You are on page 1of 10

www.ingentaselect.com=titles=09575820.

htm

09575820/03/$23.50+0.00 # Institution of Chemical Engineers Trans IChemE, Vol 81, Part B, September 2003

FUNCTIONAL MODELLING FOR A SUSTAINABLE PETROCHEMICAL INDUSTRY


G. P. J. DIJKEMA1 , J. GRIEVINK2 and M. P. C. WEIJNEN3
1,3

Faculty of Technology, Policy and Management, Delft University of Technology, Delft, The Netherlands 2 Faculty of Applied Sciences, Delft University of Technology, Delft, The Netherlands

n a transition towards a sustainable petrochemical industry, its systemic structure and functions must be considered degrees-of-freedom in the search for and the speci cation of development options that reach beyond traditional research and development (R&D) and beyond the optimization of proven technology, proven system concepts and established process networks. Functional modelling allows technology-free speci cation of systems, generalization of system concepts, and recalibration of the use or value of substances in a petrochemical network. It facilitates moving up or down the ladder of system aggregation levels, which allows one to step out of the box to the plant or chemical complex level, or alternatively to connect to the unit-operation level. Application to the aromatics industry serves to illustrate that thereby the abstraction from present system content and from proven technology is achieved, which allows the speci cation of the desired content of technological inventions that enable or rely upon changes in system structure: process system innovations. The realization of trigenerate functions through the petrochemical industry involves a paradigm-shift. Rather than considering the industry a threatening source of pollution, safety and health risks, it is an enabler of global CO2 emission reduction. Keywords: sustainability; petrochemical industry; process synthesis; process system innovation; trigeneration.

INTRODUCTION Over the past decades, the petrochemical industry has reduced its harmful emissions signi cantly, amongst others via environmental management and technological development. Meanwhile, the industry has realized 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 communications and image with local communities and the general public. While the depletion of fossil resources and the threat of global-warming imply that dramatic change is required in industry to achieve sustainability, only a few companies have acknowledged that sustainability is crucial for their long-term prosperity (Dijkema and Mayer, 2001). The Dutch Scienti c Council for Government Policy recently has categorized CO2 emission, climate change, biodiversity and continuously growing energy-use as wicked problems (Wetenschappelijke Raad voor het Regeringsbeleid, 2003), because to date these have largely escaped (inter)national policy and regulation, largely because of lack of public awareness and political consensus. These are dif cult problems to communicate to the general public, because there is no direct exposure to the associated adverse effects that develop gradually over decades in a global context. Meanwhile, many tough environmental problem have been turned into manageable problems in 331

the past 30 years, through end-of-pipe measures, technology development, process improvements, pollution prevention etc. It is, however, the wicked problems that must be addressed to enable sustainable development . . . that meets the needs of the present without compromising the ability of future generations to ful l their own needs (World Commission on Economic Development, 1987). A sustainable petrochemical industry is characterized by appropriate resource selection, effective resource utilization, the avoidance of waste and emissions and a product slate that fosters a sustainable society. Drastic reduction of resource consumption and CO2 emission, however, is a wicked problem for the petrochemical industry, as it is linked to its present use of fossil feedstock and because most companies compete in global markets where opinions on CO2 and global-warming vary. Thus, despite ef ciency gains, the industrys 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. The petrochemical industry has been classi ed a sciencebased industry (Pavitt, 1984). It has an impressive record of technological innovation at or below the level of single plants, where fundamental and applied research have fuelled continuous improvement of, for example, catalysts, reactors, separation technology. Considerable ef ciency improvements have been realized, as innovative design and clever

332

DIJKEMA et al. variety of process systems for, for example, propylene oxide or methanol production therefore are currently in use. The entire petrochemical industry and parts thereof, such as the Rotterdam Chemical Cluster, can be represented as a structured con guration of individual plants that are connected by some kind of ow. These are systems where oil products are converted to suitable feedstock for the polymer industry and other sectors, utilizing natural gas based products where feasible. Analogous to the design of individual plants, the present industry s con guration is only one instance of a very large set of possible system con gurations (Table 1). 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 organized, systematized, or even automated is at the focus of process synthesis research (Siirola, 1996: 3). Referring to industrial practice, however, Siirola hints at the relatively scarcely addressed problem of having to generate owsheets from scratch, as often owsheets from previously realized plants are available. Genuinely novel process system concepts, such as distributed chemical manufacturing (Rowe et al., 1997) are scarcely reported. As of lately, process synthesis techniques have been used for the development of new chemicals (e.g., Marcoulaki and Kokossis, 2000) and the importance of the relation between product and process research has been emphasized (Charpentier, 2002). Despite the research agenda suggested by Grossmann and Westerberg (2000), another largely unexplored white spot in process synthesis R&D is the conceptual design of petrochemical complexes comprising multiple plants, letalone true multiscale approaches that link the complex, single plant and unit operation level. A notable exception beyond the vast body-of-work on the design and optimization of energy-exchange or utility networks is the work on the synthesis of reactor and reactor=separation networks (e.g. Papalexandri and Pistikopoulous, 1996; Glasser and Hildebrandt, 1997; Sargent, 1998) and work on exploring some solution-space beyond existing system concepts where process alternatives have not been pre-postulated (Ismail et al., 2001). Contributionsthat focus on sustainable development are relatively few in number. In the past decade, the inclusion of environmental impact minimization into process synthesis was addressed (Buxton et al., 1997; Yang and Shi, 2000). The inclusion of environmental criteria into the design process was elucidated (Herder, 1999; Sharratt, 1999). The inherent multicriteria design problem of pollution prevention and waste management was addressed (Steffens et al., 1999; Mellor et al., 2002; Linke and Kokossis, 2003) as well as the incorporation of

