scope and
foundations are explored of such 'process system innovation by design'.
First the past and present sources for inventions relevant for the chemical industry
are reviewed. These range from basic science, chemistry and chemical engineering to
top-down system and policy studies. Since it is impossible to review this massive
literature base, an eclectic overview is presented. Thus, the usefulness and potential
is analysed of relevant knowledge domains.
Subsequently because process system engineering was identified as a source of
building blocks for said systemic approach, the vast body-of-knowledge of system
theory and process system engineering is explored to underpin and define the
concepts of 'system', system elements, system analysis and system design.
Third, results from a literature search are reported to unearth concepts and methods
useful for process system innovation by design. Specific attention is given to system
decomposition and to process synthesis.
2.2.1 OVERVIEW
Past innovations in the chemical industry may appear to be the result of a chaotic
process. In reality, the process of change and innovation is driven by public R&D
policy and funding, corporate R&D programmes to ensure continuity by sustained
competitiveness and by individual scientist's curiosity and preferences.
Scientific literature often only reveals the results of research programmes, projects
and initiatives. In addition, corporate R&D strategies and goals generally remain
beyond scrutiny, although some results do land in the literature or patents.
What follows, therefore is a structured overview of 'innovation sources' classified in
categories relevant for process system innovation in the petrochemical industry.
Apart from serendipity, innovations relevant for the petrochemical industry may be
seen to originate or emerge from
1. basic scientific developments
Virtually all disciplines of the natural sciences and the engineering sciences may
yield inventions that allow conceptualisation of new unit operations,
improvement of existing technologies and realisation of novel reactions towards
new or existing products. Within specific scientific or application domains,
exploratory research attempts to go beyond what is already known.
38 Process System Innovation by Design
chemistry plays a pivotal role, but it is chemical engineering science that has shaped
the industry through the development of catalyst systems, chemical reactors,
separation equipment, and integrated continuous process designs.
Innovation of chemical factories at first proceeded at a slow pace. The products
developed by 17-19th century chemists were directly put to work in simple chemical
factories. From 1880 onwards the development of chemical processes and plants
rapidly evolved ‘from art to science’. In Europe and the US chemical engineering
programs were started, and in 1923 chemical engineering was firmly established by
the publication of the “classic text ‘Principles of Chemical Engineering’ (Walker and
Lewis, 1923)” (cited in Freshwater 1997). Academic chemical (process) engineering
science was rapidly recognized as essential to carry out chemical processes safely and
competitively. Its application for the transformation of laboratory-scale synthesis
routes to industrial scale facilities became common practice.
Today, the Institution of Chemical Engineers defines chemical engineering as 'the
design, development and management of a wide and varied spectrum of industrial
processes' (IChemE 2003). In chemical engineering, methods and tools have been
developed for scientific underpinning of the understanding, analysis and design of
chemical operations at industrial scale. Continuously operated systems were
developed, initially in crude oil refining, ammonia production, sulphuric acid and
other processes. Today, chemical engineering has developed into a myriad of sub
disciplines which range from the microscopic level of underlying transport
phenomena in reactors, to the macroscopic level of equipment and plant
specification. The Ullmann encyclopaedia of chemical technology (Anonymous
1995b) now comprises 26 volumes, and hundreds of journals have been established
on every sub discipline of chemical engineering.
One fundamental concept of chemical engineering is that any process plant is
considered a structured collection of unit operations, the first paradigm for the
discipline (NRC 1988). Sub disciplines of chemical engineering focus on single
classes, or even subclasses, of unit operations, such as reactors or separation
systems, which have specific links with their underlying basic sciences. Examination
of unit operations from a more fundamental point of view became a second
paradigm for the discipline. In addition, some unit operations remain largely the
realm of mechanical engineers, e.g. 'rotating equipment' such as compressors and
pumps, whilst other such as industrial furnace and heat-exchanger design require
specialization and are increasingly supported by Computational Fluid Dynamics
models.
In the NRC report one distinguished between the microscale, mesoscale and
macroscale. At the microscale, among other molecular reactions and heat and mass
transfer phenomena are studied. At the mesoscale, the selection of a particular set of
unit operations, the specification of each unit operation and the design of the
structure between the unit operations is the realm of process design.
Conceptual process system design involves the analysis and anticipation of process
plant performance, as well as process synthesis to meet plant design objectives in a
sound and safe combination of unit operations. At the macroscale, entire plants
must be developed that satisfy the needs and requirements of a complex
environment that may require integration of multiple plants and synthesis of supply-
40 Process System Innovation by Design
chains, enterprise modelling and life-cycle models (NRC 1988; Li and Kraslawski
2004).
Notably the conceptual design of chemical plants was established as a topical sub
discipline after the seminal publication by Douglas, who put forward a structured
method to arrive at a design of a chemical plant (Douglas 1988) based on and
bundling the experience of generations of process design engineers. Today
conceptual process design is part of the body-of-knowledge of process system
engineering, which is elaborated in section 2.4.
13 The Dutch Office of Statistics, the CBS executed this study that was commissioned by the Ministry
of Vrom and RIVM. The study was completed with the help of DHV, TNO and Interduct, the Delft
University Clean Technology Institute.
14 The use of LP and related techniques in conceptual process system design is addressed in Ch. 5.
42 Process System Innovation by Design
the characteristics of new processes and products that lead to their eventual
adoption in the economy.” In order to answer these questions, they developed
and employed a linear, economic input/output model that was used to establish
the optimal structure of the petrochemical industry with respect to production
costs. Out of an extensive data catalogue on industrially applied processes, the
1977 (and 1940, 1950, 1960, 1970) structure of the industry could be
reconstructed. By changing the available technology catalogue, the impact of
novel processes becoming available was assessed. In all cases, the existing
industry proved to be a barrier to selection of novel processes, which was
modelled by applying capacity constraints to the model selections.
We adapted their model to investigate the use of novel technologies in existing
process networks (Dijkema et al. 1997). Optimisation of C4 networks in
retrospect confirmed maturity of this part of the industry. We concluded that the
existing industrial process network has a large flexibility towards feedstock
economics, which implies that new processes must offer significant advantage to
overcome this barrier to implementation. These kind of studies appeared
particularly useful to establish the scope for innovation and for some
prioritisation of R&D efforts.
4. Scenario studies to elucidate the need for resource and energy conservation.
The work reported by the ‘Club of Rome’ (1972) was the first to address the
resource situation of our global industrial society using the system dynamics
modelling approach. The energy sector, base metals, and the petrochemical
industry were explicitly recognized and modelled as an important sector
(Meadows et al. 1972). The Club of Rome warned for major resources becoming
unavailable for society in the foreseeable future. In contrast with common belief,
however, the report did not state a definite horizon but merely pointed out that
the modelling exercises gave similar scenario patterns over a very wide range of
parameter for birth rates, economic-growth, welfare etc. The authors conclude
that a drastic change in lifestyle would be needed for sustainability. Partly
responding to criticism on the original model, the authors published the results
of an improved global model in 1992 (Meadows et al. 1992), where they argued
that the model supported that a status of sustainability at the time was still
achievable, however that the 'escape'-window had become much smaller, which
implied more drastic adaptations to our global society were needed.
Governments and major oil companies have developed their own scenarios for
world energy developments to help developing business strategies. In the
Netherlands, the energy research centre ECN has developed various scenario
models, the most well-known “Nederland in Drievoud” (Gielen 1999).
5. Sector-wide studies to find general means of improvement of resource and
energy efficiency
In the Netherlands, around the time of the Club of Rome report, a first Cabinet
letter on the environment was issued (VROM 1972). Not much later results of
sector wide policy studies on efficiency improvement were published. Examples
of studies that addressed energy improvement potential are (Boot and van Wees
1982; Melman et al. 1991; Worrell 1994). In all of these studies feedstock supply
and sector structure are taken for granted, and the possibilities of improvements
Process System Innovation Sources 43
2.3.1 OVERVIEW
This section firstly introduces (general) system theory and system engineering.
Subsequently we address and define 'system', 'system elements', system assessment,
design and analysis.
The chemical engineering community has been very successful in adopting system
concepts (Villermaux 1996). Process system engineering has emerged as a sub discipline
where methods for optimal design of chemical plants have been developed, both
with respect to the selection of technology content and of system structure.
According to Prof. Villermaux, it is the mere training of chemical engineering
students in systems thinking that allows them to be productive in so many sectors of
science, engineering and business (Villermaux 1996).
After the original publications of (Boulding 1956) and (von Bertalanffy 1951), system
theory has grown into a vast area with sub disciplines that address hard- vs. soft
systems; static and dynamic systems, system analysis and system design. (Corrigan
and Kaufman 1965: 29) describe the activities of system engineering as “a progressive
definition and integration of functions and design decisions at every level of system
complexity.”
Nelson addresses system engineering in terms of functional activities performed “in
the conduct of a system definition/design (Nelson 1972 216: vi). INCOSE has
defined system engineering as “an engineering discipline whose responsibility is
creating and executing an interdisciplinary process to ensure that the customer's and
the stakeholder's needs are satisfied in a high quality, trustworthy, cost efficient and
schedule compliant manner throughout a system's entire life cycle” (Incose 2004).
According to Asbjørnsen, “systems engineering and systems thinking are formalized
common sense”(Asbjørnsen 1995).
Blanchard & Fabrycky coined the phrase “bringing systems into being” as the
emphasis in the overall activity of systems engineering (Blanchard and Fabrycky
1998: xiii). System engineering thus includes the definition of needs, requirements
analysis, functional analysis, design synthesis and finally system evaluation. The latter
is the objective of systems analysis, which prime domain of use is the improvement
of systems that are already in existence.
(Hartmann and Kaplick 1990) also distinguish between system analysis and system
synthesis steps. According to Sargent, David Rippin would have described the
objective of the systems engineering approach “as the provision of a systematic
framework for tackling the solution of complex problems” (Sargent 1998). “Process
synthesis is the invention of conceptual chemical process designs” (Siirola 1996: 10),
and the question whether “the invention of chemical process designs can be
organised, systematised, or even automated” is at the focus of process synthesis
research. (Siirola 1996: 3).
From these definitions , we conjectured that system theory, and more specifically
(process) systems engineering would provide useful insight and procedures to
address complex industrial systems.
In our case, the system constitutes the whole petrochemical industry. Its present
structure is the result of a long process that by some has been labelled chaotic. In
addition, the utilisation phase of commercial chemical plants often is very long,
“much longer than the life of artefacts15 and artefact-making machinery” (Siirola
15An 'artifact' is “… an ornament, tool, or other object that is made by a human being …” [Sinclair,
1995 #59 :82]. By using this term, Siirola (1996) succinctly captures the differences and analogy
between the production of consumer goods (including intermediates) and capital goods. The output
48 Process System Innovation by Design
1996: 2). Indeed, the economic lifespan of many a chemical facility is extended over
and over again by multiple rejuvenation phases (Grievink 1994). It is to be expected,
therefore, that elements of the present petrochemical industry will be in existence for
a long time to come. Referring to industrial practice, Siirola hints at the relatively
scarcely addressed problem of having to generate flow sheets from scratch, as often
flow sheets from previously realised plants are available (Siirola 1996: 13). Genuinely
novel process system concepts, such as distributed chemical manufacturing (Rowe et
al. 1997) are scarcely reported. It may thus be seen that system engineering per se is out
of the question because it is focused on the technology acquisition phase (Blanchard
and Fabrycky 1998:13). A more realistic objective appears to be to provide a
contribution to system re-engineering or a re-direction of the system’s general
direction of development or transition towards sustainability. The initial activity then
comprises a thorough analysis and assessment of the present system such as
presented in chapter 3, to be followed by a creative but possibly systematised phase
during which novel, innovative system concepts are specified.
of continuous chemical process plants, mass production or the assembly facilities for consumer
products are artefacts, so are the production facilities.
Process System Innovation Sources 49
Using the foundations laid out by these and other authors, Asbjørnsen (1992) coined
a definition of system that appears to avoid ambiguity and to be suitable for the
purpose of devising a modelling strategy for the petrochemical industry. Thus, a
system is defined as “a structured assemblage of elements and subsystems,
which interact through interfaces. The interaction occurs between system
elements and between the system and its environment. The elements and their
interactions constitute a total system, which satisfies functional, operational, and
physical characteristics, as defined by the user and customer needs and requirements,
over a defined total system life cycle of the system existence, including the life cycle
of bringing the system into being” (Asbjørnsen 1992).
Another example is the successful use of the ‘black-box’ approach found in control
system design (e.g. Aström and Wittenmark 1984)), where only the black-box’
response to input changes and/or measured disturbances need to be known over the
range of operating conditions for most common types of controllers. Nothing of the
physical realisation of the black-box needs to be known, its contents are incognito.
The system can equally comprise an area inhabited by rabbits and foxes, a stirred-
tank, or a ship. The dynamic response patterns can be established by observation,
analysis or experiments.
In a ‘grey-box’ approach, knowledge about the contents of the box is used to
conjecture an expected dynamic transfer function, which is validated and calibrated
by experiments. In a ‘white-box’ approach, finally, a model is constructed from first
principles and the available knowledge on the box’s technological contents. It
follows that process flow sheet simulators follow a white-box approach at the level
of unit operations, as the input-outputs relations obey the thermodynamic models
thought applicable. The actual apparatus that would constitute the unit operation,
however, is still considered a black box.
The definition of system elements by Asbjørnsen provides a means to identify
analogies, and explains the impact of information systems engineering on the
development of general systems engineering theory. “Like materials and energy,
information is acquired, transmitted, stored, retrieved, and processed (...) Like
materials, information is processed by separating, combining or otherwise
transforming its parts. (...). Information flow and processing constitute, along with
flow and processing of materials and energy, fundamental aspects of systems
operation” (Blair and Whitston 1971: 28-33).
In our modelling strategy, therefore we have adopted the system concept as
phrased by Asbjørnsen, with the specific notion that the system’s lowest level
is an element that is only defined by its functional characteristics. This then
opens the way to develop new technological concepts to bring about the
transformation between inputs and outputs.
2.4.1 OVERVIEW
Inventions may originate from basic science or R&D in chemistry, chemical
engineering and related disciplines and from studies with a system, policy or
management focus. The latter necessarily employ a meso- or macro-perspective
opposed to the micro-oriented approach in science and engineering. Any invention
becomes an innovation only when realized in an industrial plant or product.
Therefore, process system engineering is crucial for innovation in the petrochemical
industry.
Indeed, reflecting on and summarising the advancements of thirty years of process
synthesis efforts in PSE, Siirola “describes a framework for the industrial chemical
plant innovation process, showing how process synthesis fits into that structure and
how that framework has in turn influenced the development of systematic process
synthesis methods” (Siirola 1996: 2).
As a result of the work of and interaction among academia, the chemical and related
process industry and its knowledge providers, PSE has evolved as a sub discipline of
chemical engineering. Its focus is bringing into being chemical plants that exhibit
excellent operation characteristics and performance with respect to economy,
ecology and safety. The quality of a design, however, is not uniquely quantifiable.
There exists not “the quality“, but many quality factors, some of which can be
exactly quantified, while others lack quantitative precision. The result is a
“fingerprint” of quality factors of a design (Herder 1999).
52 Process System Innovation by Design
The body-of-knowledge of PSE has been developed over some thirty-five years
(Edgar et al. 1999). Today it includes (1) process systems analysis through mass and
energy balancing, second-law analysis and flow sheeting (Denbigh 1956; Rudd et al.
1973; Kotas 1985; Seider et al. 1999 respectively) (2) process system design
supported by elaborate design knowledge formalised in heuristics for conceptual
design (Sherwood 1963; Rudd and Watson 1968; Douglas 1988; Siirola 1996) and by
algorithms for process synthesis that have been developed for specific sub domains
Process System Innovation Sources 53
chemical process system design (Azapagic 1999; Burgess and Brennan 2001; Khan et
al. 2001). Others incorporated engineering thermodynamics into process design and
optimisation and suggested thorough consideration of the available energy resources
and their conversion to useful work (Kalitventzeff et al. 2001; Bakshi 2002). While
the PSE community has realised that “The process industries are facing a range of
growing commercial pressures and challenges” (Bogle and Perris 1999), PSE
research contributions that address sustainability are relatively few in number.
Need
Define Primitive
Problem
Analysis of
Problem
Gather Cognate
Facts
Statement Systematic
Create Specific
of Problem Generation
Problems
of superstructure
Optimisation
Selected Structure and Design
Schemes
Critique / Evaluate
of Alternatives
Embodiment of
Schemes
Detailing
Working
Drawings
etc.
Figure 2-1: Expansion of the General Design Paradigm for process synthesis (after Rudd
and Watson 1968; Siirola 1996; Herder 1999).
56 Process System Innovation by Design
(Kaibel and Schoenmakers 2002: 11) of BASF, however, comment that above
restrictions are disadvantages of both approaches. They recommend that “the
inclusion of a process in the company's process chain and the site conditions must
be specified. Finally, they state that “both methods are used in industry (…),
however, that neither of these methods is used regularly.” Indeed, Sargent reiterates
Rippin's concern “Will computer-aided synthesis of complete plants be used by
industry in due time (…) Are the methods too simple to cope with real industry
problems-or too complex for everyday use by industry engineers?” (Sargent 1998).
Notably Pistikopoulos et al. have addressed chemical process design, integrating four
domains of interests, viz. (a) the process system (b) the control (c) the operations -
static and dynamic optimisation thereof - (d) the production planning and scheduling
in a supply-chain (Bansal et al. 2000; Bansal et al. 2002; Ryu et al. 2004). For any pair
of domains {a,b}, {b,c}, {c,d} and {a,c} convincing studies have been published
concerning the feasibility of such approaches for realistic sections of a process plant.
Dealing with any of these domain-pairs for a full scale plant appears to be beyond
current computational capability. Similarly, whilst the concept is there, actually
dealing with all four domains for a realistic plant section is computationally too
demanding at present.
Kussi et al. of Bayer formulate a work-flow procedure for the effective use of
computer tools in the optimisation of existing chemical plants. They stipulate that
the interlinkage between the different types of modelling approaches and their tools
is a major obstacle to the effective use in industry of computer-aided process design
and optimisation. Whilst the vision has been coined to integrate simulation, design,
control and optimisation capabilities into a single model to support chemical
operations over a plant's life-cycle, they observe that rigorous modelling, black box
modelling and hybrid modelling are still 'living apart together' (Kussi et al. 2002).
Sherwood defined 'engineering design' as “the process of applying the various
techniques and scientific principles for the purpose of defining a device, a process,
or a system in sufficient detail to permit its physical realization” (Sherwood 1963). In
teaching design his approach was to continually remove biases suggested by the
terms used to get students to think outside the box (Westerberg 2004). “A
complete process design includes a critical review of the idea that there should be
any process, the invention or selection of an appropriate process (…).” (Sherwood
1963).
Rudd and Watson also emphasize the synthesis of plausible alternatives and
elaborate on generating ideas prior to engineering solutions. From an ill-defined,
primitive problem they suggest to arrive at a limited set of specific problems that can
be subject to engineering solutions (Rudd and Watson 1968). No use of
mathematical models is made in this phase; only relevant knowledge, information
and data is extracted from those concerned, experts and scientific information. This
approach fits with the General Design Paradigm (French 1999), which was expanded
by Siirola for the conceptual design or synthesis of chemical processes. The relations
are depicted in Figure 2-1. Siirola distinguishes four fundamental phases in a process
synthesis strategy, viz. (1) systematic generation, (2) evolutionary modification 3)
optimisation of structure and design parameters and (4) evaluation of best
alternatives with respect to criteria not in the optimisation problem (Siirola 1996). In
Process System Innovation Sources 57
this strategy, the first two phases have found both heuristic and algorithmic
implementations, the third is the natural domain of mathematical programming,
notably MINLP-techniques and the fourth stipulates to evaluate the mathematical
programming results in a wider context. Indeed, through the execution of every
design process, not only the particular design at hand may be improved, but also the
methods employed in each phase. Difficulties, however, arise when one seeks to
assess the quality of a particular design (Herder and Weijnen 1999).
Chemical processes can be modelled as objects that transform materials from an
initial state to a goal state. Usually, property differences are observed between raw
materials, products and intermediates. According to Siirola, the purpose of “an
industrial chemical process is to apply technologies in such sequence that these
property differences are systematically eliminated and the raw materials thereby
become transformed to the desired product” (Siirola 1996). With regard to process
synthesis steps, Seider et al. appear to define the main objective in chemical process
design as to arrive at elimination of property differences (between input,
intermediate and output materials in the process). The correction of such difference
can be seen to be effected by operators, the essence of means-ends analysis (Simon
1969). According to Seider, such formal, logic based strategies for process synthesis,
means-end analysed have only been developed for simple processes (Seider et al.
1999: 44). Meanwhile, the informal or heuristic approach initially introduced by
Rudd has been applied widely (Rudd et al. 1973). It involves the rules for structuring
process plant operations as indicated in Table 2.1.
Table 2.1: Process synthesis procedure (adapted from Seider et al. 1999: 44);
underlining added; original italics
Synthesis Step Process Operations
1 eliminate difference in molecular types Chemical Reactions
2 distribute the chemicals by matching sources and Mixing
sinks
3 eliminate differences in composition Separation
4 eliminate differences in temperature, pressure and T, P, phase change
phase
5 integrate tasks, i.e. combine operations into unit
processes
These steps appear in recommended order of execution (ibid: 112).
1999: 6). In process invention, the focus is on process creation, i.e. the formulation
and preliminary analysis of a promising base case design. The importance of
simulation tools and heuristics in process analysis and synthesis respectively is
emphasized. Whilst process invention is not defined, the goal of the procedures
outlined appears to be to create an attractive, economically feasible, environmentally
sound and safe process design package. This is in accordance with the
Schumpeterian definition of innovation (Ch. 1). In this view process synthesis
activities lead to innovation when the object of the synthesis activities, the process
plant, is realised. Process synthesis itself, however, is also the subject of innovation:
each novel method of conceptual process creation, if adopted in the community,
represents an innovation.
Beyond single plants, a vast body of process synthesis research concerns the design
and optimisation of energy-exchange or utility networks. The pinch analysis ideas
were pioneered by Hohmann and rediscovered by Linnhoff in 1978, who
subsequently pursued its adoption by industry (Hohmann 1971; Linnhoff 1993). A
massive body-of-knowledge has been developed on the optimum exchange of heat -
using temperature level and pinch based decision methods - and later the exchange
of water, hydrogen and solvents - using purity level and pinch. On an alternate,
applied thermodynamics and optimisation track, exergy analysis and exergy-based
optimisation of energy systems were developed (e.g. Kotas 1985; Yee and
Grossmann 1990; Yee et al. 1990; Cornelissen 1997).
In a recent review paper, however, Li and Kraslawski conclude that the picture of
conceptual process synthesis given in (NRC 1988) is still current (Li and Kraslawski
2004). They argue that process synthesis “at the meso-scale has attained a high
degree of scientific maturation” and that “a strong demand for new, innovative
products and processes has attracted a lot of attention to the phenomena at the
microscopic level”.
Despite the suggestions by Grossmann and Westerberg (Grossmann and
Westerberg 2000), however, process synthesis research at the macro-scale and
conscious analysis or re-design of the industry system structure beyond single plants
has only attracted limited attention.
2.5.1 OVERVIEW
In the overview of system engineering and PSE (section 2.1) definitions have been
given of 'system', 'system element', 'system engineering', and 'system assessment',
'system analysis' and 'system design'. These appear to be sufficient for a modelling
strategy that supports process system innovation of the petrochemical industry.
What is missing, however, is a conscious decision process on selection of the system
boundaries, formulation of the boundary conditions for inputs and outputs, and
system decomposition.
Since the global petrochemical industry has never been designed as a single system,
difficulties can be expected in decomposition. What is a suitable system boundary
and aggregation level for the elements of the system studied? How to break down
60 Process System Innovation by Design
the petrochemical industry system into progressively smaller units, so that the model
meets the objectives stated?
Difficulties in devising a conceptual classification scheme or typology of systems are
encountered in many sciences. In summing up the problems encountered when
addressing world agriculture, for example, Grigg mentions that criteria for
assignment to a certain class must both be developed and tested (Grigg 1974).
Reviewing the system engineering literature and the process system engineering
literature, we have searched for elements and theories that would help develop a
solution for this problem. Texts on system engineering address the topic of system
boundary selection and system decomposition in general terms only, reiterating that
in system design it is not only useful to discern between systems and system
elements, but also to group system elements into subsystems that together constitute
the total system (e.g. Blair and Whitston 1971; Blanchard and Fabrycky 1998). Where
(Corrigan and Kaufman 1965) use the analogy of the magnification setting of a
microscope as a means to increase the level of detail of analysis, Asbjørnsen quotes
the familiar top-down decomposition in trees, which says that each system can be
decomposed into ever smaller level of detail (Asbjørnsen 1992). No indication is
given, however, on how to achieve an optimum magnification-setting, when to
abstained from opening increasingly smaller boxes or when the system content has
become known in sufficient detail. In other words, no stop-criterion is given for the
decomposition-depth.
18 The improvement of ‘the refining of crude oil’, for example, requires improved and new system
elements and networked process system design. Its optimisation, modification or substitution for its
performance in a single product-chain will have an effect on all other product-chains that depend on
this operation.
Process System Innovation Sources 63
and conflict. Thus functional cohesion underpins an abstract, iterative method for
the conceptualisation of a system's functional structure19.
2.6 Conclusions
In academia presently a gap exists between studies on industrial sectors or
economies, and chemical R&D and engineering on products and industrial
processes. Integrated conceptual design of the petrochemical industry's systems
above the level of single plants is a largely unexplored 'white spot'. Conscious
attempts at structured or planned innovation for sustainability at and beyond the
level of single plants appear to be largely absent.
Process system engineering was identified as a resourceful body-of-knowledge to
bridge this gap and address the research question elaborated in this thesis: how can
technological or systemic innovation content be specified that enables a transition
towards a sustainable petrochemical industry? PSE, however, focuses on the
invention and realisation of single plants, and the central question addressed in
process synthesis research is whether “the invention of chemical process designs can
be organised, systematised, or even automated” (Siirola 1996: 3). In process synthesis
research indeed methods are being developed to automate some parts of process
invention. To date, however, the PSE-body-of-knowledge does not offer a
procedure to specify process system innovation content20. Despite the suggestions
by Grossmann and Westerberg (Grossmann and Westerberg 2000), conscious
analysis or re-design of the industry system structure beyond single plants has hardly
attracted attention.
A literature review confirmed that the conceptual design of petrochemical
complexes comprising multiple plants is a largely unexplored 'white spot' in process
synthesis R&D. Thus, it is no surprise that true multi-scale process synthesis
approaches that link the complex, single plant, unit operation and molecular level
appear only to have been suggested (Charpentier 2002). In process synthesis the
computer-based automation of the generation of and selection among design
alternatives receives much attention. It is the work on conscious articulation of
primitive problems into specific design problems (Rudd and Watson 1968; Douglas
1988) (Siirola 1996), however, that offer some building blocks for a methodology to
explore process system innovation.
From a limited literature-search focused on development of system simulation
models and tools for process system design and synthesis, we have only found
superficial coverage of the problem of devising suitable system decompositions. In
none of the literature sources is a stop-criterion given for the opening of the box, i.e.
the decomposition and the 'depth' of the model. Careful consideration of system
decomposition, however, is of utmost importance to avoid unwanted ‘lock-in’
situations that limit the applicability of the model (such as in an economic I/O
analysis and LCA). In PSE, the system decomposition of the industry adopted
19 Functional cohesion is one of the foundations of the method reported in §3.3 Functional
modelling for process system innovation.
20 These observations are largely based on the literature reviewed to unearth previous work on
methodologies useful for the specification of process system innovation content. This work is
presented in detail in Chapter 3, this thesis and has been published in part as (Dijkema et al. 2003).
64 Process System Innovation by Design
appears to be considered trivial and the aggregate level of single process plants taken
for granted. Cohesion methods, a topic in computer science, appear useful for
system decomposition and modelling. In academic R&D the industry system
structure is considered rather fixed and has remained largely unchallenged, except
for the notion that the tasks represented in this structure may be integrated (Douglas
1988; Siirola 1996). As a consequence, the search for improvements implicitly is
confined to the system boundary of the battery limits of individual plants, and the
potential of networked process system innovations has remained largely unexplored.
3 Process System Innovation by Design
Somewhere, (…) between the specific that has no meaning and the general
that has no content there must be, for each purpose and at each level of
abstraction, an optimum degree of generality (Boulding 1956: 302).
For the sanity of the engineer and his chances of success in design, we
would prefer the simplest possible mathematical model (Truxal 1972).