process modi cations are essential to the economic health of the producer of commodity chemicals (Wei et al., 1979). A transition towards a sustainable petrochemical industry, however, requires innovation at all system levels, from reaction path to process system, from chemical plant to petrochemical complex or industry. These innovations must reach beyond traditional R&D and beyond the optimization of proven technology, proven system concepts and established process networks. The question arises therefore what innovations are required to meet the challenge of a sustainable petrochemical industry and how the required technological or systemic content can be speci ed? Addressing the second part of this question, in this paper we report some of our work on the development of a systemic approach to specify the desired content of process system innovations in the petrochemical industry. These are de ned as changes in the system structure or system design of the petrochemical industry, its industrial complexes, or individual plants (Table 1). These can be enabled by technological inventions or vice versa. Firstly, innovation, technological development and process synthesis are addressed and the research question is elaborated. Secondly, it is shown that functional modelling provides a suitable mechanism for abstraction from the present system content and from proven technology. Employing the procedure thus developed, conceptual process system innovations for the case of industrial aromatics are developed. Thirdly, a case of rethinking the functions of petrochemical complexes is presented: the realization of a trigenerate industrial complex.

SYSTEMS IN THE PETROCHEMICAL INDUSTRY Petrochemicals are produced in bulk quantities in largescale out-door facilities. Whilst these are complex installations in appearance, they can all be represented as a system where a number of physical inputs are transformed to a number of outputs, for example a steam cracker transforms naphtha to ole ns and aromatics. As far as the representation of the systems design is concerned, each petrochemical plant can be completely characterized by a single conguration of a limited number of process sections or functional units (Zevnik and Buchanan, 1963) such as reaction, separation, process ows and process recycles. Since operating conditions and system design can be selected to meet a range of objectives, and a variety of apparatus can be selected to implement each functional unit, the design of any commercial petrochemical plant represents only one out of a large set of feasible con gurations. Thus, a

Table 1. Overview of system levels in the petrochemical industry. System Petrochemical industry Complex Chemical plant Unit operation Apparatus System elements Complexes plants Chemical plants Unit operations Apparatus Internals Intra-connectivity Mass ow, infrastructures Mass & energy ows; utility infrastructure Mass & energy ows; utility network Direct transfer of mass, heat, power, impulse; Mechanical, electrical, chemical Degree-of-freedom Selection & con guration of elements; Feedstock=product state and composition Selection & con guration of elements; Feedstock=product state and composition Selection & con guration of elements; Conditions & ow composition Selection & con guration of elements; Operating conditions Principle & design

Trans IChemE, Vol 81, Part B, September 2003

FUNCTIONAL MODELLING FOR A SUSTAINABLE PETROCHEMICAL INDUSTRY life-cycle assessment into optimal chemical process system design (Azapagic, 1999; Burgess and Brennan, 2001; Khan et al., 2001). Others incorporated engineering thermodynamics into process design and optimization and suggested thorough consideration of the available energy resources and their conversion to useful work (Kalitventzeff et al., 2001; Bakshi, 2002). In process synthesis research, however, conscious re-design of the industry system structure beyond single plants has hardly attracted attention. Meanwhile, in the scienti c community at large the system design of the industry has been considered xed. With the maturation of the industry (Stobaugh, 1988), changes or extensions of the industry s structure have become incidental consequences of successful fundamental R&D. Thus it may be seen that one cannot be sure that the largely bottom-up approach employed in the scienti c community will provide suf cient and timely solutions to support a transition to a sustainable petrochemical industry. Systemic innovations do require a long time period for development and implementation irrespective of the mechanism by which they have been identi ed. The focus of our work, therefore, is on early conceptual specication or invention of the desired content of process system innovations for a sustainable petrochemical industry. METHODOLOGY Conjecture How to achieve sustainability is an open problem that cannot be solved by planning and optimization or traditional R&D alone. We conjecture that in order to prepare for a sustainable future of the petrochemical industry, integrated innovation at all system levels (Table 1) is required, from material to apparatus to plant to petrochemical complex and industry con guration. Furthermore, the industry system structure must be considered a degree-of-freedom in the search and speci cation of yet unknown technological developments, as well as in the adoption of innovations that originate from outside the sector. Objective and Research Question Our objective was the development of a systemic approach to specify the desired content of process system innovations in the petrochemical industry. This preferably includes and links together analysis and subsequent synthesis. The research question addressed is whether it is at all possible and worthwhile to devise some mechanism to systematize the search for technological innovations. Is there a top-down systems approach that allows assessment followed by the conceptual speci cation of R&D options for process system innovation in the petrochemical industry? Development and Testing The method for conceptualization of process system innovation has been based on the body-of-knowledge of process system engineering, thermodynamics and functional modelling. The method has been speci ed and re ned in the course of a number of case studies. Particularly, its application to the use of fuel cells in the chemical industry has led to a number of patents (Dijkema, 1997; Hemmes and Trans IChemE, Vol 81, Part B, September 2003