3.1 Introduction
While 'sustainability' has been adopted as a strategic principle for business, it has
been difficult to translate this notion into guidelines or recipes for technology and
business development (e.g. Scholes 1999; Dijkema and Mayer 2001b).
We conjectured that a transition towards a sustainable petrochemical industry21
requires innovation at all system levels, from reaction path to process system, from
chemical plant to petrochemical complex or industry (Ch. 1). Such process system
innovation reaches beyond traditional R&D and beyond the optimisation of proven
technology, proven system concepts and established process networks.
The overview of innovation sources (Ch. 2) revealed that there presently exists a
'gap' between the aggregation level of the economy or its sectors, and the topics
addressed in chemistry, chemical engineering and process system engineering.
According to Brennan “The chemical engineering profession has thus far been slow
to develop a techno-economic taxonomy or classifications system, useful for
understanding the structure and performance of the process industry sectors”
(Brennan 1998: 255).
We conjectured that innovation opportunities must exist that have never been
explored systematically - process system innovations where explicitly the network
structure is considered a degree-of-freedom.
In this chapter, methods are presented to explore this 'innovation-space'. Thus, the
second part of this thesis' central research question is addressed: how can
technological or systemic innovation content be specified that enables a transition
towards a sustainable petrochemical industry?
The methods presented are based on system representation and characterisation
using the input-output paradigm (Appendix A. 4). It is assumed that the function of
a system, (Σ), is to generate a behaviour B to transform a set of inputs (Su) into a set
of outputs (Sy), which relate the system to its surroundings (the external world). The
performance (Ω) of a system is assessed by means of criteria or value functions (P).
In section 3.2 the focus is on system modelling and characterisation for assessment
of resource utilisation, the main indicator for sustainability (see Cadre 3-3: Effective
resource utilisation). Input/output modelling, engineering thermodynamics and the
stream-valuation concept allow formulation of value functions used in a
performance assessment procedure. A priority list of weak performing elements
provides a starting point for the search and specification of promising R&D themes.
Subsequently, in section 3.3, a new method for the specification of innovation
content is reported based on the concepts of functional modelling (Baylin 1990). Re-
addressing system modelling and system decomposition (meta-model, Ch.1, p. 35)
has led to a new modelling-decomposition-synthesis strategy, which serves as a basis
for a procedure for the specification of process system innovation content.
Throughout the chapter the industrial systems around aromatics (benzene, toluene
and xylenes) are used to illustrate and test the methods.
System Decomposition
Together, the process plants and complexes in the petrochemical industry are a
single system (see Ch. 1, Figure 1-5). This can be modelled as a collection of system
elements between which material and energy (exergy) are exchanged. Within each
element a set of inputs (Su) is manipulated and converted to a set of outputs (Sy). In
studies of the petrochemical industry, a single commercial plant within battery limits
appears (see Cadre 3.1) to be the implicit system element of choice (cf. Dijkema and
Reuter 1999). The system image of the petrochemical industry thus comprises a
networked system that is decomposed into nodes and arcs. Single plants are the
transformation nodes and the material flows are the connecting arcs. This is the
model that has been adopted implicitly in chemical engineering and in most of the
work in process system engineering.
Cadre 3-1: Individual (petro) chemical plants
Many chemical plants are part of industrial complexes, which together with 'stand-
alone' individual plants make up the petrochemical industry. The existence of
integrated petrochemical sites may both enhance and hamper the assessment of the
petrochemical industry with respect to resource utilisation. The aggregate data
available on complex performance provides a means to check the data of individual
plants for consistency. As a result of integration the performance of petrochemical
complexes is expected to be better than the simple sum of the individual plants. The
results of assessments based on literature data of individual plants therefore can be
biased. On the other hand, options for performance improvement can be masked by
the very existence of petrochemical complexes, the structure of which might change
considerably in case novel technology for existing plants is offered. The site-wide
integration of technologies, however, in many cases is not so much governed by the
scope of technology, but more by the type organisation, the quality of management,
and by the prevailing economic incentives. Thus one may argue that rather than
assessing the industry as an assemblage of individual plants, the industry assessment
must also address petrochemical complexes, which can be decomposed into a set of
interconnected plants. Thus, a more-or-less intuitive, or logical decomposition of
the petrochemical industry has been formulated as a set of individual petrochemical
plants, some of which are more closely linked by their location in a petrochemical
complex.
Cadre 3-2: Petrochemical complexes
22 In case the system is not in steady state, the effects of temporary accumulation or depletion of
balances to assess the real function of fine chemical plants (Sheldon 1993). In all
these cases, however, implicitly some stream valuation is used.
Generally, two perspectives are used to assess the performance of a particular
production system or part thereof: economy and ecology. In the Brundtlandt definition
of Sustainability, economy, ecology and society are related (World Commission on
Economic Development 1987). In a much used conception of sustainability, 'People,
Planet, Profit', these three dimensions return. We have left out the societal
dimension explicitly in the assessment of production system performance because it
would require a thorough analysis of system surroundings, economic structure,
market dynamics, division-of-wealth (equity) and ecological impact. The economy
and ecology do impact peoples lives, however - if economics are favourable,
production systems will be realised and continue operations, jobs will be created and
consumer needs will be fulfilled. If ecological characteristics are unfavourable,
people will react either because their health or environment is at risk and demand
change.
The set of stream labels presented in Figure 3-1 is an attempt to combine the
valuation of economic and ecological impacts. To allow operation of the system
material and/or energy resources are extracted from natural cycles to provide
feedstock and utilities. Their cost depends on availability or scarcity, market demand,
taxes levied and regulation. The associated ecological impact depends on the origin,
extent and characteristic of such extraction. Ubiquities such as seawater or air are
neither scarce nor does their extraction pose a significant ecological impact. A long-
term risk is involved, however because these may become subject to taxation. The
processing of waste has a positive ecological impact and results in a fee. Using an
economic perspective, in case a market demand exists, inputs are obtained at a cost,
ubiquities are for free, and waste processing brings a fee.
At the output side, main products and co-products are sold at revenue, which must
lead to gross profits, i.e. revenue exceeds average total cost. 'Wanted' by-products
yield at least marginal profits, i.e. revenue exceeds feedstock and operating cost.
'Unwanted' by-products include those sold at revenue less than marginal cost and
those that must be disposed off at a cost. Main products that are in demand bring
such revenue that a profit results under appropriate market conditions. Wanted by-
products add to the margin realised by operating the system. Unwanted but accepted
by-products remain in the cycle but do not add any margin for the operator of the
system. Emissions in many cases cannot be avoided. In case regulatory restrictions
apply, disposal requires treatment at a cost.
The ecological impact of the output categories depends on their use and end-of-life
characteristics. Unaccepted, unwanted by-products represent a loss from the system,
will leave the industrial material cycle, and thus will directly impact the environment.
Finally, emissions can be labelled constrained to limit their adverse ecological impact
at a cost, and unlimited because of their perceived lack-of-impact, which represents a
risk.
The scheme illustrates that between inputs and outputs the economic and ecological
perspectives often do not align: in the manufacture of many products, some adverse
emissions or waste to the environment are accepted.
Input Labels Output Labels
ByProducts
f (o,ch) Cost Incidental
Industrial Wanted + Margin f(u, ch,eol)
System
none Risk Ubiquties Not Desired
In electric power production, for example scarce resources are used to generate a
wanted product, electricity, at the expense of unwanted thermal waste - cooling
water effluent-, CO2 emission and solid waste - flue gas dust and slags-. Slags of
coal-fired power plants are applied as road foundation material, however. Thus these
have been 'upgraded' to a status somewhere in between unwanted waste and wanted
by-product, as there is no revenue or net disposal cost.
The classification in Table 3.1 has been based on the notion that any industrial
process has at least one main product, the reference product. To enable
reproducible calculations all process data are expressed as ratios: the quantity
required for the production of 1 kg reference product. Usually, with the reference
product a primary feed is associated which delivers the bulk of the product. Often, a
process has been specifically designed for co-products and by-products result for
which the system has not been intentionally designed. These may be chemicals that
are produced in small quantities, and that can be sold at a net profit margin, such as
for example the C4 stream in a steam cracker. Over time, stream labels around a
process system may change. By-products, for example, may become key to process
economics and thus be labelled co-products.
Table 3.1: Overview of stream labels around a system or system element
Stream Label Symbol Suffixe │ Symbol Suffixe
Inputs I Outputs O
Scarce resources S Main Product M
Primary f Reference r
Co-feed c Co-product c
Incidental i By Product B
UbiquiTies T Wanted/Desired d
Waste W Not desired
Accepted n
Wasted w
Emission E
Limited l
Allowed a
Grey-box approach -see text
Process-related p Process-related p
Utility-related u Utility-related u
Note: The set of Inputs I, for example has subset Scarce resources S, Ubiquities T
and Waste W. Adopting a grey-box model( Figure 3-2, p.74), these may be process
related or utility related.
product of steam cracking (see Chapter 4) and consists of simple aromatics, non-
aromatics, olefins (C5+) and complex aromatics (C9+). The non-aromatics, olefins and
diolefins present are recovered and recycled to the steam cracker. To that end,
olefins/diolefins are selectively reacted with hydrogen and recovered as a C5-cut by-
product. A mixture of simple aromatics is isolated and a C9+ fraction is recovered as
by-product that has fuel value. Often, aromatics plants are designed for the
production of Benzene and para-Xylene only. Sometimes, styrene recovery is
feasible. This is due to lack-of-demand for the other aromatics present in pyrolysis
gasoline, viz. toluene, m-xylene and o-xylene. Therefore, after separation of benzene
and p-xylene, these simple aromatics are converted to benzene and p-xylene. To
operate, the aromatics plant requires fuel, steam, electricity and cooling agent
(Chauvel and Lefebvre 1989a). In the past decade, in some aromatics facilities
recovery units for styrene have been implemented. From this description, the labels
given in Table 3.2 seem appropriate.
Table 3.2: Stream labels for an industrial aromatics complex
Inputs I Outputs O
Pyrolysis Gasoline S f p Benzene M r p
Hydrogen S c p p-Xylene M c p
Fuel S c u Styrene B d p
Electricity S c u C5-C8-non-aromatics B d p
Cooling agent T c u C9+ B n p
Steam S c u Condens Water B n u
Catalysts S i p Spent Catalyst B w p
In every system, some energy conversion takes place to complete the system's
operation in a finite time-span. In a chemical plant, the chemical transformation of
feedstock to desired chemical product also involves an energy conversion. The
Second Law of Thermodynamics dictates that part of the Gibb's free energy ∆G is
converted to waste-heat. When the conversion of feedstock turns some chemical
energy ∆G into power and heat and the generated amounts exceed the plant's
internal demand it becomes a net exporter of heat or power. An industrial ammonia
plant, for example, has been labelled a power plant with ammonia as a lucky by-
product (Montfoort 1993).
A distinction between 'process' and 'utility' related streams therefore appears useful,
where utility includes non-energy related streams such as lubrication, inert gas,
process water and organic solvents. Thereby a system is decomposed into a process
subsystem and utility subsystem or subsystems.
At the process-side streams are subject to some chemical reaction or some physical
operation. A significant part of the input ends up in the output. Utility streams enter
and leave the process system to supply or withdraw energy from the system or to
provide other utility functions, e.g. to facilitate separation or product purification via
solvent extraction. They are not chemically converted with the process streams.
The utility category consists of streams that enter and leave the process-side
unchanged except for their energy and related physical state (P, T, composition,
74 Process System Innovation by Design
Process
Process input 1...m 1...n Process Outputs
Side
Energy transfer
Utilities
Utilities input 1...p Side 1...q Utilities output
Mass Circulation
Effective resource utilisation expresses the notion that depletable resources that have
been extracted from the earth or from its ecosystems in order to satisfy human need
must be utilised to their full potential, both with respect to quantity and quality (see
Cadre 3-3). The development or selection of criteria suitable for an assessment of
resource utilisation is not trivial. In many (petro-) chemical operations hydrocarbon
material of fossil origin is both feedstock and energy supplier. Resources and
material flows exhibit properties that are conserved in the Universe or part thereof,
such as mass and energy. Entropy, which relates to quality, however, is a non-
conserved property. Thus, while mass and energy cannot be lost in the Universe, it is
implicit that something is 'lost' by ineffective resource utilisation: loss of quality and loss
from the system to the surroundings.
Internal energy, ∆U23, between input and output materials and some exchange of
Work w and Heat q with the system's surroundings (Figure 3-3). Internal energy is a
state property (Appendix A. 2). According to the first law of Thermodynamics,
∆U = q + w (Eq. 3.2)
Enthalpy is a state property defined as
∆H = ∆U + ∆(PV). (Eq. 3.3)
It follows from eq. 3.1 and 3.2 that in case only PV work is allowed on a system (i.e.
no shaft Work, w=0) that ∆H = qp. According to the Second Law of
Thermodynamics the sum of the entropy changes in the system and its surroundings
can only be positive or zero. Entropy, or disorder, is defined as
∆S = q / T. (Eq. 3.4)
Whenever a transformation process proceeds, changes occur both in the entropy of
the materials involved and the entropy of the system surroundings. According to the
Second Law, the created entropy σ is given by
σ = ∆Siirr., created = ∆Ssystem + ∆Ssurrounding >= 0 (Eq. 3.5)
The Gibb's free energy of a system is defined as:
23 In eqn. 3.1 - 3.10, the change in property X (X = Ex, G, H, S or U) between input and outputs is
denoted as ∆X = Xout - Xin
24 Gibb's free energy is a state variable that combines enthalpy and entropy change, accounting for
any change in conditions (P,T,x) associated with the transition; exergy or availability denotes the work
available or required when transforming from the present state to standard conditions.
Process System Innovation by Design 79
or parts thereof can be set up, however, to calculate exergy loss, ∆Exloss. The exergy
loss that is irrevocably related to the system's implementation and operation is
calculated from the exergy change represented by the material transformation,
combined with and the net other exergy inputs or outputs of the system. The exergy
of materials relates to the Gibb’s free energy at standard conditions (Po,To). Since G is a
state property (see Appendix A. 2), whether a chemical conversion takes place in the
system or whatever its exact route is irrelevant for the analysis25.
Process System
Inputs Outputs
G
Figure 3-3: Work and a process system. In case ∆G > 0, Work needs to be delivered to
the process; in case ∆G < 0 the process can deliver Work.
Results from inventorying feedstock, products and utilities for mass and energy
balancing can be used to calculate the exergy value of all streams around a process.
Exergy loss is obtained as
+ (Σwin - Σwout )
According to Eq. 3.7, the theoretical amount of Work associated with a chemical
transformation can be determined from a Gibb's free energy input/output balance
of feedstock and products by assuming 100 % reversible conversion. The delivered
actual amount of work, wact , will be less due to the process implementation selected
25
It is safe to assume for most industrial processes the final state of both inputs and outputs of a
process system are standard conditions P0 and T0 Notable exceptions of course are steam,
cogeneration and heating systems
80 Process System Innovation by Design
For streams around a process system or part thereof (Figure 3-1, p.71) all (sub)sets
of streams as indicated in Table 3.1 (p.72) can be completed by assigning valuation
labels to all streams around a system. External losses per unit main product and
system efficiencies can be defined for any conserved property. Thereby, a set of
performance indicators is defined whereby an impression or “fingerprint” is
obtained of system performance in relation to its surroundings. Each performance
indicator has a unique value function:
Reference Loss - any scarce resource that does not end up in main product:
information often can only be estimated from a system mass and energy balance and
the applicable overall stoechiometric chemical reaction26.
The indicators La…Ld can be calculated for the entire system or for the process or
utility-side respectively. Notably comparison of La…d and La…d, p will yield information
on the amount of utilities required to operate a particular process system, and where
the utility material or energy supplied does end up.
As a matter of course, the stream sets can also be used to define system efficiencies
for materials and energy use, such as:
- the reference efficiency:
26 In the petrochemical industry, oxygen from air for (partial) oxidation processes is the most
commonly used Ubiquity. Waste materials are hardly used. In the assessment of the current system,
therefore, to obtain corrected values for L can be based on an element balance for oxygen. Assuming
that any organic feedstock not accounted for in O is converted to Emission of CO2 and H2O allows
to complete the material balance for each system.
Process System Innovation by Design 83
related, utility-related emissions or that for example unwanted organic products are
downgraded to utilities, where they are eventually converted to CO2 and H2O.
Together, ηa…d and La….d provide a “fingerprint” of the system element with respect
to resource utilisation. For any higher system level (complex, industry), combined
with capacities installed the La….d provides an impression of relative importance of
system element losses.
plants. At both these aggregate levels the Laws of Conservation of Matter and
Energy apply.
The system boundary must be selected to include large-scale petrochemical plants
for undifferentiated, commodity products. Since the focus is on improvement of
current systems and technologies, the product range is fixed. Some of these products,
however, are directly extracted from intermediate oil refinery products. Therefore,
system boundary and feedstock restrictions must accommodate inclusion of, for
example, refinery production of aromatics from reformed naphtha and propylene
from cat-cracked gases.
Often in system assessments, only information is available from open literature.
Such data sets often consist of vendor-supplied data or aggregated input/output data
for a 'typical' plant. The limitation to open literature sources is hard to overcome, as
current performance data on existing petrochemical plants is considered strategic
information, because it is process innovation by which the mature petrochemical
industry competes (Ch.1). This kind of information is assembled, aggregated, made
anonymous and exchanged via consultants for benchmarking purposes and the
industry is extremely cautious to expand the exchange of quantitative information
(Dijkema and Mayer 2001a).
Using open source data a somewhat dated image of the industry's plants
performance is obtained, certainly not an image of current 'Dutch industry', nor the
performance of an individual company. Open source data on petrochemical
complexes, the industry and parts thereof is even more scarce. Construction from
individual plant data is possible, however, assuming that synergies achieved in
industrial chemical complexes are exploited and reflected in the labelling of streams
adopted (§ 3.2.1, p.66). A fairly extensive publication of 1990 name-plate27 capacities
is available (Anonymous 1990), which can be augmented with material from various
other sources. Thus, soundness, speed and simplicity can be matched to the
magnitude of the system addressed, time and data available. Using the data available,
a sufficient 'proxy' to industry performance is obtained for first prioritisation
between system elements (see § 4.3.2, p.123 for an example). To validate the results,
the data used can be compared with company or consultant data and industrial
complex or national economy aggregate data.
Complete data sets on industrial plants are extremely hard to find and there is no
clear standard of representation of performance data. Thus, Input-Output analysis
combined with the valuation of streams as introduced in this chapter is to be
adopted as a standard of representation. Data sets from the literature must be
converted prior to their use in the assessment. Data reconciliation involves a check
for data-consistency and completeness per data set on an individual plant (an
example is given in Appendix A. 6).
To construct a complete the process for useful and sufficiently reliable results the
following may serve as a guideline:
27 The so-called name-plate capacity is the production capacity for which a chemical plant originally is
constructed. Over the years, so-called 'capacity-creep' occurs for a great many plants through
operating experience, plant changes, debottlenecking and/or revamping. Such progress often remains
the proprietary knowledge of plant owners.
Process System Innovation by Design 85
28 Since it was neither the objective nor possible to judge the actual situation in individual production
locations, the data used relate to the situation in or before 1990 are considered to be suitable.
29 Of course, many processes and product flows in the petrochemical industry are at elevated
pressure. The energy-change with respect to the reference state is small, however, even for significant
pressure increase. Care has been taken to use data of the appropriate phase (gas, liquid, solid).
86 Process System Innovation by Design
B
Crude oil
Refinery BTX plant T
St. Cracker
Cokes
Coke oven oX
pX
BTXProd
PS
T ABS
Nylon
etc.
100 Potential 40
External
% Energy Loss PJ/yr
80
30
BTXCon
60
BTXProd
20
40
10
20
0
0
PyGsl BTXExtr. RefGsl CaproLtm Phenol1 PhtcAnh
HDA oXylSep BTXSep Stryren1 Styren2 Phenol2 Cumene
Scope for improvement (100-ηb) %
External energy loss per process
Figure 3-5: Results of energy analysis of the Dutch aromatics system; processes ranked
according to scope for improvement (100-ηb).
The total losses of the analysed aromatics system BTX-Tot are evenly spread over
the two parts. As part of the assessment procedure, however, also the performance
and loss of individual petrochemical plants is determined. Ranking based on the
results yields a number of relatively weak performing chemical processes, notably
Caprolactam, Phenol and Phtalic Anhydride production, as illustrated in the bar
Process System Innovation by Design 89
graph in Figure 3-5. This illustrates that over 90% of the energy loss reported for the
aromatics system can be attributed to the 13 processes in the bar graph. Of these 13,
only 8 exhibit significant scope for improvement. Thus, albeit the Dutch aromatics
system includes some 40 process-plants installed, all else being equal, in order to
improve resource utilisation as a start we only need to examine 20% of these
processes.
20
40 PyGsl 30 20 10 0
High Styren1 Small
Loss Loss
[PJ/Yr] 30 Styren2 [PJ/Yr]
Phenol2
Ineffective HDA
40
Ineffective
but small Loss
PhtcAnh
50 CaproLtm Phenol1
Large Scope [%]
Systematic search
Three main classes of improvement options exist for a petrochemical process that
has been labelled a weak system element:
Process System Innovation by Design 91
3.3.1 INTRODUCTION
In every design of a novel industrial plant for the production of commodities, e.g.
ethyl benzene, the particular selection of unit operations, their implementation and
interconnections are degrees-of-freedom. In actual practice, however, a ‘best’ or
common design appears to have been developed for a great many plants. The
design-space is limited and process flow sheets are hardly ever produced from
scratch (Siirola 1996). Designs of industrial chemical plants thus show considerable
procedural cohesion. As a consequence, by outsiders the development of the
petrochemical industry is perceived to have entered ‘the end-of-the S-curve’, where
there is only scope for 'incremental' improvements when adhering to existing plant
structures. Meanwhile, process engineers are hardly ever confronted with 'ill-defined
statements of a need', primitive problems (Rudd and Watson 1968). Despite the
research agenda's suggested (Ch.2), the focus of process system engineering still is
individual chemical plants. To ‘get out of’ this lock-in and foster process system
innovation, functional modelling is introduced, and related to 'technology-free'
systems representation (section 2.3.3) using the input-output paradigm. Functional
cohesion is used as a basis for system decomposition because modelling using logical
decomposition (§3.2) suffers from difficulties in system boundary selection and lacks
a stop-criterion for decomposition-depth.
Σ = { SU , SY , Ω, B, P }. (Eq. 3.27)
Φ = { SU , SY , Ω }. (Eq. 3.28)
Λ = { SU , SY }. (Eq. 3.29)
30This section is largely based on personal communications (Grievink 1998) and internal research
notes (Grievink 2003) that elaborate on (Dijkema et al. 2003).
Process System Innovation by Design 95
The above description is analysis driven. Given σ and T one can determine Ψ, and
hence B. In a synthesis situation one has to deal with the inverse case. No unique
inverse relationship, however, exists to map from a specified transformation (Ψ’Λ)
onto a particular technology T’:
Usually, one will set targets for the transformation from system inputs to outputs;
then, it will be possible to set targets for the required behaviour (Btarget ) and the
associated transformations: Ψ Λ , target .
31 See §2.5.3p.61 and Appendix A. 3, Overview of cohesion types for definitions and details on
logical, procedural and other types of cohesion.
98 Process System Innovation by Design
This concept of mutual contingency thus offers a more objective criterion for
adherence to the functional cohesion principle. The selection of objective-defined
functions, however, remains subject to interpretation.
When in functional modelling the condition of mutual contingency is met the
decomposition adopted is functionally cohesive. Functional modelling of the system
is successfully completed when functional cohesion has been achieved at each (sub)
system level. If not, the model and the decomposition must be modified. Thus,
mutual contingency provides both the foundation for a definition of functional
cohesion and a stop-criterion for the functional decomposition procedure.
A (pseudo) algorithm has been written for the test for the criterion for mutual
contingency in functional system decomposition (Appendix A. 5).
Process System Innovation by Design 99
Objective-Defined Functions
{S1}
Specific Specific
Objective Objective
A B
Specific
Objective
C
Objective-Defined
Functions {S3}
Objective-Defined Functions
{S2}
Specific
Objective Specific
D Objective
E
Functional Functional
Element Element
Functional Functional
Element Element
styrene respectively (Figure 3-10). In a model of the complex, these technologies can
be grouped to yield a subsystem that is similar to the PO-Styrene plant, a logical
approach for the comparison of system alternatives.
The decomposition of the complete industrial complex by these subsystems,
however, does not meet the criterion of mutual contingency! In the new subsystem,
the objectives of ‘produce PO, ‘produce styrene’ are not mutually contingent: when
the subsystem is part of the larger system, depending on the route of both PO and
EB, either one can be achieved, or both! The possible instances for the processing of
a single molecule of naphtha are listed in Table 3.7.
The solution, of course, is to modify the decomposition, and to separate the PO and
Styrene-process again. This decomposition is mutually contingent.
Isobutane
Naphtha steam cracking
Oxygen per-oxidation
Propylene
separation Oxide
Propylene epoxidation
PO-TBA process
Ethylene Benzene
Oxygen
Ethylbenzene process
Propylene
Propylene epoxidation
Oxide
dehydration Styrene
Legenda
Intermediate
Objective- Objective-
Input output PO-Styrene process
Output
Figure 3-9: Overview of the production system for styrene and propylene oxide.
Process System Innovation by Design
Process System Innovation by Design 103
Labels of instances:
Pl = (ΦX AND ΦY) (parallel)
M = (ΦX XOR ΦY) (mutual exclusive OR);
Mx = only ΦX can be achieved; both (ΦX AND ΦY) and (only ΦY) are not
possible (singularly exclusive towards ΦX - condition M is not met,
only condition Mx is met).
Ax = ΦX implies ΦY (achievement of ΦY through ΦX)
Ay = ΦY implies ΦX (achievement of ΦX through ΦY)
O = ΦX OR ΦY (inclusive OR: ΦX, ΦY or ΦX AND ΦY)
Isobutane
per-oxidation
Oxygen Propylene
Naphtha steam cracking Oxide
epoxidation
Propylene
TBA
separation PO-TBA process
Ethylbenzene
Oxygen per-oxidation
Steam Cracker
Propylene
Propylene epoxidation
Oxide
Ethylene Benzene
dehydration Styrene
Table 3.7: Overview of instances in the extended Styrene /PO production system
Objective output Petro- Ethyl Propylene- Styrene
ΦX / ΦY chemicals benzene oxide
Petrochemicals -
Ethyl benzene A1 (NAF, -
EB)
Propylene-oxide A1 ((NAF, Pl (EB, PO- -
PO-SM) or TBA) or
(NAF,PO- A1((EB, PO-
TBA) or PO- SM2) ((EB,
SM2) PO-SM),
Styrene A1 (NAF, A1(EB, PO- Pl ((PO-SM) -
PO-SM) or SM) or (EB,or (PO-SM2)
(NAF, PO- PO-SM2) or (PO-SM2,
SM2) PO-TBA)
TBA A1 Mx(EB, PO- Pl (PO-TBA) M (PO-SM,
(NAF,PO- TBA, PO- PO-TBA)
TBA) SM2)
Note: for explanation of table and symbols, see notes below Table 3.6, p.103
These result in a reversal of activity flow and introduce feedback loops, which may
introduce complex dynamics that affect system performance. An equivalent reversal
of information flow appears to be largely avoided in business information systems.
Information and control systems do create direct couplings in physical production
and couplings between production control systems, scheduling and management
information systems. At the bottom level control system feedback loops represent
reciprocal couplings that may span multiple unit operations or plant sections. The
feedback from management information systems, scheduling and optimisation also
may introduce reciprocal coupling.
The ‘process recycle’ must be considered an important part of a great many (petro)
chemical plants. Its functions often are multiple: -reprocess unconverted material;
provide initiator/seed for reaction; circulate catalyst; allow safe process conditions;
intensify the use of reactors etc. Indeed, an important step in conceptual design is
the definition of the recycle structure (Douglas 1988). At the level of industrial
clusters, industrial networks, or industrial infrastructure notably materials are
recycled through a series of networked plants, the major problem in the metals
industry being how to keep metals within the cycle, maintain their quality and
availability (Verhoef et al. 2004a).
The recycles create interdependency between the system elements. Using functional
decomposition of process systems, it may be seen that a process recycle that results
in a reciprocal coupling is only a technological implementation of some function in
the black-box of a chemical plant that purposely must convert substances A and B
to product C. In Chapter 5, Fuel cells and Trigeneration, it is demonstrated that
106 Process System Innovation by Design
Identify
1
Intermediate Objectives
Eliminate
2
Unrelated Elements
Separate
3 Modify / Adjust
Sub-groups
Check for
4
Functional Cohesion
Complete
In 1909, Weber published his theory on the location of industrial operations (Weber
1909). He decomposed the processing industry into two subsystems, viz. one where
“Reinrohstoffe” are being processed and another where “Verlustrohstoffe” are
being processed. The petrochemical industry falls completely into the first category,
since crude oil and natural gas must be considered a Weberian Reinrohstoff, i.e. a
resource that can almost completely be converted into useful products at the
expense of only limited generation of solid-waste. On the contrary, most ores are
Weberian Verlustrohstoffe, since for the greater part they consist of rock that is
considered ballast.
Weber's decomposition is still useful for the explanation of the general location
pattern of process’ industry facilities. In retrospect we conjecture that it is a
functionally cohesive decomposition: the process industries’ activities were
decomposed in a subsystem where activities are grouped together that completely
Process System Innovation by Design 109
‘pipeline’, ‘condense’. In this case, or in the case of moving vapour, the unit
operation 'compress' is considered early on in the design process because of the
substantial cost of compressors.
In conceptual plant design, pipelines are not normally considered a unit operation.
'To move substances from A to B', however, is an important functional element in
industrial process plants. This function can be implemented using current state-of-
the art technology and system or by realisation of the concept of a pipeless plant
(Drinkenburg 1999). Douglas emphasized that in conceptual design not only the
selection, conditions and arrangement of unit operations are important, but also the
recycle structure (Douglas 1988; Biegler et al. 1997). We suggested considering
'recycle' in a chemical plant to be a unit operation (Ch. 5).
In the (strategic) planning process for the erection of new chemical plants early
estimation of investment costs is of imminent importance. Due to the large-scale of
operations, it is well known that the actual investment will dominate the net earning
power of the project for a long time. Over time, various methods for the generation
of investment cost estimates have been developed (Chauvel 1981; Peters and
Timmerhaus 1991). In estimating the costs for a new plant, the design of plant must
be broken down into individual units. In their founding article on their method for
the estimation of the investment costs for a chemical plant, Zevnik and Buchanan
actually defined the system element of a chemical process system to be a functional
unit and labelled their method functional unit method (FUM) (Zevnik and Buchanan
1963). At DSM, this cost estimation method has been expanded upon the
recognition that required functions in a process plant can be identified. Each
function can be given an estimated cost prior to knowing or specifying its
technological content. (Drinkenburg 1994; van Geem 1994). Thus, even in the very
early phase of R&D, if some new technology appears to be useful to create a new
production process, a cost estimate can be generated. A drawback of the method is
that in process system design, it is often difficult to clearly define a functional unit,
which very often turns out to consist of more than one unit operation. Possibly, the
procedure outlined in Figure 3-11 offers a workable recipe for the functional unit
method for cost engineering, especially when a basic, expandable set of functional
units that is applicable to all chemical plants can be identified. When this is the case,
a (historic) database of possibly a context dependent cost factor for each functional
unit can be derived.
In simple product assessment, the performance or fit-for-purpose of two alternatives
are compared. In energy analysis, primary, secondary, tertiary and quaternary
resource/energy use may be accounted for (Boustead and Hancock 1979) (§3.2.2)
In Life Cycle Analysis (LCA), such analysis is extended to the complete production
system that is required to bring a particular product into being. Modern LCA or
cradle-to-grave analysis attempts to determine and allocate to a single product or
'functional unit' the effect of mining, operations, production, consumption and
waste processing, recycling and disposal (Setac 1994).
Whilst this decomposition commonly adopted in LCA analysis probably is
functionally cohesive at the level of production systems, the entire approach of many
an LCA-study is not. LCAs are often executed to obtain a fingerprint of a particular
product and its related production chain or network. The result is claimed to offer a
Process System Innovation by Design 111
3.4 Conclusions
In this chapter, the second part of this thesis' central research question was
elaborated: how can technological or systemic innovation content be specified that
enables a transition towards a sustainable petrochemical industry?
Firstly, a method has been developed for system modelling and characterisation for
assessment of resource utilisation, the main indicator for sustainability. Its
foundations are system representation using the input/output paradigm and
straightforward system decomposition where the petrochemical industry is
considered a structured collection of single chemical plants. Mass and energy
balancing, thermodynamics and system engineering were combined in a suitable,
reliable and versatile assessment procedure, which provides a quantitative image of
the petrochemical industry or parts thereof from information that is available in
open literature. The criteria or indicators calculated in the procedure could be
developed employing subjective but robust valuation by labelling system
input/output flows. These labels combine economic and ecological considerations.
Process System Innovation by Design 113
Together with energy values, these provide a sufficient 'proxy' for exergy. In
addition, the stream labelling approach presented provides an interface between all
stakeholders involved.
The assessment results can be condensed into an overview of 'scope for
improvement' versus 'current loss generated'. Obviously, petrochemical plants or
complexes that exhibit a large scope for improvement and a high current loss must
be high on the innovation priority list. Starting from such a priority list, a systematic
search for innovations must include identification of causes within the 'weak
element' combined with inspection of its system surroundings. Since inefficiencies
and losses are manifested as material or energy streams with inferior quality labels;
these labels may be changed to reduce the perceived loss and efficiency of process
systems or petrochemical complex involved. Thus a technology-oriented procedure
has been developed to systematically explore the 'innovation-space' for the
petrochemical industry. This includes the network structure, which is explicitly
considered a degree-of-freedom. We suspected that such a procedure would be most
effective to foster the science- and technology-fuelled innovation process (Figure
1-7), as well as help to balance short-term consumer needs, mid-term business
objectives and long-term sustainability of our industrial society at large. The
illustrations for the industrial aromatics system are no proof, but do demonstrate the
usefulness of the method compiled. Application of the procedure does yield options
for improvement viz. conceptual specification of innovation content. Elsewhere,
more extensive results have been reported that cover the entire petrochemical
industry (Dijkema et al. 1995). Overall, as stated, both the specific results per process
or complex, and more generic results, however, were meagre. This may be due to
very character of the petrochemical industry, a science-based industry with an
impressive track record of technological innovation driven by economic and
environmental incentives. Alternatively, the results can be viewed as confirmation
that reduction of resource consumption and CO2 emission indeed is a wicked
problem for the petrochemical industry (Ch. 1).
Section 3.2 represents the first completion of the circle of the meta-model of
activities (Figure 1-8) where the subsystem decomposition and boundary selection
remained unexplored. In short, a relatively straightforward systemic approach was
compiled for the structured identification of innovation opportunities in the
petrochemical industry, notably R&D options for improved resource utilisation.
In section 3.3, the modelling strategy itself was considered a degree-of-freedom,
notably system characterisation and system decomposition. Functional modelling
(Baylin 1990) was linked and formally embedded into the system representation
using the input/output paradigm. Functional cohesion was adopted as the principle
for decomposition and modelling strategy. A decomposition procedure was
developed using mutual contingency as a stop-criterion.
Our review of process system engineering literature (Ch.2) indicated that until 2002
the process system engineering community has been largely focused on the
development of “new flow sheets for existing problems,” thereby optimising known
transformations in the petrochemical industry while operating in a limited design and
solution space. Functional modelling, however, provides a mechanism for
abstraction that is expected to open the way to redefine functions of the existing
114 Process System Innovation by Design
industry and its production systems, explore novel means of achieving said functions
through system rearrangement and/or novel technology, and finally to explore
creative applications of available technology. In providing a means to decouple from
current practice it resembles and relates to state-task representation: functional
modelling allows the proper formulation of objective-defined functions of existing
operations based on past design. Hidden or new, emergent objective-defined
functions required for sustainable process systems may be identified. Thus, the
functional modelling strategy presented may become an activity that precedes
process synthesis, which mainly assists in primitive problem formulation and
specification of the design-space to be explored in process synthesis. To use the
approach requires a thorough consideration of the functional objectives of the
(production) system studied, and in essence will yield a subjective decomposition
and model.
The general system representation explains the limits of process synthesis for
process system innovation. System boundary selection and functional modelling are
a means to identify and specify novel or alternative system functions and include or
specify novel technology in the innovation space. Thus, re-addressing system
modelling and system decomposition (meta-model, Ch.1) has led to a new
modelling-decomposition-synthesis strategy or a procedure for the specification
process system innovation content.
4 Innovation around Olefins
4.1 Introduction
In this chapter, the focus is on innovation for improved resource utilisation in the
industrial system for the manufacture of olefins32 and their derivatives. The system
around olefins was selected because ethylene and propylene are the key products of
the petrochemical industry because olefins represent the major non-fuel application
of crude oil products and because the mature 'olefins' industry represents a 'perfect'
case to elucidate the specification and the potential of process system innovations.
Olefins are key products of the petrochemical industry. Since these are the building
block of many plastics33 they are of major importance to our industrial economy.
Today polymers34 are used in over 20,000 grades in a myriad of applications, which
include packaging almost everything, synthetic clothing and carpets, toys and
consumer electronics, construction materials and automobile parts. The availability
of olefins is essential for the manufacture of many of these polymers. They are the
single building block in polyethylene (PE), polypropylene (PP) and polyvinyl
chloride (PVC), while polystyrene (PS) is produced from ethylene and benzene. In
addition, olefins are essential components for the manufacture of specialty polymers
such as polyesters (e.g. PET-bottles), polyurethanes (insulation foam, cushions),
polyacetates (paints), polyacrylamides (clothing) and a vast range of intermediate
chemical products. Some 70-80% of world ethylene and propylene production is
used for polymer production (Stanley 2001). The total 79 Million Ton per Year
(MTA) world polyolefin35 production represents approximately 63% of the global
plastics business (CMAI 2001b). The group of olefin chemicals thus includes the
most important building blocks manufactured in the petrochemical industry, viz.
ethylene and propylene. In fact, only sulphuric acid and ammonia are produced in
quantities that exceed ethylene production (Brennan 1998).
Olefins represent the major non-fuel application of crude oil products. Current
nameplate capacity for ethylene is estimated at some 100 MTA worldwide, with
current capacity utilisation around 90% at prices that fluctuate between US$380 and
$550 per ton in the last decade (CMAI 2000a). These units have a capacity for co-
producing some 40 MTA of propylene. Since production quantities are large, the
associated use of feedstock and fossil energy sources is high, roughly 200 MTA.
32 'Olefins' is the common trivial name for the group of mono-unsaturated hydrocarbons with a
carbon skeleton that consist of 2 atoms (ethylene) to 5 (isoprene) or even 8 atoms (1-octene). The
presence of a double C=C bond in these molecules causes them to exhibit significant reactivity,
which makes them a suitable feedstock for a host of chemical processes. The double carbon-carbon
bond is labelled 'unsaturated' as it can be saturated by reaction with a single molecule of hydrogen.
33 'Plastic' is a trivial name for polymers that originates from the label 'thermoplastic resins', polymers
such as polyethylene that slowly weaken and melt at elevated temperature. Thermoset resins do not
exhibit this behaviour. Today, main categories of polymer use are plastics (PE, PP, PS, SAN), resins
(paints, epoxy resins), foam (PS, PUR), rubber (EPDM, SBR) and fibers (Nylon, Kevlar).
34 Polymers is the common name of molecules that are formed from many small monomer
molecules.
35 Polyolefins is the commercial group label for polymers derived from olefins, viz. polyethylenes
Today, these solely comprise non-renewables, notably oil products and natural gas
liquids (NGL). These are estimated to account for some 5-8% of global oil
equivalent consumption (Anonymous 2001a). Thus, only small improvements of the
global system around olefins (production, consumption, end-use) could result in
avoidance of considerable CO2 emissions from this system, which are roughly
estimated to total some 500 MTA36. Since steam crackers for ethylene and propylene
production exhibit an average net energy efficiency of some 70% there appears to be
scope for improvement with respect to resource utilisation.
Olefins represent a 'perfect' case to demonstrate the specification of process system
innovation content because not only is the industry's state-of-the-art rooted in a
large body-of-knowledge that was developed over a number of decades, it is also
largely a mature industry where the steam crackers built or revamped today are built
for decades of operation (Zeppenfeld and Walzl 2001).
Over time the industry has become linked and integrated at both the supply-side and
product side. Polyethylene and propylene plants do appear to be primarily located in
the vicinity of steam crackers, or ethylene / propylene pipelines. Styrene plants are
located near producers of ethylene and benzene, the exceptions obtaining their
benzene from other locations of the same company. The same holds for cumene
production plants. Vinyl chloride presents an interesting case, as both ethylene and
chlorine are costly to transport. Most sites, however, were also found to be located
near ethylene producers. The exceptions were plants that had switched from
acetylene based production, and hooked on an ethylene pipeline. These observations
led Molle and Wever to conclude that “there appears to be a very strong spatial
linkage among functionally related chemical activities” (Molle and Wever 1984).
Determining factors for the location of olefins facilities are the difficult transport of
olefins, the vertical integration in the petrochemical industry, and the pull that an
existing complex exerts on new settlers because of the agglomeration economy, the
availability of existing infrastructure, services etc. The present structure of the
industry and the technology applied are a result of a process that was driven by
economic motives, technological innovation and regulation (Dijkema and Kuipers
2001).
Olefins production by steam cracking is considered “a fairly mature technology”
(Stanley 2001). A characteristic of this maturity is that the potential impact and
associated earning power of R&D expenditure is perceived to be low compared to
other industry sectors. It is well known that such a situation is detrimental for the
rate of process innovation and that gradually the “quest for profitability
improvements” focuses on organisational restructuring to reduce cost (Hutcheson et
al. 1995). Although the industry around olefins has a history of continuous process
technology improvement, real process system innovations have been largely lacking.
This is illustrated by the structure of the industry that is dominated by steam
crackers and where novel processes adopted in commercial practice have been few.
In ethylene production, for example, production capacity has double from 50 MTA
in 1984 to 100 MTA in 2001 (Chauvel and Lefebvre 1989a; CMAI 2000a). Despite
36 Assume (1) that each ton of ethylene represents a total use of two tons of hydrocarbon as
feedstock and utility source for all steam cracker products (2) that, eventually, all hydrocarbon /
polymer derivatives degrade to CO2. Then per ton of hydrocarbon feedstock 2.5 tons of CO2 result.
Innovation around Olefins 117
this dramatic expansion, “the process chemistry and fundamental flow sheet
configuration {of the steam cracker} have remained relatively unchanged” during
the past 25 years (Stanley 2001).
The historic average growth-rate of some 4% per year of ethylene production
capacity is a trend that is expected to continue up to 2004, when world capacity is
expected to increase by 20 MTA to a total of 120 MTA ethylene installed (CMAI
2000a), and total system-related CO2 emission increases by 100 MTA to 600 MTA.
As we have argued elsewhere, however, the petrochemical industry must respond to
the call for reduction of greenhouse gas emissions, and timely anticipate the
depletion of fossil resources (Dijkema et al. 2001). Therefore, the procedures
developed to identify weak system elements and to specify process system
innovation were applied to the industrial system for olefins and their derivatives,
notably ethylene and propylene.
The olefins system in the Netherlands was used to develop the case study where it
was assumed initially that the product spectrum of the petrochemical industry
remains relatively unchanged. First, the entire system was assessed for weak elements
with respect to resource utilisation. The results provide the starting point for a
structured search for possible innovations in the olefins system: options addressed
within the boundary of single plants and changes that imply a change in network
structure or feedstock. The structured search for innovation options for the steam
cracker is reported to illustrate the scope and limitations of the methods reported in
chapter 3. Secondly, a functional model was constructed on the basis of all
information gathered in the assessment, structured search and literature review. This
model was used to categorise the innovations reported in the literature reviewed and
to explore additional process system innovation options to cater for the set of
functions of the olefins system. Thereby, changes within or beyond the boundary of
single plants are considered that rely on a change in network structure or feedstock.
Finally, economic and ecological aspects in evaluating the innovations reported in
the literature and proposed in this study are discussed and conclusions drawn.
37 Aliphatic or saturated hydrocarbons are linear or branched hydrocarbons such as methane, ethane,
propane, octane, iso-octane etc. of the general formula CnH2n+2. These do not contain double or
triple bonds, nor special reactive side-groups attached to the carbon skeleton. The larger the
molecule, the more susceptible it is to cleavage and abstraction of hydrogen (dehydrogenation).
118 Process System Innovation by Design
which may originate from natural gas winning and production or from crude oil
refining.
38 Aromatics are hydrocarbons that contain at least one triple-unsaturated C6 ring (C6H6-n)R1..n. In the
simplest aromatic, benzene, the C6H6-skeleton, is both heavily unsaturated and more stable than the
corresponding aliphatic hexane, C6H12. Aromatics in steam cracker feedstock largely yield undesired
unwanted products because of the relative stability and unfavourable C/H ratio of the aromatic
structures.
Innovation around Olefins 119
2000 1986
Feedstock Europe US World Europe US World
[%] [%] [$] [%] [%] [%]
Ethane 5 55 29 8.0 57.5 30.5
LPG 10 17 11 11.0 19.0 11.0
Naphtha 75 23 54 69.0 9.5 49.0
Gas oil 9 4 6 12.0 14.0 8.5
Others 1 1 1 - - 1.0
1990 1980
Production of [MTA] [MTA]
Propylene 30.320 18.600
% %
Polypropylene 44 30
Acrylonitrile 14 19
Oxo-alcohols 15 19
Propylene oxide 9 9
Cumene 8 10
Oligomers 4 8
Other 6 5
39 Lyondell recently announced construction of pilot facilities, however, to test said direct conversion
the improvement of gasoline quality - notably research octane number, which indicates the gasolines
anti-knock characteristics. TeL has been phased out because of health and environmental risks
associated with distributed Lead emissions from vehicles. In the US, MTBE has become the subject
of a controversy because of leakage from storage tanks and MTBE ending up in groundwater. These
spills, however, appear to have little to do with the actual use of MTBE in gasoline.
122 Process System Innovation by Design
4.2.6 OUTLOOK
Today, the production of olefins, notably ethylene, is considered a mature market
that however exhibits moderate growth rates of some 4 to 5% per year for ethylene
and propylene (CMAI 2001a; Stanley 2001). The sector is in its 'systemic phase' of
innovation (Hutcheson et al. 1995) where the focus is on cost reduction by
reorganisation of the business. Presently, this is manifested in the trend towards
global-sized sites (Dijkema and Kuipers 2001), where larger individual plants for
production and consumption of olefins are combined at lower cost per ton of
product. A second trend is the shift of the ownership of petrochemicals operations
to refiners41.
The development of competing products for the market niches of ethylene-derived
products (e.g. propylene!) has not yet resulted in a decline in demand for ethylene.
On the contrary, with the growth-rates anticipated, by 2018 ethylene production
capacity will have doubled to 200 MTA! Thus ethylene remains a keystone to the
petrochemical industry while at the same time the importance of propylene
continues to increase.
Today, a single type of commercial process dominates ethylene production, the
steam cracker, which also produces the majority of propylene. Their continued
operation worldwide depends on the availability of crude-oil and NGLs. A major
share of propylene is being produced in refineries. The increased world gas reserves
and the launch of Lurgi's MegaMethanol technology has initiated renewed interest in
methanol based routes to olefins, which have lead to successful operation of a
demonstration facility and plans for the erection of at least one commercial plant in
the coming 5 years (Bekkum 2001).
4.2.7 CONCLUSION
On the basis of the olefins system overview and assessment, we conclude that long-
term 'sustainability' is not yet the major concern in this industry today. Where health,
safety and environment issues are high on the industry's agenda to ensure
continuation of its license-to-operate, the still limited efficiency of its major process,
steam cracking, appears to be largely accepted. Its consequence, a significant
contribution to CO2 emission worldwide , is hardly penalized. Major petrochemical
corporations, however, have recognized that CO2 regulation at some point in the
future will be put into effect. Therefore, they experiment with emission trading and
other clean development mechanisms. More important, however, ethylene producers
appear to be trapped in the systemic phase of innovation, where competition
requires continuous cost reduction, which are realised amongst others by increasing
the scale-of-operations. This reduces the industry's flexibility with respect to
developments in feedstock availability, market demand and alternative technology
development.
To sustain current and future olefin production levels at some point in time
alternative feedstock must be sourced and the associated processes must be
developed and employed. Most imminent at the demand side is the changing ratio
41This is illustrated by the recent takeover of DSM Petrochemicals by Sabic and the refocusing of
Shell Chemicals on olefins ethylene and propylene.
Innovation around Olefins 123
4.3.1 INTRODUCTION
The 1990 Olefins system in the Netherlands has been assessed using the method
described in chapter 3. In this analysis, the steam cracker has been identified as the
major weak system element with respect to resource utilisation. Therefore, after an
overview of the olefins system analysis is given, the focus is on innovations in or
around the steam cracker. The steam cracker is described by a summary status of the
technology and technology development. A systematic search for
innovation/improvement options is completed for the steam cracker as per the
approach presented in chapter 3.
4.3.2 ASSESSMENT
42The DSM steam cracker installations in Geleen were sold to Saudi Aramco Basic Industries (Sabic)
as part of DSM's spin-off of its petrochemicals activities in 2002.
124
Vinylidene
PdVC Food-packaging Films
Chloride
Chlorine
EDC VCM PVC Pipes, Film, Wire Insulation
Ethylene
Diamines
Chloro- Tetrafluoro
PTFE Smooth & Medical Products
form Ethylene
Vinyl PE/VAc
Acetic Coatings, Resins
Acetate copolymer
Acid
Terephtalic
p-Xylene
Acid
Ethylene
HDPE Packaging, Toys
Ethyl-
Styrene PS Disposables,Insulation
Benzene Benzene
From C2 to C3 subsystem
From C2 to C3 subsystem
Ethyl- Ethylene
Styrene Ethylene
Benzene Glycol
EPDM
Car door & window seals
PG Ether Rubber
Innovation around Olefins
Styrene
Propylene Propylene
Oxide Glycol
Poly-
Isobutane TBA Foams; Car cushions
Butanediol Urethanes
(BDO)
ABS Car parts; Electronics
Polyether
Polyols
Iso-
SAN Transparent Car parts
cyanates
Poly
Acrylonitril Fiber clothing
Acrylamides
Isopropyl Gasolline
Acetone MIBK MTBE
Alcohol Additive
MMA
Artificial windows -
PMMA
Plexiglas
Interpretation of results
The result of the assessment is not a performance assessment of the Dutch olefins
industry as it existed in 1990 or yesterday, let alone an assessment of a single
Innovation around Olefins 127
Assessment results
The system analysed is decomposed in olefins production and olefins consumption.
Production is the subsystem illustrated in Figure 4-1; consumption predominantly
comprises conversion into polymers, often via a production chain of chemical
derivatives (Figure 4-2 and Figure 4-3), where strictly speaking polymer production is
outside the system boundary of the petrochemical industry. A complete overview of
the system analysed and results is given in Appendix A. 6.
Table 4.4: Key results of the olefin system analysis (1990 data)
Capacity Mass Loss Energy Loss
Lb Le Lb Le
[MTA] [MTA] [MTA] [PJ/Yr] [PJ/Yr]
Production Olefins 3.57 0.70 1.21 47.5 47.5
Consumption Ethylene 2.28 0.44 3.01 16.7 17.0
Propylene 1.67 0.42 4.74 32.8 32.9
Total 3.95 0.86 7.74 49.6 49.9
Polymerisation Ethylene 1.57 0.04 1.71 11.3 12.0
Propylene 0.54 0.04 0.35 4.8 4.8
Total 2.11 0.08 2.06 16.1 16.9
Olefins Total 8.61 1.29 11.01 89.4 90.5
Notes to Table 4.4:
1. Olefins production capacity is only ethylene and propylene; efficiencies and losses are listed
for a steam cracker that co-produces C4/C5 and aromatics.
2. The capacity of the olefins consumption subsystem exceeds that of ethylene and propylene
production because some derivatives produced include oxygen and chlorine.
3. The relatively small capacity of polymerisation indicates other use and export of monomers
from the Netherlands.
43In case one is particularly interested in such assessment, it is suggested to contact one of the many
consultancy firms that specialise in tracking the global (petro)chemical industry's development and
performance.
128 Process System Innovation by Design
The key results of the analysis are summarised in Table 4.4 for process net mass and
energy losses b (material/energy supplied that ends up elsewhere than wanted main
or by-products) and total mass and energy loss e (process energy/material + utilities
supplied that does not end up in products/by-products). Losses b and e have been
calculated as defined in chapter 3 (Eq. 3.21-3.24).
The entire system around olefins requires a total feedstock and energy input that is
equivalent to some 6 [MTA] naphtha, or some 250 PJ/Yr. Its chemical and polymer
represent some 135 PJ/Yr, and some 115 PJ is lost from the system annually. It
follows that the gross energy efficiency (on the basis of LHV) of the system around
Olefins is some 54%. Thus, the energy efficiency of this complex chemical
production system exceeds the efficiency of Dutch electricity generation (which
currently is some 45%).
Weak elements?
The results of the system analysis as presented in Table 4.4 can be interpreted in
various ways. Olefins production and consumption for conversion into
petrochemicals require attention when the results are analysed using the aggregate
system level perspective where the system around olefins is thought to be
decomposed into a subsystem for olefins production, conversion into
petrochemicals and final consumption for polymer production. The results of the
analysis show that both mass and energy loss Lb of olefins consumption for
conversion just exceeds the losses associated with of olefins production. The losses
associated with polymerisation are the smallest. These results should be interpreted
with care, however, as the results on polymer production do not include the losses
associated with PS and other polymers that include material with an olefin origin. In
addition, the Dutch capacity installed for polymer production does not match the
quantity of monomers produced. A considerable amount of petrochemicals is
exported. In case the exported monomers would have been converted in the
Netherlands, the associated losses would total some two-thirds of the losses in
olefins production and associated conversion respectively. The energy incorporated
into the olefins by steam cracking is largely retained when olefins are converted to
more complex monomers, as the energy-rich double bond remains unchanged in the
conversion or is converted into an epoxy structure. Eventually, however, the greater
part of the monomer's increased energy level is used to effect polymerisation, which
results in energy loss from the system as waste-heat.
Inspection of Figure 4-2 and Figure 4-3 reveals that while olefins production (Figure
4-1) only involves the steam cracker and propylene from catalytic cracking, olefin
conversion involves some 20 distinct industrial conversion processes and some 10
polymerisation processes that each contribute to their respective subsystem losses.
Therefore, a proper conclusion is to take special notice of the large loss of the steam
cracker, and to inspect the results for olefin conversion in more detail.
Indeed, when the analysis results are inspected at the aggregate level of individual
plants and when all processes in the olefins systems are ranked, the steam cracker
Innovation around Olefins 129
ranks first with respect to net energy loss and mass loss . In 1990, steam cracker
plants only had an overall energy efficiency of some 65%44.
The results can also be interpreted using an ethylene and propylene subsystem
division. In the subsystem around ethylene the total energy loss, the loss allocated to
steam cracker ethylene production, equals the combined losses associated with
EDC/VCM, ethylene oxide and ethylene glycol. In the propylene consumption
subsystem, especially propylene oxide production is the exception because its net
loss exceeds the loss allocated to steam cracker propylene production.
The energy efficiencies vary from a 39% low (ethylene diamine) to a 93% high
(PE/Vac copolymerisation). Especially endothermic processes (steam cracker),
partial oxidation (ethylene oxide, propylene oxide) and other exothermic processes
(ethylene glycol) have a relatively high scope for improvement index. In all these
processes there is a relatively large energy transformation between heat and chemical
energy where the Second Law of Thermodynamics dictates that retention of energy
quality is impossible and maximum retention of quantity is economically infeasible.
44 It is not the objective of this analysis to assess technology improvement, i.e. the effect of process
innovation during the past decade. Benchmarking has revealed, however, that the Dutch steam
crackers rank amongst the top-10% of steam cracker installations worldwide. Because of the
confidentiality of performance data on individual installed plants, however, it is not possible to verify
and compare current performance data. Suffice to say that if, through process innovation,
performance has indeed risen to some 70-75%, this is a tremendous achievement and proof of the
technological capabilities of the industry and its technology suppliers.
130 Process System Innovation by Design
The most important desired reactions are the primary reactions, where light olefins
are formed. The secondary reactions largely occur following the primary reactions.
They are responsible for the heavier components in the cracked-gas such as benzene
and butadiene. Tertiary reactions include the formation of coke on the furnace pipe
surface. Coke formation occurs via a variety of reactions, by which large carbon
networks are formed. These reactions are favoured by high-pressure. In order to
suppress coke formation, in the steam cracking process the feedstock is diluted with
steam to lower the reactants partial pressure.