333

Dijkema, 1998), which is an illustration of their applicability albeit not a formal proof of the methods. Application of the methods to industrial aromatics led to a number of new system concepts for resource utilization improvement. We used functional modelling to develop trigeneration concepts. Since also their early evaluation would help to foster sustainable development, a straightforward linear programming (LP) model was used to assess the extent of CO2 reduction of a trigenerate industrial complex. FUNCTIONAL MODELLING Traditi onal Approach Initially, our objective was the development and testing of a systemic approach to specify the desired content of innovation of the petrochemical industry. In a traditional search for improvement options, a ranking of relatively poorly performing chemical processes is the logical startingpoint. Firstly, possible causes for the losses calculated must be established. Secondly, improvement opportunities within each of these chemical processes are investigated. Thirdly, opportunities that involve system change outside the chemical plant system boundary are investigated with the explicit condition that technology and structure of subsystems within each chemical plant are a degree-of-freedom. Fourthly, the opportunities thus formulated for each single plant must be tested for applicability to other chemical plants. Those that pass the test can be considered promising opportunities that can have wider application, possibly to all weak elements of the petrochemical industry identi ed. Finally, opportunities identi ed can be subjected to expert review. Example: Indus trial Aromatics In the Netherlands, benzene (B), toluene (T) and xylenes (ortho-xylene, oX and para-xylene, pX respectively) are produced from pyrolysis gasoline and reformed naphtha respectively. These aromatics serve as intermediate for a number of chemical product chains that lead to, amongst others, polystyrene, styrene copolymers such as ABS, nylon and polyethyleneterephthalate(PET). At the time, the Dutch industrial aromatics system comprised some 40 plants. An overview is given in Figure 1. The results of an assessment of the situation of the Dutch industrial aromatics system in 1990 using process information available in open-literature data (Chauvel and Lefebvre, 1989), indicate that the energy-ef ciency in this part of the Dutch petrochemical industry is high, some 77%. Ranking, however does yield a number of relatively weakly performing chemical processes, notably caprolactam, phenol and phthalic anhydride production (Dijkema and Grievink, 2000). The annual losses from the processes employed in the Dutch system together amount to some 110 PJ. Apparently, there is scope for improvement of these processes, and because of their large production capacities world-wide the total industry improvement potential is large. Relatively straightforward opportunities can be identi ed by looking beyond the system boundary of the particular process studied. The most obvious one is to reconsider the fate of unwanted by-products. In many plants utilization of such by-products within the plant system-boundary is by

334

DIJKEMA et al.

recovering the fuel-value only. To upgrade their utilization requires fuel replacement and identi cation of some destination of the by-products. Nevertheless, to upgrade by-products and to reap the associated bene ts is a longstanding tradition in the petrochemical industry which in many cases has led to new business development. In modern styrene units, for example, the hydrogen present in recycle off-gas is captured for pure hydrogen production. Also in older styrene plants or in the hydro-dealkylation of toluene other utilization of such gases than as fuel may be contemplated. In caprolactam, the large exergy loss and ammonium sulphate production were an incentive for the development of a new generation of caprolactam processes. Another opportunity is in catalyst research, where, traditionally, much research and development is devoted to nding the ultimate selective catalyst. Looking beyond the boundary of a single plant, however, suggests the development of effective catalysts for high conversion, high yield that are not necessarily selective to a single product. Such catalysts must be designed to be employed in process systems where the combination of unselective reaction and product separation can be effectuated at operating conditions that lead to limited exergy loss only. Concurrently, the possibilities to develop some utilization of by-products and the associated business must be considered. System Elements and System Decompositi on In studies of the chemical industry, the single chemical plant is usually selected as the central system level for analysis. Indeed, this is a reasonably identi able system element in the petrochemical industry, and information on typical chemical processes for almost every commercially produced commodity chemical can be found in the open literature. This selection, however, is hardly addressed explicitly in the literature (Dijkema and Reuter, 1999). From a number of case studies we concluded that modelling the petrochemical industry as a network of single plants suf ces for a rst assessment. However, on critical re ection, and in spite of the opportunities for innovation identied, we had some concern about the methodological underpinning of the identi cation procedure developed, because the selected decomposition of the industry into individual cases and process plants can be characterized as trivial and heuristic. This is unsatisfactory, because it may be