The steam cracking reactions occur largely in the gas phase. The reaction mechanism
is via free organic radicals (Chauvel and Lefebvre 1989a). The reactions are initiated
by bond-cleavage that yields two organic free radicals. In propagation, these radicals
can react with the other components to yield desired products + additional free
radicals. Finally, in a gas mixture where the radical concentration increases,
recombination of radicals terminates the chain of steam cracking reactions. An
inevitable consequence of this mechanism is the formation of a substantial amount
of hydrogen and methane in the product mix. All of the individual reaction steps
involved have different thermodynamic and kinetic characteristics, which are a
function of Pressure, Temperature and chemical composition. The product mix of a
steam cracker furnace thus is a function of feedstock composition, furnace radial
tube temperature distribution and longitudinal profile and the residence time. In
SpyroTM all these effects are accounted for to achieve an optimal design.
The steam cracking furnaces today largely have been designed as and operate as
single-pass reactors because the substantial energy requirements to operate the cold
separation dictate that a limited share of recycles must be sent through the process.
Innovation around Olefins 133
part of the feedstock is eventually used as fuel and converted to CO2 and water
instead of desired products. This is caused by a variety of process characteristics: (1)
the reactions towards olefins are strongly endothermic, and are effected at elevated
temperatures in milliseconds. As a consequence, the energy efficiency of the furnaces
is limited to some 85-90% today. In addition, the net transfer of heat liberated from
fuel combustion to the reactor mix is limited, the remainder can only be recovered as
heat. (2) Thermal cracking leads to a mixture of products that inevitably require
separation. In this particular case, some of the products (hydrogen, methane,
ethylene) are separated in commercial installations at high pressure and very low
temperature, in large cryogenic distillation columns. Both the thermodynamic
efficiency of such columns is limited, as well as the efficiency of the cryogenic
operations. Finally, the maximum thermodynamic efficiency achievable in gas
compression is only some 35%. (3) The product spectrum of the cracking section is
such that a relatively large part of the product mix must be transported through a
large part of the separation system, which increases the section's energy-use.
Energy use
Thermodynamically, the 'process path' in a steam cracker deviates considerably from
the optimum, reversible path. One major indication for this is the extreme
differences between the temperature levels of operation. A second indication is the
requirement of rapid heating, followed by rapid cooling. Thermodynamically this
equates to the use of large driving forces, which imply large losses (Prigogine and
Stengers 1984). Apparently, economic trade-offs drives one into this particular
direction. The consequence of this is that a large amount of Work, or exergy, is used
in the hot reaction section, the compressor and the cold separation section
respectively. This is observable in the difference between feedstock and product
exergy. The difference is liberated as heat, only part of which can be utilised.
Improvement, i.e. reduction of the net energy supplied to a steam cracker of existing
design, can be achieved via three major routes:
1. Optimal matching of heat demand and heat generated as a consequence of the
process paths selected.
This is the objective in the use of Pinch technology, to achieve the
thermodynamic optimum in heat integration.
2. Modification of the temperature levels by adding equipment in various heat
streams that convert part of those streams into work: process & energy
integration.
One option is the use of a gas turbine that provides part of the heat of reaction
to the furnaces, whilst producing shaft-work (Dijkema et al. 1998).
3. Modifications in the process system design, e.g. in the cold separation section or
recycle structure.
These modifications can be enabled by the advent of new technology, e.g. in cold
separation technology. Their adoption by industry is subject to the total expected
lifecycle cost trade-off between energy-saved and additional investment. Examples
are adsorption-cooling and open cooling systems that use product streams with
direct contact.
Modifications to the system design also can comprise a combination of 'utility
optimisation' and feedstock/product utilisation. For example, styrene in the
aromatics fraction can be hydrogenated and extracted as ethyl benzene, but it can
also be separated as a by-product. In the latter case, further processing steps from
ethyl benzene are avoided, and utility hydrogen otherwise required can be saved for
other applications.
Innovation around Olefins 137
4.3.4 CONCLUSION
The use of the systematic assessment procedure and systematic search for options
for improvement (Ch. 3) has been illustrated for the case of steam cracker
technology. Since the scope of the analysis and search for improvement was limited
within the steam cracker system boundary, only intra-process improvements were
suggested. It is well known, however, that the upgrading of waste products to by
product (C4/C5 olefins, aromatics) and eventually co product (propylene) has been a
successful strategy of stream cracker operators. The concept, however, can still be
useful today with respect to ethyl benzene and styrene isolation, and the recovery of
methyl acetylenics instead of their saturation. Some of the options thus identified
will require a market to develop and further changes in the petrochemical industry to
materialise.
Function Function
1 S
Mo
s
fin
nom
le
O
ers
The Petrochemical Industry
Olefins BTX
Production Consumption
Function Function
Function M Functio
n1 P
1
Py sol
B, T , m
Ga
ro ine
oX, pX
l ys
is
BTX
X,
Production
Gas ha
LPG
Oil
ht
Function Function
Nap
1 N
ed
rm a
fo hth
Re ap
N
Refining
Function Function
1 R
Figure 4-5: Decomposition around the olefins system (reprinted from Dijkema et al.
2003).
Since the concept of functions allows technology-free specification of systems, in the
process alternative methods of fulfilment - innovations - can be explored. Thus for
example at the level of 'packaging' being the major driver of olefins production and
consumption, one may consider devoting some attention at the 'reinvention' of
packaging and the associated systems to fulfil our (basic) human needs.
Figure 4-2 and Figure 4-3, a slightly different decomposition and boundary selection
was adopted to include the system level beyond the petrochemical industry. Thereby
it may be seen that the 'olefins' or 'polyolefins system' is part of an industrial
infrastructure, which major ODF is to provide packaging material, construction
material and raw material for more durable consumer goods (Figure 4-6). With this
model we can consider the whole chain from crude oil or gas to fulfilling the
function of 'packaging' by a polyolefin. In this system the net energy consumption is
very high. All the energy put into incorporating reactivity into the olefin is lost again
in polymerisation. Subsequently, after end-use, most polymer material is either land
filled or incinerated in the waste management infrastructure. The net result is that
exergy; the potential to do work is lost.
The functional model of our present system given in Figure 4-6 also indicates that
recycling to each other element may be realised possibly via some recycling industry.
Exceptions noted, however, present implementations of this functional element
'recycling' are relatively scarce. The conversion of plastic waste to feedstock, for
example, to date has not been commercially exploited on a large-scale for the
petrochemical industry nor the polymer industry.
A continuous process of scientific and technological development has enabled the
production of thinner and thinner packaging material, thereby consuming fewer
kilograms of polymer per packaging application45. For several decades, however,
packaging and consumer product supply chains have co-evolved, which for example
led to the supermarket of today that sells pre-packed fresh pizza's in competition
with pizza-restaurants. Such development illustrates that the improvements in
material have lead to an increase in the application spectrum. The net result is a
growth in the end-use of plastic, which trickles through to, amongst others, the
production capacity employed of ethylene and propylene: the doubling of ethylene
capacity in the past 16 years cannot be attributed to the population increase and
welfare increase in emerging economies alone. This growth is accompanied by
growth of two 'non-objective outputs' (Ch 4): resource depletion and CO2 emission
(Figure 4-6). This illustrates that sustainability of the petrochemical industry requires
innovation at higher system levels.
In natural ecosystems, the 'decomposers' sustain complex ecosystems such as
rainforests by enabling the high rate of nutrient cycling at a relatively constant (solar)
energy input. Long-term innovation of the petrochemical industry should enable
realisation of a similar industrial society, where essential materials are retained and
recycled at constant net energy input and acceptable CO2 emission.
The ODF of the olefins consumption system is NOT olefins consumption, but this is a
consequence of fulfilling its ODFs.
Satisfaction /
CO2-emission End Use Fulfilment
(non-ODF)
FE 1 FE N
FE 2
Manufacturing Industry
FE 1 FE N
FE 2
ds
in .
ag or..
u er
oo
od m
G
Du cts
g
Waste
Pr nsu
Pa ials
e
bl
Co
er
ra
ck
at
M
Polymer production
Recycling
FE 1 FE N
Industry
FE 2
s
er
m
FE 1
o
on
M
Petrochemical Industry FE 2
Fuels for...
FE 1 FE N
FE 2
he
po
Fee
FE N
ati
we
d s to
r
Fuel
g
tr a
Ge
ns
ck
po
rt
Energy Industry
Waste
Depletion of Resources (Non-ODF)
FE 1 FE N
FE 2 (non_ODF)
CO2-emission
(non-ODO)
At this functional level, thus it is obvious that not only refinery fractions are eligible
feedstock, but also other sources shown, such as synthetic crude-oil fractions from
Fischer-Tropsch synthesis plants that use synthesis gas that either originates from
Innovation around Olefins 143
Coal (Sasol) or natural gas (SMDS). The processing of (mixed) polymer waste in for
example the Veba process would also provide a suitable feedstock because its main
product is a synthetic crude that contains large and aliphatic naphtha and gas oil
fractions (Dijkema and Stougie 1994). Thus, the present fulfilment of packaging
visualised in Figure 4-6 indeed can be realised in a material cycle.
As stated above, the steam cracker can be decomposed into two system elements,
viz. a hot section and a cold separation/purification section.
If we inspect the steam cracker system diagram in Figure 4-4 a bit more closely,
however, we see that a cracker complex consists of a number of distinct functional
elements, viz.
1. Feed selection and feed preparation
2. Fuel selection, preparation and application
3. Various types of cracking furnaces (optimised for various feedstock)
4. A quench operation to provide rapid-cooling of the reaction mixture
5. A stabiliser to separate gases (olefins + saturated hydrocarbons + light gases)
and liquids (aromatics, non-aromatics)
6. A compression section to bring the gases up to suitable pressure for subsequent
cold separation
7. A cold separation section that produces individual olefins and a number of by-
products
8. By product recycle(s), respectively to cracker feedstock and energy supply
9. A separation section that produces individual aromatics and a number of by-
products
Function 1 and 2 are a consequence of the operation and feedstock selected in the
cracking furnaces, while 5 and 6 are a requirement for 7 and result in 8 and 9.
(Figure 4-8).
A logical, high-level decomposition therefore would be one where this system has
four main functional elements, viz. a reactor, separation section, recycle and product
purification . Thus, we are back at the 'standard' model of any chemical plant
(Dijkema et al. 1998) . In order to enable more specific starting points for system
innovation, at the next level of detail, these three functional system elements can be
further decomposed.
From the descriptions above it may be seen that in the reactor section the range of
functions is performed consecutively and/or simultaneously as listed in Table 4.6.
Following the flow of the feedstock / reactants through the installation, functions 1
to 3 take place largely consecutively. On the level of the chemical reactions taking
place in the tubes, both decoking and coking takes place continuously, but as coking
outweighs decoking, a special furnace operation mode of decoking is required.
Obviously, functions (1,4), (2,5) and (3,6) are paired.
144 Process System Innovation by Design
CO2-emission
(non-ODF)
Aromatics
,X
B,T
Recovery Heavies
Conversion
FE 1
(non-ODF)
FE 1
FE N
Product
e
ylen
FE 2 Purification
Eth
Recycle FE 1
e
Propylen
FE N
Feed
FE 2
Nat Selection &
. Ga Preparation Product 1-Butylene
s Liq Separation
uid s
FE 1 FE N
Propan FE 1
e / Buta
ne (LP
G)
FE 2 Fuel
Naphtha Selection & FE N
Preparation
oil FE N
Gas
FE 1
FE 2
il
aso Light gases
av yG
He Natu (non-ODF)
r al G
FE N as
This is largely a mutually contingent set of ODFs in the case of steam cracker cold
separation, possibly with the exception of ODFs (1b,8) because gas-compression
serves both to increase gas pressure, and to let the gas flow through the installation.
From the table it is clear that the operation selected for olefins isolation, cryogenic
separation requires a host of supporting functional elements.
ethylene and propylene'. One of the co-products of these plants, aromatic pyrolysis-
gasoline was soon recognized as providing a source for functions, the production of
essentially pure B, T and X. Another function introduced was the isolation of pure
C4-olefins. Lately, C5-olefin isolation and hydrogen recovery have been introduced,
and processes for the isolation of ethyl benzene and styrene have been developed.
Today, the major ODF of steam crackers appears to be 'produce an economic mix
of ethylene and propylene'.
(1) Categorize/validate. Within the steam cracker flow sheet, major innovations have
been developed and implemented over the years (Hutcheson et al. 1995). In the hot
section, amongst others, these include:
• Improved furnace-tube materials and design, which led to higher olefin
selectivity, yield and reduced coking (fulfilment of functions 1&2)
• Improved furnace design for greater capacity (process intensification) and
greater furnace efficiency
• Milliseconds technology for improved selectivity and yield
All these changes and improvements fall under the umbrella of the functional model
of the cracker and its elements presented, as they deal with improved 'fulfilment' of
said functions. As stated above, these have increased the efficiency of operations
considerably, with the possible exception of the milliseconds technology. It is easily
calculated, however, that today's efficiency of a steam cracker is still only some 65-
75% (based on LHV of feedstock and (co-) products.
Process system innovations have been developed that must be categorised as 'novel
ODFs/functional elements. At the level of the flow sheet, these include the concept
of 'side-cracking' furnaces for dedicated ethane/propane recycle, the isolation of
individual butylenes and the inclusion of metathesis to shift the relative yield of
olefins (Stanley 2001)).
(2) 'Proof' of the methodology. The core of this thesis is a methodology to arrive at
specification of the content of innovations in complex systems. Using functional
modelling and decomposition the ODFs of the steam cracker have been established.
Subsequently, the expert knowledge and results combined in (Dijkema et al. 1995)
and literature cited was combined. The functional model was used to identify
alternative ODFs for the system and to identify alternatives for fulfilment of the
ODFs in the steam cracker model. Together, these have led to the following
suggestion for innovations within the system boundary of the steam cracker complex:
148 Process System Innovation by Design
5.1 Introduction
This thesis includes three chapters on application of the methodology for 'process
system innovation by design' as developed and presented in Ch. 3. In the previous
chapter, a case study on an important sector of the petrochemical industry, olefins,
has been reported (Ch.4). Process system innovation content was specified assuming
a largely unchanged system external world. In Ch. 6 the impact and opportunities are
addressed that are related to a dramatic change in the external world of the
petrochemical industry, the emergence of PEM fuel cell vehicles in automotive
transport.
This chapter focuses on using the methodology to explore and specify innovations
for chemical process systems that make intelligent use of an extra-sector innovation,
fuel cells. These are a novel class of electrochemical energy-conversion technologies,
which are being developed for the energy sector and automotive transport (see also
Ch. 6).
The crux of this chapter is the recognition that it is the combination of fuel cell
functions that must be exploited in process system innovations. The first part of this
chapter consists of an in-depth analysis of the use of fuel cells in the petrochemical
industry at three system aggregate levels - unit operation/reactor, process system
and industrial complex. This work has been presented at the 1999 Joint International
Meeting of the Electrochemical Society46.
In the second part of this chapter, the concept of trigeneration is explored.
Trigeneration is the combined production of chemicals, electric power and heat. Functional
modelling is used to develop innovative system concepts to realise trigeneration. A
paper on this work has been presented at the 6th World Congress on Chemical
Engineering, 200147.
46 Dijkema, G.P.J. (1999a,b), Fuel Cells in the Chemical Industry, Proc. Electrochem. Soc., 196, 1013
(Abstract); Carbonate Fuel Cell Technology V, 306-321 (Paper); Hawaii, The Electrochemical
Society, Inc.
47 Dijkema, G.P.J. (2001) The Development of Trigeneration System Concepts, Proc. 6th World Congress on
5.2.1 ABSTRACT
The use of fuel cells in the chemical industry leads to trigeneration systems that
produce useful chemical products, electricity, and heat simultaneously.
Firstly, stand-alone fuel cell cogeneration systems can be used to upgrade a chemical
plant’s by-product hydrogen or CO (Figure 5-1). Secondly, fuel cell stacks can be
employed as a multifunctional unit operation in chemical process system design,
which results in process-integrated applications (Figure 5-2). Finally, fuel cell
reactors combine the production of chemicals, electric power and heat in a single
electrochemical apparatus (Figure 5-3). In retrospect, the use of fuel cell technology
can thus be reviewed and categorised. In the chemical industry, however, the interest
in fuel cell application appears to be limited to date. In the paper therefore the focus
is on early identification, adequate representation, and early economic evaluation of
the possible applications of fuel cell technology in the chemical industry.
A systematic approach for the identification of fuel cell applications in the chemical
industry comprises three steps: (i) identify all functions of a fuel cell device; (ii)
identify all functions required of auxiliary system equipment; (iii) select suitable
chemical processes that can offer some of the auxiliaries’ functionality (Dijkema et
al. 1998).
A novel schematising technique is introduced that supports this method by
adequately representing the fuel cell system and the functions of its elements. The
method can be combined with elementary mass-balancing and thermodynamic
calculations.
A trigeneration system was developed for a new MCFC with separate CO2 supply
(Hemmes and Dijkema 1998). The result demonstrates that the approach developed
is a powerful tool to screen the potential of a technical innovation.
The early assessment of the scope for fuel cell application in the chemical industry is
completed by the development of an early economic evaluation method. The
additional profits from trigeneration largely determine the allowable investment for
the fuel cell (Vayenas et al. 1991b). In addition, the leverage effect on chemical plant
cost must be accounted for. The results revealed that the economics of fuel cell use
in the chemical industry are favourable compared to stand-alone fuel cell systems for
electric power generation.
In the upgrading of hydrogen by-product the fuel cell only offers a competitive
advantage when the electric power revenue exceeds the value-added of the recovery
and the subsequent conversion of the hydrogen to chemical products.
The use of process-integrated fuel cell modules is attractive because no alternative
technology exists that offers similar functionality. In addition, employing fuel cells
this way improves the performance of the chemical plant, which can provide
leverage to a fuel cell project.
Since fuel cells lack economy-of-scale, it was expected that fuel cell reactors could be
competitive only in suitable small-scale applications. Based on calculations on
potential electric power generation from a number of industrially important
Fuel Cells and Trigeneration 155
reactions, however, it was concluded that only relatively small-scale fuel cell units (6-
40MWe) are required for common large chemical plants. In addition, fuel cell
reactors appear economically competitive for most partial oxidation reactions.
Notably the partial oxidation of toluene to phenol stands out as a prime candidate
for further exploratory research.
Chemical plant
Feed Reactor Separator Product
Byproduct H2 / CO
Fuel cell
Electricity
system
Heat
Figure 5-1: Trigeneration through by-product utilisation with a cogeneration fuel cell
system.
Chemical plant
Feed Reactor Separator Product
Fuel cell
Electricity
stack
Heat
Chemical plant
Heat
5.2.2 INTRODUCTION
Traditionally, fuel cells have been promoted because of their high thermodynamic
efficiency for the generation of electric power. Their second advantage proclaimed is
their cleanliness (Hirschofer et al. 1994).
As of late, it has been realised, however, that the best fuel cell system not necessarily
is the thermodynamically most efficient system. Rather fuel cell systems must
uniquely meet customer demands at acceptable cost.
In the chemical industry ample opportunities exist for such fuel cell applications.
Not only because equipment can be shared, but also because the fuel cell application
can have an effect on the performance of the entire plant and thereby provides
economic leverage to a fuel cell project (Dijkema et al. 1994a; Schinkel et al. 1994).
The overall impression of the literature on fuel cell application in the chemical
industry is one of incidental, fragmented publishing of possible uses. The
trigeneration system concept therefore is used in this chapter to present all
prospective fuel cell applications in the chemical industry under a single systematic
framework.
Firstly fuel cell system design is addressed and the “functional scheme technique” is
presented. Subsequently, an overview of the literature is given per category of
prospective fuel cell use in the chemical industry (see Figure 5-1, Figure 5-2and
Figure 5-3). Thirdly, the functional scheme technique is used to augment the
identification of fuel cell applications (Dijkema et al. 1998). The merits of the
combination are illustrated by the development of a trigeneration system for a the
newly invented improved MCFC with separated CO2 supply (Dijkema 1998;
Hemmes et al. 1998). Finally, a method for early economic evaluation of fuel cell
applications in the chemical industry is presented and results are reported.
Air supply
Air Exhaust
Energy
Air Energy transformation
Exhaust
supply transformation Air
Step IIa
E
Fuel T Power
E Power
Abstraction
Fuel
E Power
T Air
Power
Fuel
Current Realisation Alternative Realisation II
case of the fuel cell, the flue gas will contain unconverted feedstock. In all cases we
have assumed the feed is at the right condition (P, T).
We suggest, therefore, that at each level of system complexity, the following
procedure can be applied:
1. Review an existing (similar) system design and identify the crucial system
functions;
2. Draw a functional scheme
3. Briefly describe each system element
4. Explore alternative technical realization options per system elements
5. Explore possible combinations of functions in a single apparatus
6. Consider alternative recycle schemes to achieve 5&6; the flow sheet recycle
structure needs careful attention (Douglas 1988).
In Figure 5-5 an early fuel cell system design is represented that operates on natural
gas. In this drawing, the system decomposition has been visualised. The system
elements have been enclosed by grey boxes. Each box’ title represents an element’s
function. The representation essentially is technology-free: the drawing is still valid
when the detail in the boxes is left out. The content depicted represent only one of
the possible technical realisations of the system elements. The particular
arrangement of system elements also represents a single possibility to ‘materialize’
the fuel cell system. Indeed, the second important feature of the drawing is the
system boundary combined with its title, which is a system element at one
aggregation level up. Note that all feed and product streams have been labelled
outside the system boundary.
Feed Exhaust
saturation Off-gas
Fuel cell module
processor
Water reformer
Anode
Power
burner Cathode
Fuel
Fuel processor
Compression
Air
Cathode
cooling
Cold Hot
medium medium
Figure 5-5: Example: functional scheme for a stand-alone fuel cell system.
A third important feature of this drawing technique is that system elements where
multiple, physically unseparated functions are performed have been enclosed by an
equal number of system element boundaries (effectuation of step 6). In this way
Fuel Cells and Trigeneration 159
alternative arrangements for the design of the system offered by the integration of
functions are visualised.
Besides the fuel cell module, the stand-alone system includes 4 other elements,
steam supply, fuel processor, oxidant supply, cell cathode cooling, and off-gas
processor.
Steam supply is a function implied by the particular selection of fuel processing in
this system, steam-reforming. Prior to reforming, methane needs to be mixed with
steam. The option shown is to integrate the steam supply in the fuel processor by
installing a feed-saturator that employs waste-heat.
Fuel processing consists of hydrogen generation and fuel purification. Since no
natural resources of hydrogen are known, it must be generated from hydrogen-
containing species.
Fuel
Anode
Cathode Hot exhaust
Feed preheat
Feed compression
Fuel processor
burner Power
Air
Off-gas processor
Air supply O2 supply
CO2 supply
Cathode preheat
required for the steam-reforming reaction. In an external reforming fuel cell system
the off gas can be recycled to the reformer inlet.
The other essential feature of the Smarter system concept is the integration of the
off-gas processor and cathode oxidant and CO2 supply. The outlet of the off-gas
processor, a catalytic burner, is sent to the cathode of the direct internal reforming
MCFC together with excess air for quenching.
In retrospect, the system design procedure developed clearly illustrates the
advantages of the Smarter System compared to a common stand-alone fuel cell
system. The integration of multiple functions in a single apparatus and the
usefulness of recycles are apparent. The Smarter system is a dramatic breakthrough
in system simplification, and is a case in point for the use of the system design
approach presented.
Electrochemical reactors have long been applied on an industrial scale, notably for
the production of aluminium, zinc, steel, chlorine and adipic acid. According to
Yentekakis et al., however, the feasibility of electrochemical fuel cell reactors was
demonstrated for the first time in 1980, notably for the conversion of ammonia into
nitric oxide, one of the prime steps of nitric acid in fertiliser production, in a SOFC
(Vayenas et al. 1992). Other industrially important reactions include the oxidative
dehydrogenation of ethyl benzene to styrene, and of 1-butene to butadiene
respectively, the ammoxidation of methane to hydro cyanic acid, the partial
oxidation of methanol to formaldehyde, and the oxidation of H2S to SO2(Yentekakis
and Vayenas 1989). (Malhotra and Dattra 1996) mention tentative development of a
novel fuel cell for the trigeneration of electricity and acetaldehyde from ethanol.
The applications of the electrochemical reactors described above comprise
trigeneration of electricity and a useful chemical by controlled partial oxidation.
Synthesis gas-proper is another candidate product that can be produced by partial
oxidation. (Alqahtany et al. 1993) investigated the production of synthesis gas in a
solid electrolyte cell that incorporates electrochemical oxygen pumping (EOP). They
conclude that EOP allows control of CO-formation. Since H2 formation remains
largely unaffected, the H2 /CO ratio can be controlled by EOP. Finally, they suggest
to use their results for the development of an in-situ methane reforming SOFC,
which the authors claim to perform equally or better than existing industrial
methods under the co-production of electrical power.
Mazanec et al. were issued a patent on the use of certain materials in membrane
reactor or fuel cell arrangements for oxidation reactors. They list examples of
methane to synthesis gas or unsaturated compounds, the partial oxidation of ethane,
substitution of aromatic compounds, and, notably the extraction of oxygen from
oxygen-containing gases which they deem useful for exhaust gas cleanup (Mazanec
1993).
The iMCFC
During the initial testing of the iMCFC concept, the question emerged “Do market-
niches exist where only the iMCFC would fit, and what is the scope of the iMCFC
with respect to system improvement.”
recycle
CO / H2
Anode
CO2 / H2O / inert
Matrix
<CO, H2>
Cathode
N2 / O2 N2 / O2
recycle
Not only the cell’s performance improves, however, but also a 3 input - 3 output
fuel cell results. These unique features must be exploited in the design of a system to
fill some market-niche unique to the iMCFC. To be competitive, any system design
for the iMCFC must consist of a minimum amount of equipment, notably gas-gas
heat exchangers.
Off-gas cooling
Feed preparation Syngas production Fuel Cell Module Syngas
production
Feed
Fuel Anode
saturation
Gas-heated
reformer Matrix POX
Water
Cathode
Oxydant
preparation
Air Oxygen
plant
Syngas
favoured by low operating pressure and high operating temperatures. Increasing the
pressure from 1 to 10 bar dramatically reduces the conversion of methane at
otherwise equal conditions. At 10 bar and iMCFC inlet temperature of 650 °C and a
Steam-to-carbon (SC) ratio of 2 the methane conversion is as low as 35%. As a
consequence, however, the heat load of the GHR is also reduced! The remainder of
the methane is converted in the partial oxidation reactor (POX), or autothermal
secondary reformer: oxidation of part of the gas drives the reforming of the
methane present. Thus, by sandwiching the iMCFC between GHR and POX heat-
load is shifted away from the GHR. Because the GHR involves physical heat-
transfer, and the POX ‘chemical’ energy transfer, this shift increases system
efficiency. The net result is that with the same amount of feedstock, synthesis gas
can be made and electricity produced. The iMCFC operation at an elevated pressure
is advantageous for the sandwich system design. An elevated pressure previously
was limited by cathode dissolution, but is no issue in the iMCFC because of the
decoupling of CO2 and oxygen supply to the cell. pCO2 can be kept at a very low
value (10-3 atm.) to reduce NiO dissolution.
Conclusion
The early identification method combined with the functional scheme technique has
proven to be useful to exploit fuel cell innovations by the early development of a
trigeneration system concept.
Finally, it may be seen that in all the economic methods for early project evaluation
the problem is reduced to a closed problem by sufficiently defining the system. In
the early economic evaluation of fuel cell applications in the chemical industry in
general we are after some general insights into the attractiveness of fuel cell
technology in the chemical industry, an open problem where we attempt to explore
an endless solution-space to help define promising directions for R&D!
In words: the total additional revenue from trigeneration throughout the technical
lifespan of the fuel cell must exceed the difference in initial investment.
required, in some cases not even a reformer. This type of fuel cell application can
provide an additional competitive edge over stand-alone fuel cell systems by
improving chemical plant efficiency and/or increasing plant capacity. In other
words, the fuel cell module incorporation results in a leverage effect on plant
performance. In the methanol plant, for example, the energy consumption of the
syngas compressors is significantly reduced, and plant capacity is increased by some
8%.
As we have shown in earlier work (Dijkema et al. 1994b; Dijkema et al. 1998), the
main advantage of a process-integrated fuel cell originates from the leverage effect
on chemical plant performance.
When we consider the function of hydrogen in chemical plants there are four
possibilities:
A) Hydrogen is a feedstock
B) Hydrogen is part of an (intermediate) feedstock in the plant, for example in
ammonia or methanol production via synthesis gas.
C) Hydrogen is a by-product of the chemical reactions only, for example in
dehydrogenation processes or dealkylation of toluene to benzene.
D) Hydrogen is a process agent (catalyst, diluent, promotor, reductor etc.)
But do all these options require their own modified equation (6.2)? In case A and D
it may be argued that there is no case for trigeneration using process-integrated fuel
cells: hydrogen is required for the chemical reactions themselves, any consumption
by the fuel cell would only add to the consumption of the plant. Note that in case D
usually some waste-hydrogen stream results, which then can be used in a fuel cell
system as has been analysed in the previous section. The same applies to case C.
In case B, however, an important incentive for trigeneration is the sharing of
facilities, which can offer advantages of economy-of-scale. In commercial plants,
however, hydrogen-production facilities often already are scaled to the maximum
technically feasible size. Only if there is a mismatch between maximum-size of
hydrogen generation and hydrogen consumption part of a chemical plant is there
scope for trigeneration, e.g. in ammonia production (Dijkema et al. 1996).
What remains is the incentive created by a leverage effect on plant performance, as
illustrated by the integration of a fuel cell stack in a methanol plant. In this case, (Eq.
5.2) can be extended with a term that quantifies this incentive, Rchem, trigen – Rchem, plant
The use of fuel cells to boost the performance of the plant at present appear the
most attractive market niche for fuel cells in the chemical industry. The exact
incentive, however, can only be estimated when performance data are available from
a preliminary process system design.
Because the investment cost normally quoted for fuel cells is in [$/kW], whilst that
for chemical reactors is quoted in [$/(ton/day)], we introduce α to relate the two:
Example evaluation
Consider, for example, the direct electrochemical partial oxidation of propylene to
propylene oxide, PO (Table 5.1, reaction 12). From the thermodynamic data in (Stull
et al. 1969), the Gibb’s free energy change of the reaction at standard conditions can
be calculated, -21.2 kCal/mole. Using equation (5.5) results in α = 8.8
kW/(ton/day).
The allowable investment at a technical lifespan of 5 years (5*365 days) and at an
electricity revenue of 5 ¢/kWh follows from equation (5.14). The allowable
investment is 10 k$ per (ton/day) of propylene oxide production capacity. By
division with α, we obtain the familiar result of 1095 [$/kW] allowed investment for
the fuel cell reactor.
Table 5.1: Fuel cell reactor evaluation for the chemical industry
Chemical Operation Fuel Cell Reactor Trigeneration system data
Max. Typical Fuel Cell Fuel Cell/ Typical ∆I as % of
Allowed plant Power Conventional Plant Inv. Plant Inv. I
Reactor Max ∆I
∆Gr M α If Co Rating
Reaction Kcal g/mole k$/ ton/day MW Million $ Million $ %
/mole (ton/day)
1 Hydrogen oxidation H2 ½ O2 → H2O -54.6 18 73.4 80 n.a. n.a. 0.0 0 0
2 Carbon partial oxidation C + ½ O2 → CO -32.8 28 28.3 31 n.a. n.a. 0.0 0 0
3 Carbon oxidation C + O2 → CO2 -94.3 44 51.8 57 n.a. n.a. 0.0 0 0
4 Syngas by partial oxidation CH4 + ½ O2 →CO + 2 H2 -20.7 30 16.7 18 2000 33.3 36.5 80 46
5 Syngas by reforming CH4 + H2O → CO + 3 H2 34.0 70 -11.7 -13 2000 -23.5 -25.7 0 0
6 Methanol by direct partial CH4 + ½ O2 → CH3OH -26.7 32 20.2 22 1800 36.3 39.7 130 31
oxidation
7 Formaldehyde CH4 + O2 → CH2O + H2O -14.1 30 11.4 12 285 3.2 3.6 6 59
8 Acetaldehyde C2H5OH + O2 → C2H4O H2O -46.3 44 25.4 28 285 7.2 7.9 19 42
9 Acetic acid via CH3OH + CO → CH3COOH -18.4 44 10.1 11 230 2.3 2.5 28 9
carbonylation
10 Ethyleneoxide C2H4 + ½ O2 → C2H4O -19.4 44 10.7 12 400 4.3 4.7 58 8
11 Ethyleneglycol C2H4O + H2O → C2H5(OH)2 -34.4 62 13.4 15 285 3.8 4.2 12 35
12 Propyleneoxide C3H6 + ½ O2 → C3H6O -21.2 58 8.8 10 285 2.5 2.8 102 3
13 Butadiene C4H8 + ½ O2 → C4H6 + H2O -34.4 54 15.4 17 140 2.2 2.4 26 9
14 Phenol by toluene C6H5(CH3) + 2O2 → C6H5OH + CO2 + -131.3 78 40.7 45 140 5.7 6.2 39 16
oxidation H2O
15 Styrene by C6H5(C2H5) + ½O2 → C6H5(C2H3) + -34.8 104 8.1 9 1000 8.1 8.8 52 17
oxydehydrogenation H2O
16 TPA C6H4(CH3)2 + 3O2 → C6H4(COOH)2 + -279.6 166 40.7 45 285 11.6 12.7 85 15
2H2O
17 Nitrous oxide NH3 + 1 ¼ O2 → NO + 1½ H2O 20.7 30 0.0 0 0 0.0 0.0 0 0
Fuel Cells and Trigeneration 173
A common capacity for a commercial industrial facility is some 285 ton per day PO,
or 100.000 ton per annum (Chauvel and Lefebvre 1989b), which implies that the
total allowable investment for the fuel cell reactor amounts to 2.85 Million dollars
(assumptions: investment conventional PO reactor zero; electricity revenue 5
¢/kWh, minimum lifespan fuel cell 5 years).
Any company involved in PO-production can easily insert the correct figures for
refinement of the result, for example:
◊ Actual feasible single train conventional reactor capacity must be used
◊ A portion of conventional reactor investment can be added
◊ Fuel cell lifespan can be corrected to compare with conventional catalyst
replacement cost and frequency
Evaluation of results
The calculation was completed for a number of important partial oxidation reactions
as they are carried out in the petrochemical industry today (Chauvel and Lefebvre
1989b). In addition, results are listed for candidate reactions to replace presently
employed reactions, for example methanol by direct partial oxidation.
A high value of α indicates that along with chemical production a high amount of
electric power can be potentially extracted. The combination with economic figures
allows additional assessment of the fuel cell reactor’s attractiveness: the higher the
maximum allowable investment, the higher the possible benefits. In the table, not
only reactions are included that result in some useful product, but also the
conventional fuel cell reaction, the oxidation of hydrogen appear favourable, as
shown by their high values of α. The following observations can be made by
inspection of these results:
1. The conversion of light molecules leads to high values of α.
2. CO2 formation in partial oxidation yields high α’s
3. The power rating of a trigeneration fuel cell reactor calculated, 6-40MW is
reasonable for world scale, chemical plants.
In addition, the allowable investment calculated for the fuel cell reactors offers
substantial scope for development of the technique for a number of reactions,
notably the partial oxidation of toluene to phenol. The electrochemical trigeneration
of formaldehyde by partial oxidation of methane appears attractive because of the
allowable investment represents a high portion of the plant’s total investment.
Since partial oxidation reactions involve the use of oxygen, the production of pure
oxygen is avoided in the case of employing a fuel cell reactor that uses air. This must
be taken into account, e.g. by adding an additional term in equation 5.2, (Rc,O2 –
Rr,O2), where R is negative, i.e. a cost.
The results also support the current focus on cogeneration fuel cell reactors, as the
figures for hydrogen and CO oxidation are most favourable. These are precisely the
overall reactions carried out in currently developed fuel cells. One could think of
selling pure water or CO2 as a trigeneration product of the fuel cell reactor. Note
174 Process System Innovation by Design
that the equation developed remains valid because of the assumed zero-cost of the
chemical reactor.
Please note that in the calculation of all these results, it was assumed that the cost of
a conventional reactor equals zero, i.e. economy-of-scale does not play a role!
5.2.7 CONCLUSIONS
The trigeneration concept is a suitable basis for a systematic framework to evaluate
fuel cell application in the chemical industry. The development of a trigeneration
system for a novel fuel cell type, the iMCFC has demonstrated the value of system
design in the earliest phase of R&D.
The functional scheme technique developed has proven to be valuable in the
identification and development of novel fuel cell systems. It is expected that it will
also be useful in the development of process systems, and that it may serve as a
communication framework for system development and system integration.
We expected that only qualitative early economic assessment of fuel cell use in the
chemical industry would be possible, and that the analysis would confirm the meagre
prospects of fuel cell use in the chemical industry. On the contrary, however, a
simple criterion could be derived for the evaluation of fuel cell reactors as the heart
of a chemical plant. The results show that the prospects for fuel cell reactors as the
heart of trigeneration systems are worthwhile exploring. Trigeneration yields
additional profits even when substantial investment is required, the allowable
investment in fuel cell reactors calculated is in the range of 15-60% of total plant
investment.
The use of by-product hydrogen has been shown to be only a last resort, when all
other options are exhausted. Finally, process-integrated use of fuel cells appears to
be most attractive when some leverage effect on plant performance or cost can be
realized.
Fuel Cells and Trigeneration 175
5.3.1 ABSTRACT
Trigeneration is the simultaneous production of chemicals, heat, and power. A
procedure is presented and illustrated for the development of trigeneration system
concepts from the industry level down to the level of individual plants and parts
thereof. It is shown for trigenerate petrochemical complexes that the selection of
system elements for minimal fuel consumption depends on the amount of
trigenerate electric power that can be exported. The avoidance of some 190 kg CO2
per MWh can be achieved for a typical complex. This net reduction in the total CO2
emission related to chemicals, heat and power production can be achieved because
the chemical industry acts as a heat sink. Finally, the realisation of trigeneration in
the chemical industry involves a shift in perception of the industry: rather than
considering it a threat because of pollution, safety and health risks, it must be
considered an enabler of a reduction of total human CO2 emission.
5.3.2 INTRODUCTION
Trigeneration systems are facilities for the simultaneous production of useful
(chemical) products, electricity, and heat. One may argue that today such facilities do
not exist, but actually, they do! Today, for example, national economies or even the
EU are systems where a net amount of chemicals, electricity and heat are produced.
At numerous petrochemical sites the initial application of cogeneration systems has
led to the development of the on-site utility island concept.
Current petrochemical systems, however, have not been 'designed' for trigeneration.
In addition, these systems do not operate somewhere close to the maximum
thermodynamic efficiency achievable. In accordance with the Second Law of
Thermodynamics, in each process the total entropy of a system and its environment
increases -some minimum amount of Exergy is lost- and a net amount of feedstock,
fuel or heat is expended to provide a process' driving-force. Since the economic
operation of a great many petrochemical processes requires operation at elevated
temperature and pressure, the total exergy loss increases. Thus, whilst a large part of
the chemical reactions completed in industry is exothermic and the net change in
free energy could allow production of Work, a net input of heat and Work is
required. Trigeneration offers a means to minimise the combined inputs for
electricity, chemicals and heat production, albeit at the expense of additional
investment.
Today, the threat of global warming has led many people to believe that some
reduction of human CO2 emissions into the global atmosphere is urgently required.
Therefore we explored whether the adoption of trigeneration in the petrochemical
industry can help reduce total human CO2 emissions by improving total fossil
resource utilisation.
Outline of section
In the past edition of the World congress, we presented a paper on the integration of
energy-conversion devices in chemical plants (Dijkema 1996). This was largely
'technology-oriented' and focused on single chemical plants. Since then we have
been elaborating the concept of 'trigeneration' and developed functional modelling
(Dijkema and Reuter 1999; Dijkema and Grievink 2000) to help identify system
innovation options. This method was successfully applied to develop and evaluate
trigeneration system concepts for a range of chemical processes and a number of
energy technologies, notably fuel cells (Hemmes and Dijkema 1998; Dijkema 1999),
gas turbines and cogeneration systems.
In this chapter, we will address the development of trigeneration system concepts
for the level of industrial complex to individual plant level. Firstly, some background,
the problem addressed and the definition of trigeneration will be elaborated, and the
research approach summarised. Subsequently, functional modelling will be
introduced to help understand and develop trigeneration concepts. Third,
trigeneration as 'leitmotiv' for an industrial complex or area is investigated from an
ecological and economic perspective, and the potential of CO2 emission reduction
will be analysed. The prospects of trigeneration in the European Electricity markets
are briefly discussed. Finally, some conclusions are presented.
Definition
We have defined trigeneration systems as facilities for the simultaneous production
of useful (chemical) products, electricity and heat (Dijkema 1996). Notably in utility
engineering, sometimes the term trigeneration is used for the simultaneous
generation of electric power, heat and cold. In chemical engineering and R&D, the
term chemical cogeneration has been used to describe systems that produce electric
power and chemical products, notably fuel cell reactors (Vayenas 1988).
Fuel Cells and Trigeneration 177
In our view, however, trigeneration best coins the view of any production plant as a
possible supplier of not only products, but also energy products viz. electric power
and heat.
Research approach
We conjectured that the development of trigeneration may contribute to achieve
sustainability, notably by avoidance of CO2 emissions and by increasing total
resource utilisation. Our first objective was to apply the systemic approach
developed (Dijkema 1996) for identification of trigeneration options and the
specification of desired system content. Our second objective was to demonstrate
that options for system improvement in the petrochemical industry exist beyond the
straightforward application of cogeneration.
Thus the research questions addressed are (1) whether this method is applicable and
(2) whether trigeneration is worthwhile with respect to sustainability not only
ecologically but also economically.
We used functional modelling to develop trigeneration concepts. A straightforward
LP-model was constructed to demonstrate the effect of trigeneration on the
selection of energy conversion facilities. The model was also used to assess the
extent of CO2 reduction when effecting trigeneration in an industrial complex. Early
economic evaluation was used to obtain a first impression of the economics
involved.
Functional modelling
Applied thermodynamics teaches us that trigenerate conversions must be direct, in
order to obtain the highest efficiency, and traditionally an ideal fuel cell reactor is
assumed to prove the case (e.g. Denbigh 1956)). At the other end of the spectrum is
isolated or stand-alone conversion of fossil feedstock into chemicals, electric power
or heat respectively. Present petrochemical systems are somewhere in-between,
largely through the longstanding tradition of process- and energy-integration, while
the introduction and spread of the use of cogeneration also has improved its fuel
utilisation. This use of cogeneration can be considered a form of trigeneration.
In addressing the analysis and formulation of trigeneration system concepts we have
used functional modelling. The central idea of functional modelling is to enable
'opening-up' established or proven systems or technology concepts. System
engineering has been defined as “a progressive definition and integration of functions and
design decisions at every level of system complexity” (Corrigan and Kaufman 1965).
Systems can be decomposed into progressively smaller subsystems, where system
elements are the lowest level of detail of a system structure. System complexity can
now be dealt with effectively when the focus is on system functions rather than
system content because the functional characteristics of a system element are
Fuel Cells and Trigeneration 179
technology-free (Asbjørnsen 1992; Dijkema and Reuter 1999). In other words, a system
or system element can be described completely by its inputs and outputs and the
functions it performs. Specification of its technological content is not required;
rather the functional description is complete and independent from the actual
technical realisation of the element. For complex systems, functional modelling
offers a means to arrive at some functional decomposition (Dijkema and Grievink
2000). It may be seen that a technology-free description enabled by functional
modelling opens the way to search for alternatives in trigeneration system design.
In the case of the gas turbine, for example, flue gas is produced that contains a
significant amount of waste heat at a temperature of some 450-500 oC, instead of
transferring heat to some cold stream - e.g. cooling water. In the case of the fuel cell,
apart from energy the flue gas will contain unconverted feedstock and some cell
cooling may be required. For the sake of simplicity, in all cases we have assumed that
the feed is at the right conditions (Pressure, Temperature, composition).
We suggested that in process system design at each level of system complexity, the
same procedure can be applied: (Dijkema 1999). Trigeneration system development,
however, implies that there must be an initial focus on what possible alternative
realisations of a particular system (step 4 and 5 in the procedure given below) may
exist involving trigenerate functions. Thus, such a procedure must be extended with
(1) determination of the aggregate system level used to start the analysis, (2)
assessment of the (possible) system environment and (3) selection of the most
appropriate system boundary and system products.
In marketing studies, the latter is certainly done for a plant's main product, however,
a mechanism for identifying and evaluating other possible products should be
included. In trigeneration in the petrochemical industry, for example, CO2-credits
may be obtained that can be marketed under an Emission-trading system.
A procedure for trigeneration system development thus may look something like:
I. Review an existing system design by functional modelling:
1. Determine the aggregate system level
2. Identify the crucial system functions
3. Draw a functional scheme
4. Briefly describe each system element
II. Explore the options for trigeneration by
5. Assessment of the (possible) system environment
6. Addition of system outputs
7. Modifying the system boundary
8. Addition of system elements
9. Alternative realisations per system element or combination thereof
III. Specify and select system concept(s) for early evaluation
In addition to the global chemical and polymer industry that largely represent the
markets for petrochemicals, the system environment assessed for the case of
trigeneration must include the electric power consumers or traders, and possible
consumers of heat. Apart from surplus electric power and heat, system outputs to be
considered are tradable CO2 emission rights. The system boundary of the petrochemical
complex changes because facilities for the generation of power and heat are
included. These include stand-alone power and heat generation respectively,
cogeneration options and others. At the single plant level, subsystem boundaries
182 Process System Innovation by Design
may shift, e.g. because of the integration of gas turbines in furnace design for
optimal preheat and power generation. Finally, gas turbines, steam turbines, fuel cells
etc. represent possible added system elements and these may offer new options for
realisation or combination of system elements.
x = [c r p ]T Eq. 5.18
The optimal network design variables x0 are found by optimisation.
minx fT x subject to Ax = 0 Eq. 5.19
f is a vector of some objective function’s coefficients
The coefficients of the A-matrix listed in Table 5.2 reflect the performance or
efficiency of 'state-of-the-art' technology. These have been expressed in megawatts
(MW). A negative sign indicates an input, a positive sign an output. To avoid to
optimistic results with respect to trigeneration, the efficiency of stand-alone power
generation has been set at 50%. In this case, f is set such that r, which represents
total fuel input, is minimised. Thus, under the assumption that only fossil fuel
combustion drives these energy conversions in the installations listed, the net CO2
emission will also be minimised in this optimisation. Obviously, the actual CO2
emission depends on the fuel-mix employed. A first indication is obtained in all
cases by considering natural gas (CH4) as the fuel of choice.
Table 5.2: Coefficients used in the A-matrix. Includes net loss per element (MW)
System elements
Energy carriers Cogen-3 Cogen-4 Power StmGen-1 StmGen-2 Stmtrbine
Fuel -1 -1 -1 -1 -1 0
Electricity 0.24 0.19 0.5 -0.02 -0.02 0.14
Heat - 600 oC 0.71 0.76 0 0.85 0 -1
Heat - 200oC 0 0 0 0 0.85 0.76
LP model Results
In Figure 5-9 the results of the optimisation are given for the case where there is a
net demand in the Petrochemical complex of heat at 600 oC and 200 oC respectively
in a fixed ratio of 4:8:E (E=electric power) [MW].
From these figures it may be seen that:
- when export of electricity is not allowed, Cogen-4 caters for internal electricity
demand (0.5 MW), and there is stand-alone generation of 600 oC (HT) and 200
o
C (LT) steam
- when electricity export is allowed but limited, first Cogen-4 capacity is expanded,
subsequently Steam Turbine capacity is added
- allowing more electricity export brings Cogen-3 into being, thereby providing the
optimal match between electric power and heat required
184 Process System Innovation by Design
10
8
StmTurbine
6 Power Plant
4 Cogen-4
Cogen-3
2
0
0 1 2 3 4 5 6 7 8 9 10 11 12
-2
Total Electric Power Export Allowed [MW]
In words: the total (discounted) additional revenue flux R from the trigenerate
petrochemical complex throughout its technical lifespan θ must exceed the
difference in initial investment I for trigeneration and stand-alone power generation.
The initial investment for the trigenerate petrochemical complex may include both
energy-conversion equipment and cost of adaptation of petrochemical installations.
feasible, total CO2 emissions are reduced, which provides a clear incentive to convert
chemical complexes to trigenerate chemical complexes.
5.3.8 CONCLUSIONS
By introducing functional modelling, the options for the simultaneous production of
chemicals, power and heat can be explored at each aggregate level by translating
existing system concepts to a technology-free specification. Thereby, an opening is
created towards development of innovative concepts. A procedure to systematically
complete this task was formulated and illustrated for the case of trigenerate
petrochemical complexes.
Fuel Cells and Trigeneration 187
6.1 Abstract
Functional modelling has been applied to develop innovative system concepts for
the chemical industry to anticipate and exploit changes in logistic fuel demand
related to the market penetration of fuel cell vehicles. Within a decade polymer
electrolyte membrane (PEM) fuel cell vehicles may enable a true transition in
'automotive transport'. We conjectured that part of these will be methanol-fuelled.
Scenarios were constructed for automotive PEM fuel cell market developments,
technological progress and fuel diet. Simulations up to the year 2030 visualise the
dynamics and scope relevant for the chemical industry. In the conservative 'Superior
PEM vehicles' scenario, 10% of the annual sales of passenger vehicles and heavy-
duty vehicles are fuel cell powered by 2020. In addition, all combustion engine
passenger vehicles sold have a 42V PEM fuel cell auxiliary system on board. The
share of methanol in the fuel diet is expected to drop from an initial 100% to only
50% in all these markets. Technology advancement yields a 70% reduction in PEM
platinum-loading and an extension the economic lifespan of fuel cell stacks to 5
years. Technology to recycle fuel cells and recover platinum is developed.
In this scenario the production of the PEM fuel cell systems causes a peak in gross
annual platinum (Pt) demand of some 105 tons/year. Since fuel cell stacks are being
recycled, net demand for virgin Pt peaks at 95 tons per year around 2018, after
which it drops to a negligible level. Sustainable Pt use requires that a stack/Pt
recycling industry is developed in time. The main business opportunities lie in the
development of a stack platinum recycling industry and the associated transport
infrastructure because the long term dynamics of Pt-recycling appear much more
favourable than the dynamics of virgin Pt production.
In the chemical industry the main opportunity is the supply of methanol. The
scenario results in an annual demand of 70 MTA methanol after a period of high
demand growth. Using the methodology for process system innovation by design,
new objective-defined functions of the petrochemical industry and possible system
concepts for implementation in the Rotterdam industrial cluster were explored. One
option is the realisation of a unique integrated industrial complex around the flexible
manufacture and use of synthesis gas. The initial feedstock can be natural gas, but oil
residue, coal, biomass and biomass waste can also be converted to synthesis gas.
Methanol can consume a major share of the synthesis gas, but hydrogen
manufacture, metals processing, olefins and other chemical processes also benefit
from the variety of syngas qualities available in such a 'green' complex.
The only real barrier to these developments is the advancement of PEM fuel cell
technology and its final breakthrough to the marketplace.
The networked process system innovations presented illustrate the use of functional
modelling. Combined with scenario-building and dynamic modelling, a powerful set
of methods results to anticipate and respond to changes in the world outside the
industry. Thus, it enables the chemical industry to reap-the-benefits of a transition in
automotive transport.
Fuel Cell Vehicles and Industry Development 189
6.2 Introduction
The automotive transport sector worldwide today includes automobiles in all their
varieties, low- and heavy-duty trucks, buses, vans, and utility vehicles. Transport
accounts for roughly one-third of world energy resource use (WRI 2003).
Worldwide 20 million passenger vehicles and 600,000 heavy-duty vehicles (trucks,
buses) are sold each year (Anonymous 2003a; van den Hoed 2004).
The automotive sector represents a major market to the chemical industry. Today,
synthetic materials such as polymers, fibres and ceramics increasingly replace steel
bodywork and engine parts. Automotive fuel additives such as oxygenates are
produced in bulk quantities to augment gasoline quality. Its development continues
to create incentives for innovation and growth of the (petro)chemical industry.
Apart from the continuous stream of largely 'incremental' innovations, a transition
of 'automotive transport' initiated by the availability and breakthrough of polymer
electrolyte membrane fuel cell (PEMFC) technology could be realised in the near
future. Such transition will have major consequences for the petrochemical and
related industry. By first inspection of the requirements of PEM fuel cell vehicles
and the associated industrial activities, we conjectured that development of PEMFC
powered transport markets will open novel opportunities for networked process
system innovations around methanol49, hydrogen and carbon dioxide (CO2)
production. Available technologies may become part of an integrated system that
caters for transport, electricity, and product needs. These products are manufactured
via the generation of synthesis gas, which provides a generic link to a range of
conventional and novel feedstock, such as natural gas, coal, crude oil and biomass.
The use of biomass and CO2 sequestration contributes to sustainability. Thus, a
route towards a greener chemical industry can be opened.
Within a decade polymer electrolyte membrane (PEM) fuel cell vehicles may enable
a true transition in automotive transport, which is one of the most important outlets
of the chemical industry. We conjectured that alternative fulfilment of the objective-
defined function 'automotive transport' may be (in part) by methanol-fuelled PEM
fuel cell vehicles.
In this chapter, we report the use of functional modelling (Ch. 3) to explore the
possibilities, implications and opportunities related to the transition. This represents
a test case for the functional modelling approach and procedures developed:
Does functional modelling provide an adequate means for abstraction
from currently employed system concepts and do the procedures enable
or facilitate the specification of innovative system content?
The qualitative analysis is augmented with a quantitative assessment based on system
dynamics modelling and scenario analysis, in order to obtain a first impression of
the impact and constraints in potential transition scenarios.
49Methanol can be used in combustion engines but has never penetrated the logistic fuel markets on
a significant scale. Amongst others the Association of Methanol Producers for many years has
advocated the use of methanol as a suitable fuel for passenger vehicles. Such use of methanol
represents a major opportunity for those involved in the methanol supply-chain - natural gas
producers, methanol companies and suppliers of technology and equipment - to grow their business.
190 Process System Innovation by Design
'Automotive transport' today comprises 'individual' vehicles that are able to move as
stand-alone items and 'system-linked' vehicles that are always attached to some
infrastructure. Existing and future realisations of individual vehicles (aircraft, marine
vessels) can be categorised around (1) the conventional combustion engines, (2)
electrochemical fuel cells, (3) batteries, (4) pneumatic systems and (5) solar or wind-
powered systems. System-linked vehicles comprise (6) combustion engine powered
and (7) electrically powered. The system-link may be rail tracks (diesel trains),
electricity supply from the grid (trolley-buses), or both (electric trains, metro
systems, city trams). Special system-links are sky trains and automatically guided
vehicle systems (AGVs).
As a matter of course, (1) represents current realisation of ' individual vehicles' such
as automobiles, trucks and auto buses. Presently, gasoline, diesel and LPG are
employed worldwide as fuel, while ethanol, bio-diesel and natural gas are employed
in various regions at a limited-scale only. The use of methanol, diesel/methanol
Fuel Cell Vehicles and Industry Development 191
mixtures and dimethylether (DME) have been demonstrated. Only recently, BMW
has developed a prototype of their 730 model that operates through internal
combustion of pure hydrogen.
Options (3) and (4) appear to be developments for transport with limited range.
Vehicles for city transport have been developed and tested that use compressed air
to provide shaft power. The use of battery-powered vehicles is largely limited to
utility vehicles. The weight to energy-storage ratio is less favourable for long-range
transport using the current and emerging generation of batteries. Advanced battery
subsystems are being incorporated in PEM fuel cell systems to cater for peak- and
start-up power. Options (6) and (7) are the well-known realisations of public
transport.
Compared to transport based on (1) combustion engines, options (2) around
electrochemical fuel cells are expected to achieve much higher 'well-to-wheel'
efficiencies whilst largely avoiding harmful emissions (Geissler et al. 2001). Initial
developments on alkaline fuel cells and phosphoric acid fuel cells were not
successful. Today, polymeric electrolyte membrane, PEM, appears to be the fuel cell
technology most suited for automotive application. Various demonstrations have
been launched, which all featured conventional vehicles propelled by PEM fuel cell
systems instead of combustion engines.
Since PEM fuel cells electrochemically convert hydrogen, on-board storage of
hydrogen and an infrastructure for hydrogen distribution is required. Many studies
and efforts of governments, fuel cell technology developers and automobile
companies focus on on-board storage of hydrogen and the associated infrastructure
for hydrogen supply. In the refining and petrochemical industry, however, it is well
known that hydrogen is expensive to manufacture, to handle and to transport. In
contrast, the use of logistic fuels for fuel cell vehicles that incorporate on-board
conversion to hydrogen would require only limited adaptation of the present
industrial infrastructure (refining, gasoline stations etc.). Therefore, developments
are under way to demonstrate that on-board fuel conversion to hydrogen is
technically and commercially feasible. Fuel candidates are (clean) gasoline, diesel,
LPG, ethanol, methanol, and DME. The best result would be the realisation of so-
called Flexible Fuel Reformer (FFR) technology, because this would open the way to
economy-of-scale in the mass production of only a single type of reformer for
various fuels50.
or both. One may think of, for example, bicycles, solar powered vehicles, aircraft,
marine vessels, sailing ships, magnetic trains, hot-air balloons, rockets, zeppelins,
snow gliders and sky trains.
Therefore, we prefer the concept of 'automotive transport'. Its functional model
essentially consists of the functional elements (1) 'realising availability': some
(energy) resource is made available to be used for automotive transport and (2)
'realise automotive transport': employ or convert the energy resource to (shaft)
power that drives a suitable vehicle. Thus, the functional model for automotive
transport is similarly decomposed as the model for petrochemical systems around
aromatics (Ch. 3) and Olefins (Ch. 4), where 'production and consumption' systems
were discerned.
Emissions(non-ODF)
FE 1
FE 2
Fossil reso
Provide suitable
Provide Transport
urces energy source
automotive of goods
transport
Renewable FE 1
resources Energy Transport
FE 1
Source of people
FE 2
es FE 2
sourc
Flow Transpo
rt
FE N of livesto
FE N ck
Emissions(non-ODF)
FE 1 FE 2
Foss FE 3
resou il
Convert to Distribute Provide Transport
rces suitable power power source automotive of goods
source
transport
FE 1
FE 1 FE 1 Transport
Energy Source of people
ble
ewa FE 2 FE 2 FE 2
Ren urces Trans
p
reso of live ort
stock
FE N FE N FE N
so low
s
ce
F
ur
Reduction of fossil fuel use and CO2 emission of 'automotive transport' requires a
transition in vehicle technology (FE 3) and an alternative realization of resource
conversion and distribution (FE 1 and 2; see Figure 6-3).
Cadre 6-1: From Well-to-wheel with methanol and PEM fuel cells
One option is the fulfilment of the ODF 'automotive transport' by PEM fuel cell
vehicles. Since on-board conversion of methanol to hydrogen is technically and
commercially feasible, PEMFC vehicles may be built with on-board storage of
hydrogen or methanol. In both cases, the associated fuel supply and distribution
infrastructure must be developed. In contrast, the use of logistic fuels - (clean)
gasoline, diesel, LPG - would require only minimum changes in supply and
distribution.
Figure 6-4: 'Automotive transport' realisation by PEM fuel cells and methanol.
196 Process System Innovation by Design
As a matter of course, the use of methanol and logistic fuels will result in
production- related CO2 emission and diffusive emission from the vehicles. The use
of methanol is expected to yield the smallest amount of total emission. In addition,
where presently almost all methanol is produced from natural gas, an efficient
system would result when combining the production of methanol from natural gas
and biomass (Stikkelman 2001).
High
Competitiveness
Unimproved Improved
Economics Economics
'PEMFC, 'PEMFC
the eternal promise' Outcompeted'
Science & technology Market power of
cannot deliver feasible vested players
solutions remains
unchallenged
Low
Competitiveness
In 'Superior PEM vehicles' the economics are sufficiently improved and incentives
have been created so that PEMFC vehicles are considered superior goods. Thus,
they will obtain a share of the market that depends on economic prosperity.
Conventional gasoline and diesel vehicles continue to dominate the world's vehicle
fleets. Many automobile companies do introduce novel models with PEM fuel cells
to deliver auxiliary power.
'PEM Breakthrough' materializes when general awareness of the threat of Global
Warming, technological progress, venture capital investment and clever marketing
result in a sweeping introduction and acceptance of PEMFC vehicles worldwide.
Subsequently, in OECD economies, in a short period of time legislation is passed
and enforced to phase out gasoline and diesel operated vehicles. The time of
abundant, inefficient use of oil products in transport ends relatively abruptly. A
massive restructuring of the refining industry worldwide follows to supply logistic
fuels suitable for use in PEMFC vehicles. This includes natural gas and crude oil
conversion into methanol and hydrogen combined with carbon fixation or CO2
sequestration.
At present, it may be seen that PEMFC vehicle development resides in the lower-
left quadrant. Indeed, many are of the opinion that fuel cell technology will remain
an eternal promise. Recently, however, considerable investments in R&D
programmes, demonstration and commercialisation have been launched to move the
technology to the upper-left quadrant. Whether the 'Superior PEMFC vehicles' is
the most plausible scenario remains unknown. There is, however, considerable risk
that said programmes are not sufficient, which would imply that 'PEMFC
outcompeted' is the near future where 'PEMFC-the eternal promise' remains the
perception of many believers. A dramatic success of R&D would open the way to
realise the ultimate scenario, 'PEMFC breakthrough', which is also the most
dramatic of possible futures.
51The model equations for each market segment are the same; only the model parameters differ
between the market segments. The model is given in Appendix A. 8.
202 Process System Innovation by Design
The balance of inputs and outputs is used to calculate the 'presence' of systems
produced in each previous year. In the 'Superior PEMFC vehicles' scenario (see
below), for example, in 2003 the first 10 light-duty passenger vehicle 30 kW systems
are introduced. These have an expected technical lifespan of 1000hrs of operation.
They are used 200 hrs per year to drive some 16,000 kilometres. Thus their average
technical lifespan is five years. One of the systems, however, breaks down to a status
beyond repair in 2003, the others are in operation successfully until 2007 (1), 2008
(2), 2009(4) and 2010 (2). Thus the actual operational lifespan shows some variation
around the 1000hrs design value, which reflects the effects of variations in style of
use.
The phase-out or 'end-of-life' scenario and the technology status at time of
production determine the net volume of platinum to be recycled. In the model,
therefore, the variation by which systems break-down can be specified. Also the
fraction can be specified of systems that are removed from the world-wide fleet in
operation due to other reasons than PEM fuel cell breakdown (accidents, fire,
change of lifestyle, misuse etc.).
The combination of scenarios for market demand (input), system use and end-of-life
(output) provides a projection of the running fleet. From this dynamic forecast of
the size of the running fleet, its characteristics and use, the implications for
methanol and platinum demand can be computed. In addition, the net amount of
CO2 and other emissions can be calculated.
In order to get a realistic impression of the dynamics and scope for platinum and
methanol, the key parameters that are affected by technological progress have been
incorporated in the model. These are the total Pt-loading required (expressed as g Pt
per net kW system output), the system expected operational-economic lifespan (<
5% power degradation), and the overall system efficiency of methanol-to-mileage.
From the typical system power-rating and use pattern the net demand for methanol
is calculated. The total methanol demand is determined by the sales volume. In the
model, the fraction of methanol-fuelled PEMFC can be varied.
The typical power-rating and technology status with respect to required Pt-loading
determine the net Pt-demand for a system produced in a particular year. Once
produced, the net amount of Pt remains unchanged until the system ceases
operation and thus becomes available for recycling. In the calculations, a 100%
platinum recovery from these systems has been assumed.
In summary, the dynamics of future PEMFC-related methanol demand, the net
'virgin' platinum demand, the demand for Pt-recycling and the net amount of Pt-
stored in the PEMFC fleet are calculated. The model can be extended to visualise
future (avoided) CO2 and other emissions, the net content of Pt (g/kg) in systems
recycled (requires forecast of development of PEMFC specific weight development),
and the amount of hydrogen required for the hydrogen-fuelled fraction of the
PEMFC fleet.
Fuel Cell Vehicles and Industry Development 203
52 At a typical Pt price of some $15/g. Current prices can be found e.g. at www.kitco.com.
204 Process System Innovation by Design
5 Tech.
Lifetime
Oper.
4 [Years]
[g/kW], [Years]
2 Platinum
Loading
[g/kW]
1
0
2005 2010 2015 2020 2025 2030
Year
Figure 6-6: Stack development scenario: Platinum loading and technical lifespan for
automotive PEMFC systems.
Market development expressed as the number of systems sold determines the order-of-
magnitude of the platinum and methanol demand, respectively. Today, worldwide
20 million passenger vehicles and 600,000 heavy-duty vehicles (trucks, buses) are
sold each year (Anonymous 2003a; van den Hoed 2004). In the conservative
'Superior PEM vehicles' scenario, 10% of these annual sales are fuel cell powered by
2020.
200 Auxiliary
15
Million
Million
150
10 Heavy-
100 Duty*100
5
50 Light-Duty
0 0
2005 2010 2015 2020 2025 2030 2005 2010 2015 2020 2025 2030
Year Year
This would imply that some or all of these markets may never blossom. In other
words, the limited size of the world platinum market appears to be a serious
constraint for PEMFC vehicle market development.
Stationary - Small
(1-5kW)
PEM FC
Passenger Vehicles
Production (30 kW)
Trucks, Buses
(300 kW)
Vehicles Auxiliary
(0.2 kW)
Natural Supply MeOH Distribution
Winning
Gas Infrastructure production Infra
MTBE>Gasoline
additive
Platinum
Pt- Platinum Catalysts for oil refining MMA>PMMA windows
Mining Markets /
Ore Extraction
Distribution Catalyst for chemical plants Other: (plastic)
materials
Computer harddisks
Figure 6-8: Overview of related platinum, methanol and PEM fuel cell markets.
Therefore, we calculated a forecast for a period of 30 years. The results are
visualised in Figure 6-9. The combined developments - the number of systems sold
increases while the amount of platinum per system decreases - result in a peak of
total Pt-demand of 160 tons per annum. Due to the lifespan of PEMFC power
systems, Pt-supply from PEM Pt-recovery operations trails behind some 5 to 7
years. In the scenario, however, the Pt recycled from one early vehicle can be used in
multiple new vehicles that use more advanced technology. As a consequence, in the
scenario virgin-Pt demand peaks at some 100 ton per annum in 2012 and decreases
rapidly to extinction around 2020 when markets and technology have matured.
Finally, the amount of platinum from recycling will match or exceed net Pt demand
for new systems, in this scenario after some 30 years. Thus, a net inflow of Pt from
the PEMFC fleet changes to a net outflow of Pt from the PEMFC fleet. This
implies that Pt-production capacity must be carefully planned and that capacity
required for Pt recycling is fairly predictable and reaches some steady-state level.
Obviously total PEM related platinum demand and virgin demand levels will go
down if technology advancement with respect to Pt-loading is more successful. The
shape of the curve, however, remains the same: total Pt-demand will rise until
PEMFC markets have matured (in our scenario: around 2022, i.e. in 20 years time),
and then drop to a quasi steady-state level that is a function of technology status.
In Figure 6-9 the total amount is plotted of Pt stored in the PEMFC automotive fleet
as well: the first 20 years, it grows dramatically, then levels off to a steady-state of
some 800 tons (!). Thus, the long-term dynamics of the worldwide Pt material cycle
have been elucidated.
An important question is: can the systems in the worldwide platinum material cycle
respond adequately to the dramatic dynamics resulting from only moderately
Fuel Cell Vehicles and Industry Development 207
125
100 Pt - demand
75
[Ton/Year]
50 Pt - recovered
25
0 Pt - virgin
2005 2010 2015 2020 2025 2030
-25
Year
Figure 6-9: Model forecast for world Pt-demand, 'Superior PEM vehicles' scenario.
From statistics compiled by Johnson & Matthey (www.johnsonmatthey.com), it can
easily be seen that Pt supply has doubled since 1975, from a net supply of 81
ton/year to the current level of 174 ton/year. This development has occurred largely
in response to continuously increasing demand. In their most recent yearly outlook
(2001) Johnson Matthey quote that for example South Africa have huge yet
undeveloped Pt resources, and that both prospecting activity worldwide has
increased as well as expansion of mining and refining planned. In addition, where
(e.g. Hampden-Smith 2000) states that Pt markets are extremely volatile and its price
unpredictable, Johnson Matthey indicates that the relatively large jewellery market
share effectively results in a price-cap at some $650/troy oz. where jewellery
consumers will simply stop buying the material, which would allow a shift from e.g.
jewellery to fuel cell manufacture by Pt. For example, in case jewellery consumption
would go back to 1991 levels, some 44 tons of Pt would become available on the
market. Supply and demand data also show that with the introduction of auto-
exhaust catalysts and the development of a new market outlet for Pt, a few years
later also Pt becomes available from recycling of said catalysts for, amongst others,
the recovery of Pt. The historic data suggest that a net decline in Pt-demand for
automobile exhaust catalyst is foreseeable. Together with the continuous growth in
supply, there is thus no reason to believe that Pt availability is a constraint for the
development of a fuel cell vehicle fleet!
208 Process System Innovation by Design
80
60 MeOH demand
[MTA], [MTA/Yr]*10
40
20
MeOH new
0 capacity*10
2005 2010 2015 2020 2025 2030
-20
Year
Figure 6-10: Total automotive methanol demand and yearly capacity growth.
An increase in PEM fuel cell system efficiency will only imply a somewhat slower
growth in methanol demand after 2010, when the more efficient systems are put on
the road. The effects are minimal in relation to the uncertainties in market
development and mileage covered per vehicle. The shape of the curves remains the
same: methanol -demand will rise until PEM fuel cell markets have matured (in our
scenario: around 2022, i.e. in 20 years time), and then level or drop slightly to a quasi
steady-state level that is a function of technology status and fuel diet.
There are three underlying questions to ask whether such development of methanol
demand represents a constraint to PEMFC vehicles market development:
1. What are the long term dynamics of the worldwide methanol market?
Can this system adequately respond to the forecast - temporary - dramatic
growth in methanol demand for PEM fuel cell related use?
2. Does automotive transport through PEM offer a “sure” future methanol outlet?
Although investments in PEM fuel cell research and development have been
large and an increasing number of companies are entering the PEM fuel cell
community, the current status of the technology is still not good enough to
make large-scale introduction a commercial success (see Appendix A. 9). This
would imply that the 'PEM outcompeted' scenario or 'PEM - the eternal
promise' become reality Figure 6-5, p. 204).
3. What are the costs and risks of a methanol distribution infrastructure?
Can a feasible strategy be developed to bring-into-being a methanol distribution
infrastructure that supports the operation of a PEMFC vehicle fleet?
(1) Long term dynamics. In the Superior PEM vehicles scenario, the methanol required
for PEM automotive fuel cell systems (Figure 6-10, p. 208) represents a completely
new type of demand. Ultimately a demand of 180 MTA methanol is created, which
is about 5 times the current methanol output levels and between 2015-2030 100 new
world scale plants would have to be realised. The effects on world oxygenate
production for the gasoline is considered to be minimal.
The present worldwide nameplate capacity for methanol is some 35 MTA, and
present level of the demand54 is some 30 MTA (Tazelaar 2001). This is a normal
balance between capacity and demand at an utilisation of some 85%. Supply and
demand data also show that with the introduction of oxygenates-enriched fuel,
methanol demand in existing fuel has steeply increased in the past two decades. In
the period 1994-2000 for example worldwide capacity increased from 26 MTA to 35
MTA, or some 1.5 MTA per year.
(2) Uncertainties At present, some older methanol capacity is being phased out, whilst
a few new large scale plants are under construction and various studies for
“Megascale” methanol have been initiated (Tazelaar 2001). There is a concern,
however, that a possible ban on MTBE in the US will lead to a situation of serious
oversupply. As a consequence, world methanol prices may become depressed just at
a time when demand from PEM fuel cell cars may be growing. Today, a new
methanol plant can be erected and operational in some 3-4 years time. The question
really is: can new capacity added match the demand growth-rate .
54 At present, methanol is used in a variety of chemical and fuel applications. In Western Europe,
only the volumes used for formaldehyde manufacture exceed those for gasoline additive production.
MTBE is the most important oxygenate that replaces lead-additives in gasoline to ensure proper
combustion (anti-knock) of the fuel. Other important -chemical- applications include the production
of acetic acid and methyl methacrylate. The chemical markets can be considered mature. The MTBE
market faces some uncertainty because of possibly forthcoming legislation banning its use in gasoline
in the U.S. Substitute oxygenates are being introduced, but some of these also require methanol.
210 Process System Innovation by Design
The forecast calculated for automotive methanol demand alone implies a definite
change of the character of methanol from chemical to fuel. PEMFC methanol
demand may have dramatic impact on the structure of the worldwide industry. The
trend towards megascale methanol plants will be accelerated and older plants that
have operating capacities well below the megascale plants may be phased out.
As in platinum, the evolution of worldwide methanol production capacity must be
carefully considered before adding new methanol capacity. Contrary to Pt, in such
an analysis, the effects of existing methanol plants being outcompeted and
subsequently being phased out must be taken into account, as well as the possible
ban on MTBE, the world's largest current methanol outlet.
(3) Distribution infrastructure. The development of methanol distribution to fuel service
stations for the consumer market has the characteristics of a catch-22; because the
number of methanol-fuelled cars is limited, it is not attractive to invest in
distribution infrastructure; because of lacking distribution infrastructure, consumers
do not invest in methanol-fuelled cars.
In recent years by introducing Shell PuraTM this company has shown that in a
relatively short time period an additional fuel outlet can created. The World
Methanol Institute (1999) has calculated the installation costs for methanol outlet
per station to some 80,000 euro55 (Stikkelman 2001). In the Netherlands the track
record of change of fuel stations is very good. Companies, for example, have
invested significantly in environmental precautions induced by legislation. The
government funded part of the changeover costs. There is no reason why a similar
scheme could not be developed for the expansion of fuel stations to include
methanol outlets.
The Netherlands is home to a total of 2600 fuel service stations (CBS). The total
cost of a maximum fine-grid of methanol distribution in the Netherlands therefore
would amount to 2,600 * 100,000 = 260 Million Euro, which is about half the
investment for a single world scale methanol plant. Together with the present
growth in natural gas supply, there is thus reason to believe that
1. The development of a methanol distribution system is no less feasible than the
development of any other fuel distribution system, and thus with careful
business development is not expected to present a problem.
2. Ample methanol supply capacity can be developed in time, and thus methanol
shortage or high prices do not by themselves constrain the development of a
fuel cell vehicle fleet.
3. The status of the technology and the pace of technological progress largely cause
uncertainty with respect to the development of methanol PEM fuel cell markets.
PEMFC vehicle development may result in methanol becoming a new 'logistic fuel',
which opens the way for dramatic investment in methanol production capacity and
distribution facilities.
55This quote includes all the necessary work and hardware per station. As a matter of course, other
costs will be incurred to create a methanol-to-consumer distribution system, e.g. for the set up or
adaptation of administrative and other information systems.
Fuel Cell Vehicles and Industry Development 211
Figure 6-11: The opportunities for the PoR in the world platinum material cycle.
214 Process System Innovation by Design
Today, the industrial markets for hydrogen represent a large-volume market56, where
regional, national and international pipeline grids for hydrogen transport have been
established. In the PoR area, for example, Air Products exploits a regional hydrogen
infrastructure, and Air Liquide exploits a 'basin' hydrogen infrastructure that spans
Rotterdam, Antwerp up to Lille in France.
The network, however, largely has the function of a backup facility for local
hydrogen production facilities, which indicates that hydrogen transport over large
distances by pipeline is still uneconomic.
Worldwide pure hydrogen is produced predominantly by steam reforming of natural
gas, which has favourable economics compared to the use of other fossil resources.
The syngas produced is subject to a number of CO-shift and purification processes.
In crude oil refineries hydrogen make-up (95% hydrogen) is presently also produced
via residue gasification followed by treatment of the product gas. Due to the nature
of refinery processes, a significant level of (CO) content is allowed, which makes
this route economic. Depending on the CO-tolerance of PEMFCs, both hydrogen
production from natural gas or oil products may become economic in the future.
Alternative sources of (relatively pure) hydrogen are a number of petrochemical
processes that include dehydrogenation, notably styrene.
The existence of hydrogen pipeline infrastructure in the PoR offers the possibility to
tie in both 'virgin' and 'waste' hydrogen production, as has been done with the Air
Liquide network, which takes in purified waste hydrogen from the Dow Terneuzen
styrene plants and the General Electric chlorine plants in Bergen op Zoom.
The possibilities with respect to hydrogen in relation to fuel cells have been
indicated in Figure 6-11 and Figure 6-12. Hydrogen is part of the precursor of
methanol, synthesis gas that is a mixture of CO, CO2 and H2. As the demand for
hydrogen increases, possible synergies may emerge in the production of synthesis
gas for methanol and hydrogen production respectively. In both cases, synthesis gas
can be produced out of a variety of resources, the synergy being that for each mix of
resources available an optimal mix of syngas qualities can be produced: for hydrogen
production the syngas H2 :CO ratio needs to be as high as possible, whilst methanol
synthesis requires an optimal ratio of 2:1, and allows the presence of some CO2.
Thus, natural gas and biomass can be combined to yield an optimal syngas
composition.
In the PoR and in the region the hydrogen can be distributed via pipeline or bottles.
A facility for small containers can readily be added to the existing facilities. In other
studies, the use of hydrogen as an additive to the Dutch natural gas grid for
households has been advocated. Hydrogen capacity thus can be built to first cater
for this market, and later to cater (partly) for the fuel cell market, thus providing a
smooth growth and/or transition. The combination of a hydrogen pipeline
infrastructure and a growing market for hydrogen may provide synergy to other
industries that generate hydrogen as a co-product that in other locations must be
used for (furnace)-fuel only.
56 The small-volume market of hydrogen is mainly catered for by bottled hydrogen in industrial
cylinders. These are transported by road-cargo. An intermediate-volume market also exists, where
hydrogen is generated at a customer-site.
Fuel Cell Vehicles and Industry Development 215
57At the beginning of the twentieth century a similar opportunity in Rotterdam was seized by the
Bataafsche Petroleum Maatschappij. The construction of its refinery in Pernis initiated the
development of the Port of Rotterdam into one of the world's major transport hubs and conversion
centres for crude-oil and oil products. Shell Refining Company still operates what today is one of the
four worldscale oil refineries located in the Port of Rotterdam area.
216 Process System Innovation by Design
vehicles, the fuel mix to be produced will change. Possibly, aromatics will be phased
out in logistic fuels for PEM altogether, as these require more sophisticated on-
board reforming, if at all feasible. Similarly, MTBE, which is required for anti-knock
properties (RON), may not be required anymore as a gasoline constituent.
A second important sector that will be affected is the world's platinum material cycle
(and Ruthenium) because PEM fuel cells do contain significant amounts of these
precious metals. The key parameter is the total amount of Pt per fuel cell system.
Today, the major uncertainty is the speed of technology advancement required to
bring down this amount. Backed by a scenario study, we conclude, however, that
stack lifetime and stack Pt-content do not pose a serious problem with respect to
bringing-into-being a worldwide PEM fuel cell system.
We conclude that the only real barrier to all these developments is the technological
advancement of PEM fuel cell technology and consequently the final breakthrough
of the technology to the marketplace. Not the platinum content, not the fuel
consumption, nor the emissions, but simply the total anticipated lifecycle cost to a
PEM vehicle owner is crucial to final demand for these systems. The present barrier
may very well be overcome by the already initiated introduction of small-scale
systems for vehicle auxiliary power generation. PoR must then be prepared to
facilitate the launch of the development of said synthesis gas cluster and further
develop its position in the world's platinum markets by developing the PEM Pt-
recycling infrastructure.
The only reason technology advancement is mandatory is to bring down system cost
significantly, and improve user characteristics. In our analysis, we have concluded
that, with adequate planning and management of the world's platinum material cycle,
availability and these precious metals need not be a problem.
waste. The possible 'course-of-events' sketched illustrates that this could very well
provide a route towards real greening of the Port's chemical industry.
Interpreting these developments in combination with the existing Port facilities and
Industrial Cluster in the PoR area, the timely launch of a PEMFC-AGV project can
provide a nucleus and a driver for a methanol or synthesis gas based industrial
infrastructure or complex in the Port of Rotterdam.
7.1 Introduction
Process system innovations are defined as changes in the system structure or system
design of an industry, its industrial complexes, or individual plants that can be
enabled by technological inventions or vice versa. The central theme of this thesis is
the specification of process system innovation content for a sustainable
petrochemical industry. The main result is a methodology for said specification that
is based on system representation using the input-output paradigm and functional
modelling.
In the subsequent sections, the innovations specified are reviewed. The
contributions to (engineering) science are discussed and the overall research
approach is reflected upon. Finally, recommendations are given.
59A premises of the M.Sc. in Systems Engineering, Policy Analysis and Management curriculum of
TU Delft is that multi-stakeholder decision processes can be engineered to address effectively a
variety if problems and opportunities that for example exhibit a varyiety of scale, complexity and
impact over a number of timescales,
226 Process System Innovation by Design
7.5 Recommendations
Focus on sustainable “supply chain level” or “networked industrial systems.” Sustainability
cannot be achieved in petrochemical plants or by promising technologies alone.
These must leverage system sustainability of supply chains or networked industrial
systems that adequately link with multiple interconnected material cycles.
Integrate or relate process system engineering and industrial ecology. The methodology
developed and the theoretical formulation provide a bridge between industrial
ecology concepts and the body-of-knowledge of process system engineering. Only
one example is the conceptual specification of a coal-based industrial complex for
base metal, hydrogen and synthesis gas production that is currently contemplated by
Port of Rotterdam and the industrial stakeholders.
Explore the implications of biotechnology for the petrochemical industry. In the long run, fossil
resources must be replaced by renewable bio-resources. If the slate of high-volume
petrochemicals remains and define the sector, the nature of the conversion
technologies must change from a chemical conversion paradigm to a biological
paradigm. Or some hybrid conversion paradigm may be adopted. The related
innovation space around biotechnology must be explored.
Efficient fossil resource use by trigeneration. Trigeneration would contribute significantly to
CO2 reduction. The steam cracker is one of many candidate processes. The
feasibility of trigeneration in liberalised and re-regulated energy markets must be
evaluated, the incentives of CO2 emission-trading elucidated whilst taking into
account the possibilities of modern furnace technology and design.
Build a coherent set of models and tools for innovation. Qualitative tests involving experts
should be developed to further underpin the approach. Quantitative evaluation
should be extended and incorporate dynamics and economics of system innovation
concepts. Thus, a portfolio of generic tools to foster innovation would result.
Transfer and use of research results in industry and society at large
The methods can be used to address 'wicked' environmental problems of industrial
sectors and society at large. They will foster the specification of innovation content
and will particularly be transferable to the sectors characterised procedurally
cohesive as the 'process industry': refining, base metal, pulp and paper, agro- and
food industry. Their use need not be limited to these sectors or sustainability!
Link to policy analysis and strategy Functional modelling must be linked to
(inter)national public policy development and regulation and corporate strategy
development. While the conceptual process system innovations require further
detailing, we trust the examples presented stimulate you as a stakeholder in any
process industry characterised by maturation to help reinvent your industry.
Incorporate the methods in transition management
As one of 'the industries of industry', a petrochemical industry is essential to achieve
sustainability. Its relation with the public at large is indirect and characterised by
lack-of-interest, discomfort and lack-of-acceptance. Use of the methods, however, is
expected to forge a link between societal needs, business strategy and process system
innovation as these can facilitate interaction in process development teams and
between stakeholders involved in a transition towards a sustainable society.
References
Abernathy, W.J. and J.M. Utterback (1978). "Patterns of Industrial Innovation." Technology
Review 80: 40-7.
Albright, L.F., B.L. Crynes and S. Nowak (1992). Novel Production Methods for Ethylene,
Light Hydrocarbons, and Aromatics. Dekker, New York.
Allenby, R.B. and J.B. Richards, Eds. (1994). The Greening of Industrial Ecosystems. National
Acadamy Press, Washington.
Alqahtany, H., D. Eng and M. Stoukides (1993). "Methane Steam Reforming Over Fe
Electrodes in a Solid Electrolyte Cell." Energy & Fuels 7(4): 495-504.
Anderson, J.L., Ed. (1996). Process Synthesis. Academic Press, New York.
Anonymous (1990). Chem-facts Netherlands. Chemical Intelligence Services, London.
Anonymous (1991). Chem-facts Ammonia. Chemical Intelligence Services, London.
Anonymous (1995a). Opportunities in Refining and Feedstocks beyond 2000. Report for Rotterdam
Municipal Port Management, ChemSystems. London.
Anonymous (1995b). Ullman's encyclopedia of Industrial Chemistry.
Anonymous (2000). Ullmann's Encyclopedia of Industrial Chemistry.
Anonymous (2001a). BP Statistical Review.
Anonymous (2001b). "Special Issue South East Asia." European Chemical News.
Anonymous (2003a). Scania Annual Report 2003. english.annualreport2003.scania.com.
Anonymous (2003b). Survey. European Environmental Agency.
Anonymous (2004). Statistical Review of World Energy. BP. www.bp.com.
Arnbak, J. and J.H. Weber (1999). Transmission systems engineering. lecture notes et4-036;
TU Delft, Faculty of Information Technology and Systems, Department of Electrical
Engineering.
Arthur D. Little (2000a). Cost Analysis of Fuel Cell System for Transportation - Baseline
Cost Estimate. Report to Dept. of Energy.
Arthur D. Little (2000b). "Distributed Generation: Understanding the Economics."
www.arthurdlittle.com.
Asbjørnsen, O.A. (1992). System Engineering Principles. Skarpodd. Houston.
Asbjørnsen, O.A. (1995). System Engineering at the Norwegian Institute of Technology.
Annual Report. Trondheim.
Asselbergs, A. (1998). "De chemische industrie ontmengt." NPT-Procestechnologie
6(November/December).
Aström, K.J. and B. Wittenmark (1984). Computer Controlled Systems. Prentice-Hall,
Englewood Cliffs, N.J.
Avramenko, Y., L. Nystrom and A. Kraslawski (2002). Selection of Internals for Reactive
Distillation Column - Case-based Reasoning Approach. European Symposium on Computer
Aided Process Engineering 12 (Escape-12), 157-162. Elsevier Science, The Hague.
Ayres, R.U. and L.W. Ayres (1996). Industrial Ecology; Towards Closing the Materials Cycle. Elgar,
Cheltenham.
Ayres, R.U. and L.W. Ayres, Eds. (2002). A Handbook of Industrial Ecology. Elgar, Cheltenham.
Azapagic, A. (1999). "Life cycle assessment and its application to process selection, design
and optimisation." Chemical Engineering Journal 73(1): 1-21.
228 Process System Innovation by Design
Bakshi, B.R. (2002). "A thermodynamic framework for ecologically conscious process
systems engineering." Computers & Chemical Engineering 26(2): 269-282.
Bansal, V., J.D. Perkins and E.N. Pistikopoulos (2002). "A case study in simultaneous design
and control using rigorous, mixed-integer dynamic optimization models." Industrial &
Engineering Chemistry Research 41(4): 760-778.
Bansal, V., J.D. Perkins, E.N. Pistikopoulos, R. Ross and J.M.G. van Schijndel (2000).
"Simultaneous design and control optimisation under uncertainty." Computers & Chemical
Engineering 24(2-7): 261-266.
Baylin, E.N. (1990). Functional Modeling of Systems. Gordon and Breach Science Publishers,
New York.
Bekkum, H.v. (2001). TU Delft. Personal Communication.
Biegler, L.T., I.E. Grossmann and A.W. Westerberg (1997). Systematic Methods of Chemical
Process Design. Prentice Hall PTR, Upper Saddle River, NJ.
Blair, R.N. and C.W. Whitston (1971). Elements of Industrial Systems Engineering. Prentice Hall,
Englewood Cliffs, NJ.
Blanchard, B.S. and W.J. Fabrycky (1998). Systems Engineering and Analysis. Prentice-Hall,
Upper Saddle River.
Bogle, D. and T. Perris (1999). "CAPE.NET - Responding to the business challenges."
Computers & Chemical Engineering 23: S779-S782.
Boot, A.H. and F.G.H. van Wees (1982). Industrial Process Heat in Relation to the
Temperature Level. ECN. Petten.
Boulding, K.E. (1956). "General Systems Theory - the Skeleton of Science." Management
Science 2 (3): 197-208.
Bouma, J.M., A.G. Cramwinckel and P.C. Klein (1996). De produktie van epichloorhydrine
volgens het Showa-Denko proces. Fabrieksvoorontwerp FVO Nr. 3176, TU Delft.
Delft.
Boustead, I. and G.F. Hancock (1979). Handbook of Industrial Energy Analysis. Horwood,
Chichester.
Bras-Klapwijk, R.M. (2001). Adjusting Life Cycle Assessment Methodology for Use in
Public Policy Discourse. Dissertation, TU Delft. Delft.
Brennan, D. (1998). Process Industry Economics: an International Perspective. IChemE,
Rugby, UK.
Bretz, R. and P. Fankhauser (1997). "Life-cycle assessment of chemical production
processes: A tool for ecological optimization." Chimia 51(5): 213-217.
Burgess, A.A. and D.J. Brennan (2001). "Application of life cycle assessment to chemical
processes." Chemical Engineering Science 56(8): 2589-2604.
Buxton, A., A.G. Livingston and E.N. Pistikopoulos (1997). "Reaction path synthesis for
environmental impact minimization." Computers & Chemical Engineering 21(1): S959-S964.
Carnot, S. (1824). Re'flexions sur la Puissance Motrice du Feu, et sur les Machines Propres a
De'velopper cette Puissance. Sadi Carnot
Carson, R.L. and L. Darling (1962). Silent Spring. Riverside Press, Cambridge, Mass.
Caserza, G., T. Bozzoni, G. Porcini and A. Pasquinucci (1996). Hydrogen Fuel Cells in
Chemical Industry: The Assemini Project. 1996 Fuel Cell Seminar, 747-749. Courtesy
Associates, Orlando, Florida, US.
CBS (1992). Energiestatistiek Nederland 1990. Voorburg.
CBS (2002). Energiestatistieken Nederland. www.cbs.nl
References 229
Dijkema, G.P.J. and J. Grievink (2000). Technological change by system design - the
industrial production of aromatics. European Symposium On Computer Aided Process
Engineering, Escape-10, 631-636. Elsevier Science, Florence, Italy.
Dijkema, G.P.J., J. Grievink and M.P.C. Weijnen (2001). Process System Innovation for a
Sustainable Petrochemical Industry 6th World Congress of Chemical Engineering, Melbourne,
Australia.
Dijkema, G.P.J., J. Grievink and M.P.C. Weijnen (2003). "Functional Modelling for a
Sustainable Petrochemical Industry." Transactions of the Institution of Chemical Engineers Part
B: Process Safety and Environmental Protection 81(5): 331-340.
Dijkema, G.P.J. and B. Kuipers (2001). De chemie van Mainport Rotterdam - Middellange
termijn ontwikkeling van de chemie in de Rotterdamse haven: analyse, beleid en
stimulering. TU Delft, Faculteit Techniek, Bestuur en Management. Delft.
Dijkema, G.P.J., C.P. Luteijn and M.P.C. Weijnen (1998). "Design of trigeneration systems -
Process integrated applications of energy conversion devices in chemical plants." Chemical
Engineering Communications 168: 111-125.
Dijkema, G.P.J. and I.S. Mayer (2001a). Corporate Strategy and Product-Oriented
Environmental Policy. 6th World Congress of Chemical Engineering, Melbourne, Australia.
Dijkema, G.P.J. and I.S. Mayer (2001b). Information Exchange on Sustainable Product
Chains in the Dutch Chemical Industry? 3rd European Congress on Chemical Engineering,
Nuernberg, Germany.
Dijkema, G.P.J. and M.A. Reuter (1999). "Dealing with complexity in material cycle
simulation and design." Computers & Chemical Engineering 23: S795-S798.
Dijkema, G.P.J., M.A. Reuter and E.V. Verhoef (2000). "A new paradigm for waste
management." Waste Management 20(8): 633-638.
Dijkema, G.P.J., J.N. Schinkel and M.P.C. Weijnen (1994a). "Doelmatig gebruik van
grondstoffen in de procesindustrie." NPT-Procestechnologie 2(5).
Dijkema, G.P.J., J.N. Schinkel and M.P.C. Weijnen (1995). Synthese, Technologische
mogelijkheden voor een doelmatig beheer van primaire en secundaire grond stoffen, dl.
1. Interduct. Delft.
Dijkema, G.P.J., E.M. Steenkamp and P.J.T. Verheijen (1997). "Implementation of novel
processes in existing process networks: Evaluation of the industrial production of C-4
chemicals by optimization." Computers & Chemical Engineering 21: S493-S498.
Dijkema, G.P.J. and R.M. Stikkelman (1999). A Decision-Support System for Eco-Industrial
Parks - Development of an Agro-Industrial Complex -. Int. Conf. on Industrial Ecology,
Troyes, France.
Dijkema, G.P.J. and L. Stougie (1994). Environmental and economic analysis of the Veba
Oel option for the processing of mixed plastic waste. Interduct. Delft.
Dijkema, G.P.J., A.F. Veenstra, M. Vriesendorp and H.C. de Lathouder (1994b). Design of a
MCFC-upgrade kit for a steam-reforming methanol plant. 1994 International Fuel Cell
Seminar, 299-302. Courtesy Associates Inc., San Diego.
Dijkema, G.P.J., J. Vervoort, R.J.E. Daniëls and C.P. Luteijn (1996). Ammonia synthesis and
er-mcfc-technology - a profitable combination? 1996 Fuel Cell Seminar, 303-307. Courtesy
Associates Inc., Orlando, Florida.
Douglas, J.M. (1988). Conceptual Design of Chemical Processes. McGraw-Hill, New York.
Dow Chemical Benelux (1994). Schone cumeenproductie door procesontwikkeling. DSM
Milieutechnologie prijsuitreiking. TROS Aktua in Bedrijf.
References 231
Glasser, D. and D. Hildebrandt (1997). "Reactor and process synthesis." Computers &
Chemical Engineering 21: S775-S783.
Goossens, J.C.J.M. (1995). Conceptual Design of a Fuel Cell System for an Electrothermal Phosphorus
Process. Faculty of Chemical Engineering and Material Science. Delft.
Govaerts, F. (1999). Largest air-separator in the World. Presentation. AIChE-NLB company
visit to Air Products, Rotterdam-Botlek,
Graedel, T.E. (2003). Cross-scale Analysis of the Contemporary Copper and Zinc Cycles.
ISIE '03: International Society for Industrial Ecology Second International Conference, 1.
International Society for Industrial Ecology, Ann Arbor, Michigan MI, USA.
Graedel, T.E. and R.B. Allenby (1995). Industrial Ecology. Prentice Hall, Englewood Cliffs,
New Jersey.
Greer, D., R. Houdek, R. Pittman and J. Woodcock (2001). Maximizing Propylene Yields in
Fluid Catalytic Cracking and Related Technologies - Solutions for Increasing Propylene
Production. DGMK, Hamburg, Germany.
Grievink, J. (1994). Ontwerpen voor Procesvoering, Schakel tussen Molekuul en Markt. TU
Delft.
Grievink, J. (1998). Discussion on Symbolic System Representation.
Grievink, J. (2003). Symbolic System Representation.
Grievink, J. and J.v. Schijndel, Eds. (2002). European Symposium on Computer Aided Process
Engineering 12. Elsevier, Amsterdam.
Grigg, J.B., Ed. (1974). The Agricultural Systems of the World - an Evolutionary Approach.
Cambridge University Press.
Grossmann, I.E., J.A. Caballero and H. Yeomans (2000). "Advances in mathematical
programming for the synthesis of process systems." Latin American Applied Research 30(4):
263-284.
Grossmann, I.E. and A.W. Westerberg (2000). "Research Challenges in Process System
Engineering." AIChE Journal 46(9): 1700-1703.
Grubbard (1992). Energie & Milieuspectrum.
Grunert, K.G., H. Harmsen, M. Meulenberg, E. Kuiper, T. Ottowitz, F. Declerck, B. Traill
and G. Goransson (1997). A Framework for Analysing Innovation in the Food Sector.
Product and Process Innovation in the Food Industry. B. Traill and K. G. Grunert, Eds.: 1-33.
Blackie Academic and Professional, London.
Hamel, G. and C.K. Pralahad (1994). Competing for the future. Harvard Business School Press,
Boston.
Hampden-Smith, M. (2000). The Design & Performance of Ultra Low Pt Electrocatalysts
produced by spray based routes for fuel cell applications,. paper presented by Superior
MicroPowders, LLC, Albuquerque, 2000.
Hartmann, K. and K. Kaplick (1990). Analysis and Synthesis of Chemical Process Systems. Elsevier,
Amsterdam.
Heijnen, P.W. and Z. Verwater-Lukszo (2003). Batch Plant Integration in Need of more
Support. Foundations of Computer Aided Process Operation FOCAPO-2003, Florida.
Hemmes, K. and G.P.J. Dijkema (1998). Method of operating a molten carbonate fuel cell, a
fuel cell, a fuel cell stack and an apparatus provided therewith. TU Delft, NL-2628 BL
Delft.
Hemmes, K., W.H.A. Peelen and G.P.J. Dijkema (1998). New MCFC with separate CO2
supply, Proof of Principle. 1998 Fuel Cell Seminar.
References 233
Roussel, P.A., K.N. Saad and T.J. Erickson (1991). Third generation R and D; managing the link
to corporate strategy. Harvard Business School Press, Boston.
Rowe, D.A., J.D. Perkins and S.P. Walsh (1997). "Integrated design of responsive chemical
manufacturing facilities." Computers & Chemical Engineering 21: S101-S106.
Rudd, D.F., S. Fathi-Afshar, A.A. Trevino and M.A. Stadtherr (1981). Petrochemical Technology
Assessment. Wiley, New York.
Rudd, D.F., G.J. Powers and J.J. Siirola (1973). Process Synthesis. Prentice-Hall, Englewood
Cliffs.
Rudd, D.F. and C.C. Watson (1968). Strategy of Process Engineering. John Wiley & Sons. Inc.,
New York.
Ryu, J.H., V. Dua and E.N. Pistikopoulos (2004). "A bilevel programming framework for
enterprise-wide process networks under uncertainty." Computers & Chemical Engineering
28(6-7): 1121-1129.
Sargent, R.W.H. (1998). "A functional approach to process synthesis and its application to
distillation systems." Computers & Chemical Engineering 22(1-2): 31-45.
Sas, H., J.P. van Soest, M.J. Zegwaard and G.P.J. Dijkema (1994a). Disposal of plastic household
waste: analysis of environmental impacts and costs. Centrum voor Energiebesparing en Schone
Technologie. Delft, The Netherlands.
Sas, H., J.P. van Soest, M.J. Zegwaard and G.P.J. Dijkema (1994b). Verwijdering van
huishoudelijk kunststofafval: analyse van milieu-effecten en kosten, hoofdrapporten en
bijlagen. CE. Delft.
Satterfield, C.N. (1991). Heterogeneous catalysis in industrial practice. McGraw-Hill, 2nd New
York.
Schinkel, J.N., G.P.J. Dijkema and M.P.C. Weijnen (1994). Novel opportunities for fuel cell
commercialization: process integrated applications,. International Conference 'Bringing Fuel
Cells down to Earth', 239-249. South Coast Air Quality Management District, Los Angeles.
Scholes, G. (1999). Integrating Sustainable Development into the Shell Chemicals Business.
Environmental Performance Indicators in Process Design and Operation. M. P. C. Weijnen and P.
M. Herder, Eds. EFCE Event 616: 69-82. Delft University Press, Delft.
Schumpeter, J. (1942). Capitalism, Socialism and Democracy. Harper and Row, New York.
Seader, J.D. (1982). Thermodynamic Efficiency of Chemical Processes. Massachusetts Institute of
Technology,
Sederquist, R.A., J.C. Trocciola, J.F. Farris and M.J. Smith (1993). Method of Generating High-
purity Nitrogen Gas. US5330857. Int. Fuel Cells Corporation (US).
Seider, W.D., J.D. Seader and D.R. Lewin (1999). Process Design Principles: Synthesis, Analysis,
and Evaluation. Wiley, New York.
Setac (1994). LCA manual oid.
Sharratt, P. (1999). "Environmental criteria in design." Computers & Chemical Engineering
23(10): 1469-1475.
Sheldon, R.A. (1993). Selectiviteit in de Organische Synthese: Schoonheid in de Chemie. TU
Delft, Delft.
Sherwood, T.K. (1963). A Course in Process Design. The MIT Press, Cambridge, MA.
Siirola, J.J. (1996). Industrial Applications of Chemical Process Synthesis. Process Synthesis. J.
L. Anderson, Ed. 23: 2-61. Academic Press, New York.
Simon, H.A. (1969). Science of the Artificial. MIT Press, Cambridge, MA.
Sinclair, J.M., Ed. (1995). Collins Cobuild English Dictionary. HarperCollins Publ., London.
References 237
Smith, J.M., H.C. Van Ness and M.M. Abbott (2001). Introduction to chemical engineering
thermodynamics. McGraw-Hill, 6th New York.
Smith, R. (1995). Chemical Process Design. McGraw-Hill, New York.
Smits, D.F. and G.P.J. Dijkema (1994). Techniekbeschrijvingen, opzet en inhoud. Interduct. Delft :
Interduct.
Spillman, R.W., R.M. Sotnitz and J.T. Lundquist (1984). "Why not make chemicals in fuel
cells." ChemTech 15: 226.
Spiro, T.G. and W.M. Stigliani (1996). Chemistry of the Environment. Prentice-Hall, Upper
Saddle River, NJ.
SRI (1989). Mass Conversion Factors of Industrial Chemical Processes. Stanford Research
Institute, Calif.
Stadtherr, M.A. and D.F. Rudd (1976). "Systems study of the petrochemical industry."
Chemical Engineering Science 31(11): 1019-1028.
Stadtherr, M.A. and D.F. Rudd (1978). "Resource use by the petrochemical industry."
Chemical Engineering Science 33(7): 923-933.
Stanley, S.J. (2001). Recent Advances in the Manufacture of Light Olefins by Steamcracking.
DGMK, Hamburg, Germany.
Steffens, M.A., E.S. Fraga and I.D.L. Bogle (1999). "Multicriteria process synthesis for
generating sustainable and economic bioprocesses." Computers & Chemical Engineering
23(10): 1455-1467.
Stephanopoulos, G. (1984). Chemical Process Control. Prentice-Hall, Englewood Cliffs.
Stikkelman, R.M. (2001). Groene Methanol uit Groningen. Interduct rapport. Delft.
Stobaugh (1988). Innovation and Competition. The global management of petrochemical
products. Harvard Business School Press, Boston.
Stull, D.R., E.F. Westrum and G.C. Sinke (1969). The Chemical Thermodynamics of Organic
Compounds. Wiley, New York.
Südholter, S., M.A. Reuter and J. Kruger (1996). "Eco-techno-economic synthesis of process
routes for the production of zinc using combinatorial optimization." Metallurgical and
Materials Transactions B-Process Metallurgy and Materials Processing Science 27(6): 1031-1044.
Tazelaar, F. (2001). Nieuwe methanolcapaciteit 2001-2006. Memorandum RMPM.
Thompson, R.G., J.A. Colloway, L. Nawalanic and J. Becker (1978). The Cost of Energy and a
Clean Environment Gulf, Houston.
Truxal, J.G. (1972). Introductory System Engineering. McGraw-Hill, New York.
Twiss, B.C. (1992). Managing Technological Innovation. 4th edition, Pitman Publishing, London.
van Breda, L.M. and G.P.J. Dijkema (1998). "Eia's contribution to environmental decision-
making on large chemical plants." Environmental Impact Assessment Review 18(4): 391-410.
van den Hoed, R. (2004). Driving Fuel Cell vehicles. How Established Industries React to
Radical Technologies. Dissertation, Delft University of Technology. Delft.
van der Graaf, M.H.K. (1980). Psychologische Aspekten van de Organisatie. Samsom, Alphen a/d
Rijn, Brussel.
van Geem, P. (1994). "Methanol to Olefins." NPT Procestechnologie.
van Klaveren, R. (2003). BDO and derivatives business Europe. Presentation, AIChE-NLB
company visit to Lyondell Rotterdam-Botlek, Spijkenisse
van Schaik, A. and M.A. Reuter (2004). "The time-varying factors influencing the recycling
rate of products." Resources Conservation and Recycling 40(4): 301-328.
238 Process System Innovation by Design
van Zanten, D.J. and G.P.J. Dijkema (2002). Public roles and private interests in
petrochemical clusters - Model-based decision support of a Regional Development
Agency. Internal Research Paper, TU Delft, Fac. TBM. Delft.
Vanderborght, B. and G.P.J. Dijkema (1996). Perspectives for the use of combustible waste in the
Belgian and French cement industries. School of Systems Engineering, Policy Analysis and
Management / Interduct, Delft University of Technology.
Vayenas, C.G. (1988). "Catalytic and electrocatalytic reactions in solid oxide fuel cells." Solid
State Ionics 28-30: 1521-1539.
Vayenas, C.G., S. Bebelis, I.V. Yentekakis and H.G. Lintz (1992). "Non-Faradaic
Electrochemical Modification of Catalytic Activity - a Status-Report." Catalysis Today
11(3): R5-438.
Vayenas, C.G., S.I. Bebelis and C.C. Kyrizais (1991a). ChemTech 21: 50.
Vayenas, C.G., S.I. Bebelis and C.C. Kyrizais (1991b). "Cogeneration - Electricity +
Chemicals .1. Solid Electrolytes." ChemTech 21(7): 422-428.
Verhoef, E.V., G.P.J. Dijkema and M.A. Reuter (2001). Managing the Dynamics of Waste
Management. Proc. 3rd Annual Symposium Delft Interfacultary Research Center Design and
Management of Infrastructures 125-152. Delft University Press.
Verhoef, E.V., G.P.J. Dijkema and M.A. Reuter (2002a). Waste co-incineration in the
European Community - is the sky the limit? Int. Conf. on Resources, Recovery and Recycling
RRR2000, 1028-1036. Toronto, Canada.
Verhoef, E.V., G.P.J. Dijkema and M.A. Reuter (2004a). "Knowledge, waste management
and metal ecology." Journal of Industrial Ecology 8(1-2).
Verhoef, E.V., J.A. van Houwelingen, G.P.J. Dijkema and M.A. Reuter (2002b). Industrial
Ecology and Flexible Waste Infrastructure. 5th Int. Conf. on Technology, Policy and Innovation,
The Hague, Delft2001 Critical Infrastructures, 238 and CD-ROM. Lemma, Utrecht, The
Netherlands.
Verhoef, E.V., J.A. van Houwelingen, G.P.J. Dijkema and M.A. Reuter (2004b). "Industrial
Ecology for Waste Infrastructure Development. A roadmap for the Dutch Waste
Management System." Technological Forecasting and Social Change In press.
Villermaux, J. (1996). Opening Address. 5th World Congress of Chemical Engineering, AiChE, San
Diego, California.
VNCI (1989). Integrated Substance Management. VNCI. Leidschendam.
VNCI (1998). Jaarverslag. Leidschendam.
Vogel, F. (2001).
von Bertalanffy, L. (1951). "Problems of General Systems Theory." Human Biology.
von Weiszacker, E.U. (1996). Factor Four: Doubling Wealth, Halving Resource Use - A
Report to the Club of Rome. Earthscan, London.
VROM (1972). UrgentieNota Milieuhygiene. Tweede Kamer 1971-1972, 11906, Den Haag
Weber, A. (1909). Über den Standort der Industrien DII. Tübingen.
Wei, J., T.W.F. Russell and M.W. Swartzlander (1979). The Structure of the Chemical Processing
Industries, Function and Economics. McGraw-Hill, Inc., New York.
Weijnen, M.P.C. (1991). Naar een doelmatiger beheer van primaire en secundaire
grondstoffen. Projectvoorstel, Interduct. Delft.
Weijnen, M.P.C. (1999). The Greening of Natural Gas. Research Proposal, Faculty of
Technology, Policy and Management. Delft.
References 239
Weijnen, M.P.C. and A.A.H. Drinkenburg, Eds. (1993). Precision Process Technology, Perspectives
for pollution prevention. Kluwer Academic Publishers, Dordrecht.
Weissermel, K. and H.-J. Arpe (1994). Industrielle organische Chemie; bedeutende Vor- und
Zwischenprodukte. VCH, 4. Aufl. Weinheim.
Welford, R. (1997). Hijacking Environmentalism: Corporate Responses to Sustainable
Development. Earthscan, London.
Westerberg, A.W. (1989). "Synthesis in Engineering Design." Computers & Chemical
Engineering 13(4-5): 365-376.
Westerberg, A.W. (2004). Comment on the design course thought by Prof. T.K. Sherwood.
Westerberg, A.W. and O.M. Wahnschafft (1996). Synthesis of Distillation-Based Separation
Systems. Process Synthesis. J. L. Anderson, Ed. 23: 64-170. Academic Press, New York.
Wetenschappelijke Raad voor het Regeringsbeleid (2003). Naar nieuwe wegen in het milieubeleid.
Sdu Uitgevers, Den Haag.
Whitehead (2000). "Looking from the outside in: Achieving value growth in asset-intensive
industries." Chimica Oggi 18(5).
Witlox, F.J.A. (1998). Modelling Site Selection : a Relation Matching Approach Based on
Fuzzy Decision Tables. Dissertation, TU Eindhoven, Faculteit Bouwkunde. Eindhoven.
Woltjer, G. (2002). De economische manier van denken. 2e druk, Coutinho, Bussum.
Wood, D., Ed. (2001). Proceedings of the 6th World Congress on Chemical Engineering,
Melbourne, Australia.
World Commission on Economic Development (1987). Our Common Future. Oxford
University Press, Oxford, UK.
Worrell, E. (1994). Potentials for improved use of industrial energy and materials.
Dissertation, Utrecht University. Utrecht.
WRI (2003). 2004: Decisions for the Earth: Balance, voice and power. World Resources 2002-2004.
World Resources Institute. governance.wri.org.
Yang, Y. and L. Shi (2000). "Integrating environmental impact minimization into conceptual
chemical process design -- a process systems engineering review." Computers & Chemical
Engineering 24(2-7): 1409-1419.
Yee, T.F. and I.E. Grossmann (1990). "Simultaneous-Optimization Models for Heat
Integration .2. Heat-Exchanger Network Synthesis." Computers & Chemical Engineering
14(10): 1165-1184.
Yee, T.F., I.E. Grossmann and Z. Kravanja (1990). "Simultaneous-Optimization Models for
Heat Integration .1. Area and Energy Targeting and Modeling of Multi-Stream
Exchangers." Computers & Chemical Engineering 14(10): 1151-1164.
Yentekakis, I.V. and C.G. Vayenas (1989). "Chemical Cogeneration in Solid Electrolyte Cells
- the Oxidation of H2s to So2." Journal of the Electrochemical Society 136(4): 996-1002.
Zeppenfeld, R. and R. Walzl (2001). Steamcracker Revamp Projects: Challenges and
Technologies. DGMK, Hamburg, Germany.
Zevnik, F.C. and R.L. Buchanan (1963). Chem. Eng. Progr. 59(2): 70.
Symbols
Chapter 1
I total plant Investment
C plant capacity
p parameter
Suffix denotes:
0 standard or reference condition
Chapter 3
A Availability [J]
B By-product output ( Stream label )
B behaviour of a system
D Domain
∆ Difference or change (∆G, ∆H, ∆U, ∆S)
E Emission output ( Stream label )
Ex Exergy [J]
ηa Reference efficiency
ηb Product efficiency
ηc Emission efficiency
ηd Total emission efficiency
H Enthalpy [ J/mole ]
I Inputs ( Stream label )
Λ system boundary
Σ sum
Σ system
σ set of streams
σ Entropy created [ J/K ]
S Scarce resources input ( Stream label )
S Entropy [ J/K ]
Su set of inputs
Sy set of outputs
S1 set of objective-defined functions
S2 set of objective-defined functions
T Technology
T Ubiquities input ( Stream label )
T Temperature [K]
t time [ time unit ]
κu,y(t) vectors of conversion factors [ unit/flow unit ]
κ• net accumulation of chemical element, mass or energy [ unit ]
La Reference Loss [ unit ]
Lb Product Loss [ unit ]
Lc Emission Loss [ unit ]
Ld Total Emission Loss [ unit ]
Le Balance Loss [ unit ]
M Main Product output ( Stream label )
242 Process System Innovation by Design
Suffix denotes:
0 ‘at standard conditions’
a allowed ( Stream label )
c co- (feed or product) ( Stream label )
d desired ( Stream label )
f primary ( Stream label )
i incidental ( Stream label )
irr. irreversible
l limited ( Stream label )
m material
n accepted ( Stream label )
p process-related ( Stream label )
r reference ( Stream label )
r reaction
u utility-related ( Stream label )
w wasted ( Stream label )
Chapter 5
Ac matrix of input-output coefficients
α electric power per ton unit capacity [ kW/(ton/day) ]
β conversion factor 1,16 *103 [ J/Cal]•[g/ton]•[kJ/kWh ]
c vector of capacities
d vector of demands
f vector of some objective function’s coefficients
∆G0 Gibb’s free energy change at standard conditions [ kCal/mole ]
∆Gr Gibb’s free energy change of reaction [ kcal/mol ]
ε electrical efficiency [ 0..1 ]
γ fraction of fuel cell system investment [ γ ≤ 1 ]
I Unity matrix
I Investment, expressed as dollars per unit capacity [ $/(ton/day) ]
Ie Investment, expressed as dollars per unit power generated [ $/kW ]
λth ratio of the electricity revenue and value added of the chemical
Symbols 243
Suffix denotes:
c ‘conventional reactor’ or ‘conventional plant’
e ‘expression in electric dimensions’ (e.g. Ie)
f ‘fuel cell reactor’ or ‘fuel cell use in plant’
s ‘stand-alone power generation’
t ‘trigeneration’:
Abbreviations
ABS Acrylonitrile Butadiene Styrene copolymer
AC Alternating Current
AIChE American Institute of Chemical Engineers
Akzo Akzo Nobel (company, NL)
Al Aluminium
AVR Afvalverwerking Rijnmond (company, NL)
B Benzene
BASF Badische Anilin und Soda Fabrik (company, Germany)
BDO Butanediol
BMW Bayerische Motorwerke AG (company, Germany)
BTU British Thermal Unit
BTX Benzene, Toluene and Xylenes
BTXCon Consumption system BTX
BTXExtr. BTX Extraction process
BTXProd Production system BTX
BTXSep BTX separation
BTXTot Total system BTX
C Carbon
C1 Hydrocarbons with only one C-atom
C2 Hydrocarbons with two C-atoms (e.g. ethane, ethylene)
C3 Hydrocarbons with three C-atoms (e.g. propane, propylene)
C4+ Hydrocarbons with four or more C-atoms
C5+ Hydrocarbons with five or more C-atoms
CaCO3 Lime or Calcium Carbonate
CAPE Computer Aided Process Engineering
CBR Case-Based Reasoning
CBS Centraal Bureau voor de Statistiek (organisation, NL)
CE Centrum voor Energiebesparing (company, NL)
CEO Chief Executive Officer
CFC Chlorinated Fluoro Carbons
CH4 Methane (natural gas)
Cl2 Chlorine
CMAI Chemical Marketing Associates Inc. (company, US)
CO Carbon monoxide
CO2 Carbon dioxide
CPC Computer and Process Control
Cu Copper
DC Direct Current
DHV The Dutch company DHV
DME Dimethylether
DMT Dimethyl Terephtalate
DSM The Dutch company DSM
DTO Duurzame Technologie Ontwikkeling
EB Ethylbenzene
ECN Energie Centrum Nederland (research centre, NL)
EDC Ethylene Dichloride
EOP Electrochemical Oxygen Pumping
EPDM Ethylene Propylene Dimonomer
246 Process System Innovation by Design
MTBE Methyl-tert-butyl-ether
MW Megawatt
MWh Megawatt hour
mX meta-Xylene
NaCl Sodium Chloride (rock salt)
NGL Natural Gas Liquids
NGO Non Governmental Organisation
NH3 Ammonia
NIMBY Not In My Backyard
NL The Netherlands
NOx Nitrogen oxides
NRC National Research Council (organisation, US)
ODF Objective-defined Function
OECD Organisation for Economic Co-operation and Development
OPEC Organisation of the Oil Exporting Countries
oX ortho-Xylene
oXylSep ortho-Xylene separation
PAFC Phosphoric Acid Fuel Cell
pc Personal Communication
pCO2 Partial CO2 pressure
PdVC Polydivinylchloride
PE Polyethylene
PEM Polymer Electrolyte Membrane
PEMFC Polymer Electrolyte Membrane fuel cell
PET Poly Ethylene Terephtalate
PG Propylene Glycol
Phenol1 Phenol process 1
Phenol2 Phenol process 2
PJ Pèta Joule (1015 Joule)
PMMA Polymethylmethacrylate
Pmt Product-market-technology combination
Pmt-R Product-market-technology-Resource combination
PO Propylene Oxide
PoR Port of Rotterdam
POX Partial Oxidation
PP Polypropylene
PS Polystyrene
PSA Pressure Swing Adsorption
PSE Process System Engineering
Pt Platinum
PtcAnh Phtalic Anhydride
PTFE Poly tetra fluoro ethylene
PUR Polyurethane resin
PVC Poly vinylchloride
pX para-Xylene
PyGsl Pyrolysis Gasoline production
R&D Research and Development
R/P Resource to Production ratio
RC Responsible CareTM
RefGsl Refinery Gasoline production
RIVM Rijksinstituut voor Volksgezondheid en Milieu (research centre, NL)
ROACE Return on Asset Capital Employed
248 Process System Innovation by Design
A. 2 Applied thermodynamics
State properties
In thermodynamics, the concept of state properties has proven to be very useful. A
state property of a particular substance, mixture or stream only depends on the
particular state (pressure, temperature, composition), and not on how that state has
been realised. The value of a state property is always quoted relative to some
reference state60.
Important state properties are internal energy U, enthalpy H, free energy G (all
[J/mol] and entropy S [J/mol/K]. These have been defined in chapter 3:
∆U = q + w (Eq. 3.2)
The internal energy of a system changes in case heat q or work w flow in or out of
the system.
Properties correctly derived from combinations of state properties also are state
properties. Enthalpy is a state property defined as
∆H = ∆U + ∆(PV). (Eq. 3.3.)
It follows from eqn. 3.2 and 3.3 that in case only PV work is allowed on a system
(i.e. no shaft Work, w=0) that ∆H = qp.
60 For thermodynamic properties such as enthalpy, this is the state of the elements and standard
conditions (P=1 atm, T=298 K). For industrial properties such as the Lower Heating Value and
exergy, one prefers to use the lowest energy state found on Earth (i.e. CO2 + water). The definition
of a reference state for other elements is somewhat arbitrary, as their natural appearance may vary.
As a consequence, discussions have been ongoing in the scientific community to arrive at a consensus
that can be used for consistent energy and exergy efficiency calaculations (e.g. Kotas 1985).
Appendices 253
Manifestations of energy
Energy has many forms or manifestations. The best known manifestation of energy
is heat, in the form of radiation, convection, or direct heat transfer. Other energy
forms are movement (kinetic energy), gravitational potential energy, and of course
electrical potential energy.
254 Process System Innovation by Design
Powering
Coal Equipment
Powering
Equipment
7
Propulsion
8
Heat Industrial
Tmedium Processing
3 9
Heat Space-
Tlow heating
4 10
Energy sources
or
stocks Forms of energy or energy-carriers Energy use
to be used in central heating systems, for example, misuse of LHV leads to common
'deceit'. Boiler efficiencies greater than 100% are proclaimed for systems that utilise
the energy released during condensing the water vapour present in the flue gas. The
energy content of the natural gas consumed is listed as the Lower Heating Value, a
measure which assumes that water vapour is not condensed. Thus, both exergy and
LHV are thermodynamically sound properties once the 'dead' reference state has
been defined, which must include the phase state of the components included.
The work of Denbigh (1956) provides a link between energy systems and chemical
reaction systems as he derived the equation
wt = To σ + ∆A
This equation states that the total Work wt required to let a process progress at a
finite speed equals the wasted work To σ or irreversible entropy production plus the
change in of availability the process materials ∆B. In case the transformation occurs
and T0, this equals ∆G0
256 Process System Innovation by Design
61 Only if the initial and final (P,T) of the materials processed is the same, and equals (P,T) of the
Base Metal
Coke
Inorganic Chemicals Food
factories
industry
Fertilizers
Large-scale process industry
Other Organic
Chemicals
Power Plants
Polymers
Petroleum Refineries
Petrochemicals
62 Methanol To Olefins
63 Shell Middle Distillate Synthesis
Appendices 259
64 This appendix is largely based on internal discussions and research note (Grievink 2003).
260 Process System Innovation by Design
Now that the behaviour of the system has been defined, the next item is the measures
for the performance of the system.
The performance will be measured by so-called values ω ∈ Ω. These values are
elements of the performance space Ω. Again this space is given an internal structure
to make sure that the elements have a size and that a distance measure is in place.
The performance values are obtained from a performance functions (P) , which is a
mapping from the inputs, outputs and, possibly, the states over a certain time
interval of evolution to the performance measures:
Pstate: SU * SX * SY * T => Ω
The mapping P to the performance space is called the value map.
The values can be written as a functional over time interval [t0,t1]:
ω = P(∫t0t1 p(u(t),y(t),x(t),t) dt
Some examples of performance values are: controllability, controllability and
economic profit, technological indicators, like exergy losses. Hence, the system Σstate,
is now fully characterised by:
Σstate = { SU , SX , SY , T, Ω, Bstate, Pstate }
An example of such a system description will now specified for linear dynamic
systems.
The behaviour, Blin, can be characterised by a triplet of matrices:
Blin = {B, A, C}.
The corresponding behaviour is given by:
d x(t) /dt = A.x(t) + B.u(t) and y(t) = C. x(t)
The performance of such linear systems is often specified in a quadratic fashion:
ω(t1 ,t0 ) = x(t0)T.W0.x(t0) + x(t1)T.W1.x(t1)
+ ∫tot1[u(t)T.Wu.u(t) + y(t)T.Wy.y(t)] dt
Hence, the performance functional can now be characterised by the matrix quartet:
Pquadratic = {W0, W1, Wu, Wy }
For a black-box type of input – output description, where the internal state is not
considered, the behavioural description will reduce to:
BI-O: SU * T => SY * T
When evolution in time is not considered (e.g. representing the steady state of a
system) the description further simplifies to:
BI-O,ss: SU => SY ,
PI-O,ss: SU * SY => Ω
ΣI-O, ss = { SU , SY , Ω, BI-O,ss, PI-O,ss }.
Table A.6.3: Capacities of propylene derivatives in the Dutch industrial olefins system
Product Capacity [ton/year] Company Remarks
Propylene 1.220.000 Dow, DSM, Shell
Polypropylene 480.000 DSM, Basf, Shell Expansion: 190.000
Propylene oxide 370.000 Shell, Arco
Propylene glycol 60.000 Arco
Propylene glycol ether 100.000 Arco
Polyols 180.000 Dow, Du Pont, Estimate
ICI, Shell
Acrylonitrile 175.000 DSM
acrylamide 15.000 Cytec Estimate
polyacrylamide 15.000 Cytec Estimate
isopropylalcohol 290.000 Shell
diisopropylether ? Shell
Acetone 150.000 Shell
PMMA 9.500 ICI
Methylisobutyl ketone 60.000 Shell Estimate
Methylisobutyl carbinol
Allylchloride 80.000 Shell
Epichloro hydrin 80.000 Shell
Bisphenol A 90.000 Shell
Epoxy resins 55.000 Shell
65 Data and information were supplied with the understanding that these would be treated as
confidential.. Therefore, analysis results are only presented in aggregate form. Details on the data set
used and the reconciliation process applied are available from the author.
Appendices 265
The data sets obtained were converted prior to their use in the assessment. The data
reconciliation involved a check for data-consistency and completeness per data set
on an individual plant. Intricate process knowledge and mass and energy balancing
were used to correct or complete data sets. To be able to complete the process the
following assumptions had to be made:
- with respect to additions for completion of plant's mass balances, is was
assumed that a deficiency (i.e. more mass going into the process than going out)
is caused by the generation of light gases that are emitted as CO2 and H2O. This
assumption can be considered valid, since most process schemes include an off-
gas treatment section that remains otherwise unspecified, whilst 'hydrocarbons'
can always be utilised as fuel after proper treatment. The carbon and hydrogen
thus lost will eventually end up as CO2 and H2O.
- the specific energy content of streams not reported in a particular data set can be
estimated. However, since often nothing is known about the energy content or
the application of these streams, these are assumed to be zero in the analysis.
Mass loss La, Lb and mass efficiency ηa, ηb do not require complete datasets, they
immediately give an impression of the performance of the process. Calculation of
these indicators does not require an assumption with respect to the fate of the
external mass loss!
Computation of Industrial System Performance indicators
The energy content and exergy content of the input and output streams can
conveniently be estimated from thermodynamic data reported in the literature (Stull
et al. 1969). The computation of Lower Heating Value and exergy require adherence
to a suitable thermodynamic reference state (P0, T0, standard composition of the
environment). This is acceptable for the analysis performed, which is intended to
obtain an overall picture in a reasonable amount of time, which is reasonably
accurate and which can serve to make a first selection of starting points for an
assessment of the (technological) potential for improvement. Because most
production plants take-in feedstock at ambient temperature. In specific cases,
however, the actual losses may differ substantially from the ones calculated in the
analysis.
As with mass balances, however, hardly ever the data reported on industrial
processes result in a closed energy balance because streams are missing. Energy
leaving an industrial process, for example, as flue gas, cooling water, or condensate
are hardly ever listed. In addition, some losses occur by radiation. Properly defined,
energy is a thermodynamic property of state. Energy content, however, also
represents a valuation of streams. In chapter 3, it was argued it is a prime candidate
for use in process system assessment for sustainability.
266 Process System Innovation by Design
Mass
Efficiency Loss
0…100% [kton/year]
a b e a b e
System Subsystem Process name Capacity
Production Ethylene Steam cracker 73 73 44 2550 350 350 1211
Production Propylene Steam cracker 73 73 44 1220 350 350 1211
Production C4/C5/PyGsl/Steam cracker 0 0 0 1043.1 0 0 0
Consumption Ethylene EDC+VCM 85 85 47 480 88 88 549
Consumption Ethylene Ethylene diamine 40 40 23 49 74 74 164
Consumption Ethylene Ethylene oxide 54 54 11 285 242 242 2255
Consumption Ethylene Ethylene glycol 81 92 92 345 79 32 32
Consumption Ethylene Glycol ethers 81 81 94 74 17 5 5
Consumption Propylene Propylene oxide 91 94 90 370 151 93 166
Consumption Propylene Propylene glycol 81 90 90 60 14 7 7
Consumption Propylene Propylene glycolether 81 90 90 60 24 13 13
Consumption Propylene Polyols 96 97 88 180 8 6 25
Consumption Propylene Acrylonitril 64 72 20 175 97 77 760
Consumption Propylene Acrylamide 75 75 39 15 5 5 23
Consumption Propylene Isopropylalcohol 89 89 57 290 37 37 216
Consumption Propylene DIPA 0 0 0 0 0 0 0
Consumption Propylene Acetone 90 90 64 150 16 16 85
Consumption Propylene MEK 75 75 75 60 10 10 10
Consumption Propylene Allylchloride 49 93 93 80 84 12 12
Consumption Propylene Epichloro hydrin 38 38 2 80 131 131 3400
Consumption Propylene Bisphenol A 86 86 83 90 15 15 18
Consumption Propylene VinylAcetate 0 0 0 60 0 0 0
Polymerisation Ethylene LDPE 97 97 52 580 18 18 536
Polymerisation Ethylene LLDPE 98 98 90 380 9 9 35
Polymerisation Ethylene HDPE 97 97 91 110 3 3 11
Polymerisation Ethylene UHMWPE 97 97 64 5 0 0 3
Polymerisation Ethylene EPDM Rubber 98 98 90 65 1 1 7
Polymerisation Ethylene PE/Vac Copol 98 98 56 20 1 1 19
Polymerisation Ethylene PVC 99 99 27 410 4 4 1099
Polymerisation Propylene Polypropene 96 97 86 480 21 16 77
Polymerisation Propylene PMMA 100 100 8 9.5 0 0 116
Polymerisation Propylene Epoxy resins 68 68 26 55 26 26 159
Appendices 267
Energy
Efficiency Loss
0…100% [PJ/Yr]
a b e a b e
System Subsystem Process name Capacity
Production Ethylene Steam cracker 65 65 65 2550 23.7 23.7 23.7
Production Propylene Steam cracker 65 65 65 1220 23.7 23.7 23.7
Production C4/C5/PyGsl/Steam cracker 0 0 0 0 0.0 0.0 0.0
Consumption Ethylene EDC+VCM 64 64 64 480 5.2 5.2 5.2
Consumption Ethylene Ethylene diamine 39 39 39 49 2.3 2.3 2.3
Consumption Ethylene Ethylene oxide 65 65 64 285 4.2 4.2 4.4
Consumption Ethylene Ethylene glycol 52 61 61 345 5.4 4.4 4.4
Consumption Ethylene Glycol ethers 65 75 75 74 1.0 0.7 0.7
Consumption Propylene Propylene oxide 76 80 80 370 15.5 13.1 13.1
Consumption Propylene Propylene glycol 61 70 70 60 0.8 0.6 0.6
Consumption Propylene Propylene glycolether 69 78 78 60 1.2 0.9 0.9
Consumption Propylene Polyols 78 79 79 180 1.3 1.2 1.2
Consumption Propylene Acrylonitril 51 56 56 175 5.3 4.9 4.8
Consumption Propylene Acrylamide 71 71 68 15 0.1 0.1 0.2
Consumption Propylene Isopropylalcohol 63 63 63 290 5.3 5.3 5.3
Consumption Propylene DIPA 0 0 0 0 0.0 0.0 0.0
Consumption Propylene Acetone 50 50 50 150 3.3 3.3 3.3
Consumption Propylene MEK 97 97 97 60 0.0 0.0 0.0
Consumption Propylene Allylchloride 58 68 68 80 1.4 1.0 1.0
Consumption Propylene Epichloro hydrin 54 54 54 80 1.2 1.2 1.2
Consumption Propylene Bisphenol A 73 73 71 90 1.1 1.1 1.2
Consumption Propylene VinylAcetate 0 0 0 60 0.0 0.0 0.0
Polymerisation Ethylene LDPE 83 83 83 580 5.1 5.1 5.1
Polymerisation Ethylene LLDPE 86 86 84 380 2.3 2.3 2.7
Polymerisation Ethylene HDPE 86 86 84 110 0.8 0.8 0.9
Polymerisation Ethylene UHMWPE 83 83 83 5 0.0 0.0 0.0
Polymerisation Ethylene EPDM Rubber 78 78 78 65 0.8 0.8 0.8
Polymerisation Ethylene PE/Vac Copol 93 93 88 20 0.0 0.0 0.1
Polymerisation Ethylene PVC 76 76 75 410 2.3 2.3 2.4
Polymerisation Propylene Polypropene 82 83 83 480 4.7 4.5 4.5
Polymerisation Propylene PMMA 87 87 86 9.5 0.0 0.0 0.0
Polymerisation Propylene Epoxy resins 88 82 81 55 0.3 0.3 0.3
268 Process System Innovation by Design
Once the composition of a stream is known, the LHV and the exergy value can
easily be computed from thermodynamic basic data and the reference state. In oil
and petrochemicals, however, often only an approximate composition is known and
thermodynamic basic data may be lacking or inaccessible. Therefore a pragmatic
approach has been used: for hydrocarbon mixtures such as 'naphtha', exergy has
been equated to LHV reported in reputable sources66.
In the computation of Industrial System Performance indicators as suggested in
Chapter 3, energy is calculated as Lower Heating Value (LHV) with respect to a
standard reference environment, the so-called dead state (see e.g. Kotas 1985)). The
LHV and exergy of the dead state are zero, i.e. energy is at a minimum, and entropy
is at a maximum. As a result, LHV and exergy of most hydrocarbons greater than
zero, which allows us to define exergetic efficiencies.
The assessment results are given in Tables A.6.4 and A.6.5on the previous pages.
66 Reconciliation was done per individual plant. Details of thermodynamic calculations can be
products, i.e. the main product and some valuable by-products, and the unwanted
products, such as low-quality by-products, emissions, and wastes.
From the summary of main process steps of the Showa-Denko process, we
conclude that propylene, oxygen, chlorine, and calcium hydroxide are feedstock,
whereas water, steam, hydrochloric acid, and acetic acid are processing agents. Water
also is a reactant of the first process step, but the overall process reaction does not
include water.
propylene + ½ oxygen + chlorine + ½ calcium hydroxide -->
ECH + ½ calcium chloride
The main product is ECH, there are no wanted by-products. Propane, waste gas,
waste water, and organic chlorines are all labelled as unwanted by-products.
Table A.6.6: Overall mass balance for a Showa-Denko preliminary process design (from
Bouma et al., 1996)
In [kg/s] Out [kg/s]
Oxygen 1,06 ECH 3,44
Propylene 2,18 Propane 0,69
Water 1,06 Waste gas 1,15
Acetic acid 0,004 Waste water 38,16
(CaCl2) (2,2)
Hydrochloric 0,13 Organic 0,39
acid chlorines
Chlorine 2,91
Aqueous lime 32,1
(Ca(OH)2 (1,6)
Steam 4,37
Total 43,83 Total 43,83
The direct efficiency expresses how fit for purpose the process evaluated is. The
Showa-Denko process design has an ηa of only 0,44. The product efficiency of the
process ηb is equal to ηa, as there are no wanted by-products. Both indicators ηa and
ηb can also be expressed as the quotient of products and all inputs, i.e. feedstock,
and processing agents combined. Only ηc provides additional information, however.
In case of the Showa-Denko process design the total efficiency is only 0,078, which
indicates that the process uses a lot of additional resources that are degraded to
waste. Most of these unwanted by-products can be recovered, however, at the
expense of additional process equipment and energy input.
270 Process System Innovation by Design
Thereby a proxy normal distribution of cohort phase out is modelled around the
average lifespan, which standard deviation increases with average lifespan.
Appendices 273
∇k|i≤j
∇k|i>j
Finally, with respect to health and safety methanol does not have the disadvantage of
high-temperatures and it avoids storage of hydrogen. Compared to gasoline, health
issues appear to be rather similar (Stikkelman, 2001).
Summary
In the petrochemical industry large-scale chemical processes are employed to
convert petroleum-derived feedstock into base chemicals. Huge industrial complexes
of many interconnected processes are the blue-print image of this industry, which is
linked to virtually all important economic sectors such as agriculture, automobiles,
construction, electronics, housing, pharma, textiles, consumer durables and
packaging. The core business of this industry of industries is to transform finite
resources of fossil origin into building blocks of polymers - plastics, resins, fibres.
Petrochemical operations also generate by-products that can be detrimental to our
health, safety and environment if emitted or wasted. In the 1960-70's, the state-of-
the-art industrial practice led to the persistent label ‘Dirty, Damaging, and
Dangerous’ of 'chemical industry'. Over time, however, the control and prevention
of harmful emission and waste management have improved dramatically. Carbon
dioxide, however, continues to be emitted because CO2 formation is linked with
petrochemical process energy supply, feedstock, product and polymer degradation.
To accommodate stable or modest growth in petrochemicals demand, the industry
must timely respond to the prospects of fossil resource depletion and a cap on CO2
emission induced by global-warming concerns. A petrochemical industry is required
that is characterised by appropriate resource selection, effective resource utilisation,
the avoidance of waste and emissions and a product slate that fosters a sustainable
society.
How can these objectives be met? In addressing this question, one must realize that
the petrochemical industry has the structure of an interconnected network. The
nodes represent the processing plants, the vertices are the exchange flows of
intermediate products and energy.
The networked petrochemical industry has an impressive record of innovation of its
industrial plants by invention and improvement of system components - e.g.
catalysts, reactors, separation technology- and products enabled by fundamental and
applied science. Industry maturation, however, has reduced the scope for process
innovation by “inventing new flow sheets for existing problems.” Changes of its
structure caused by R&D breakthroughs or by the use of extra-sector innovations
have become incidental.
The system structure of the industry generally has been considered fixed and the
effect of most R&D on industry network structure is not planned. A literature
review revealed that conceptual design of the industry structure and the specification
of R&D themes for multiscale systemic industry change is a 'white spot' on the
science and engineering map. The discipline of engineering science that comes
closest to dealing with these issues is process system engineering. PSE has evolved
as a sub discipline of chemical engineering that “is concerned with the understanding
and development of systematic procedures for the design and operation of chemical
process systems, ranging from microsystems to industrial scale, continuous and
batch processes.” In PSE multiscale process synthesis approaches that “span a
chemical supply chain -from mastering the molecules to chemical enterprises-” only
have been suggested, however.
280 Process System Innovation by Design
Starting from a higher level of aggregation , economy and policy studies could offer
options, but these suffer from technological lack-of-content. We conjectured
therefore a sustainable petrochemical industry requires process system innovation,
which is defined as “change in the system structure or system design of the
petrochemical industry, its industrial complexes, or individual plants. These can be
enabled by technological inventions or vice versa.” In technology and systems
development and in the adoption of extra-sector innovations the industry structure
explicitly must be considered a degree-of-freedom. Therefore, the central theme of
this thesis is the specification of system innovation content for a sustainable
petrochemical industry.
Systemic innovations do require a long time for development and implementation.
Early identification of promising concepts thus is preferred to facilitate sustainable
development at affordable and acceptable costs and to maximise the effect of R&D.
Therefore, the research questions addressed in this dissertation were (1) what is a
suitable method to assess the resource utilisation and the scope for improvement of
the production systems in the petrochemical industry, (2) can the search of the
innovation space be structured to foster the specification of process system
innovation content and (3) what is the usefulness and the scope of application of the
methods thus developed?
These research questions have been addressed using system theory,
thermodynamics, PSE and computer systems modelling. The concept of and the
methodology for 'process system innovation by design' was developed, validated and
tested in case studies that each span multiple system aggregate levels.
In a technology-free systems representation using the input/output paradigm, any
system is considered a ‘black-box’, which upon opening may consist of a structured
collection of system elements or sub-systems. At each system aggregate level it is
assumed that the function of a system is to generate a behaviour to transform a set
of inputs into a set of outputs in order to satisfy an objective. Together, these relate
the system to its external world. Its performance is assessed by means of criteria or
value functions. The petrochemical industry is a system embedded and linked to the
global material cycles. In a layered networked system view, its system elements are
petrochemical complexes and individual plants. In chemical engineering, these plants
are modelled as structured collections of unit operations connected by flows.
To define indicators for sustainability, mass and energy balancing and stream
valuation on the basis of input scarcity, output fate, economy and ecology were
combined in criteria for system resource utilisation. In quantitative analysis these
indicators yield an image of performance. Together with known plant capacities,
such quantitative system assessment results in a ranking of petrochemical plants with
respect to performance and scope for improvement. This ranking is a starting point
for systematic search for R&D options and themes.
An early revelation of our research was that inclusion of fuel cells in chemical plants
would have a dramatic effect on plant performance by exploiting hitherto
unrecognised fuel cell functionality, which allows efficient interconversion of
chemical, electric and thermal energy. The functional description of the fuel cell as a
unit operation and the subsequent recognition petrochemical plants throughout the
Summary 281
industry as candidates for fuel cell integration, however, requires some suitable
method for system decomposition.
Functional modelling for computer system design, was related to and adapted for the
technology-free petrochemical system representation using the input-output
paradigm. It enables cohesive system decomposition using 'mutual contingency' as a
stop-criterion, which indicates the system decomposition exhibits loose coupling and
sufficient relation between the system elements. The criterion is formalised in a
logical algorithm. In the multiscale modelling exercise subjective functions describe
the system and parts thereof. Upon completion of a mutual contingent
decomposition the functional model is built.
The modelling of system functions rather than system technology provides a
mechanism for abstraction from current design and physical realisation of industrial
systems. Stream valuation and functional modelling were the missing links between
resource utilisation assessment and the systemic exploration of the multiscale
innovation space for chemical conversion and energy interconversion. In system
synthesis, these links enable the specification of process system innovation content -
or the formulation of primitive problems under the General Design Paradigm. The
main result is a methodology for said specification that is based on system
representation using the input-output paradigm and functional modelling.
The use of fuel cells in the petrochemical industry leads to trigeneration systems that
produce useful chemical products, electricity, and heat. Functional modelling helps
to specify new trigenerate system concepts at the aggregate level of reactors,
petrochemical plants and complexes, one of which was patented. Addressing
innovation for the industrial olefins system showed that ongoing R&D and
innovations reported can be categorised and augmented. Quantitative assessment
using stream valuation allows one to focus on 'the weakest link'. Functional
modelling enabled specification of a true process system innovation, the integration
of olefins and syngas production from aliphatic hydrocarbons. To exploit a radical
extra-sector innovation, the advent of methanol-fuelled fuel cell vehicles, process
system innovations were specified around the Rotterdam petrochemical cluster,
whereby the Rotterdam process industry becomes a key industrial infrastructure in
methanol, olefins and platinum.
The results of these case studies illustrate that the methodology fosters 'out-of-the-
box' thinking at the aggregate level of unit operations, single plants, industrial
complexes and clusters alike. Since quantitative evaluation of each innovation
specified requires extensive work, some proxy methods for evaluation have been
adopted, such as economic assessment of fuel cell reactor concepts and dynamic
simulation of fuel cell vehicle adoption and its consequences for methanol and
platinum. The innovation results and proxy evaluations do not constitute formal
proof or validation of the methodology. It remains possible that the concepts
specified could have been developed otherwise. Extensive literature research,
however, did not reveal similar structured methods that link creative thinking with
assessment of performance, structure and technology of existing complex systems.
The methodology and concepts presented in this dissertation can be improved and
extended. Qualitative tests that involve experts can be developed to further underpin
the innovations specified. Thereby process system innovation can be linked to
282 Process System Innovation by Design