argued that the opportunities and research themes identi ed are largely a function of the decomposition of the industry applied. Therefore we investigated the possibilities of functional modelling (Baylin, 1990) to model petrochemical systems for the purpose of speci cation of the content of process system innovations. Functional Modelling, a Useful Concept? In functional modelling a function is concerned with what should be achieved, not how. The key feature of functional modelling is that the speci cation of systems, subsystems or system elements is completely technology-free (Asbjrnsen, 1992). As in system engineering and input=output analysis the particular means of action of any of these items is not described nor prescribed. At a particular system level functional modelling thereby provides a mechanism for abstraction of present system content in the search for new system con gurations. One of the functions within the subsystem where aromatics are consumed, for example, is the production of styrene. Presently, in the aromatics production system this function requires the production of essentially pure B. When, however, some direct conversion of 2 C4s (e.g., butadiene) to styrene becomes technically and economically feasible, the link between the production of B and production of styrene will be removed. Meanwhile the functional models of both the aromatics consumption and production system do not change because of such development (Figure 2). Functional modelling enables the search for opportunities for R&D, because the concept of functions allows: (a) technology-free speci cation of systems and alternatives. The technology content has thus effectively become a degree-of-freedom in petrochemical system design for intraprocess innovationsimprovement of currently employed processes; (b) suitable generalization of a system concept, because the abstraction provided by functional modelling allows expansion of an initial set of functions identi ed. Rather than limiting the function of a chemical process to the production of one or a few petrochemicals, a plant can be classi ed a multifunctional unit. Thereby, a perspective on new opportunities is opened for utilization and change of process system con gurations; and (c) optionally recalibration of the use and=or value of substances or ows in a petrochemical network. This opens the way to Trans IChemE, Vol 81, Part B, September 2003

FUNCTIONAL MODELLING FOR A SUSTAINABLE PETROCHEMICAL INDUSTRY

335

Figure 2. Change in functions.

inter-process innovations, as a chemical process systems waste products change into co-products, thereby increasing its net resource utilization. When confronted with the task of modelling an existing complex system, such as the petrochemical industry or any part thereof, the abstraction provided by functional modelling allows one to: (1) generalize at some top system level: for example the primary function of the petrochemical industry is to produce suitable feedstock for polymer production; (2) visualize the link of this subsystem of our industrial economy to other industrial sectors or parts thereof, thereby eventually allowing some linkage with needs of society catered for by industry, for example the need for hygienic packaging of food by the supply of suitable and affordable polymer lm; (3) arrive at a suitable functional decomposition of the petrochemical industry that nally allows the speci cation of technological or system content of innovations for sustainable development. In the industrial aromatics system, for example, at the top-level of abstraction, we can distinguish a production system for essentially pure B, T, oX and pX, and a system where their use as intermediate feedstock or consumption of these aromatics is a common denominator (Dijkema and Grievink, 2000); and (4) Adapt system boundaries. Because the decomposition is based on system and subsystem boundary selection it opens the way for consideration of alternative systems by expanding the set of functions in the model at a particular system level, or to view the system as a part of a larger system. Example: Functional Model of the Industri al Aromatics System Functional modelling of the aromatics system illustrates these items when modelling an existing system: (1) generalize: the general primary function of this subsystem of the petrochemical industry is similar to the primary function of the industry, produce a number of suitable feedstocks for polymer production; (2) visualize: the links of this subsystem with the petrochemical industry and beyond are Trans IChemE, Vol 81, Part B, September 2003

visualized in Figure 3. This indicates that the subsystem is largely dependent on the functions of the re ning sector and the associated technological implementations, as well as on the subsystem for the production of ole ns. The demand for automotive fuels that are void of volatile carcinogenic compounds has resulted in a reduction in allowable aromatics levels in gasoline (Auto Oil I and II). This can have a dramatic effect on the links between these sectors, as the modi cations in re ning change the supply and demand balance of aromatics world-wide, which will initiate process system change and development in all sectors; and (3) decompose. The aromatics system can suitably be decomposed in a production subsystem (BTXProd) and a consumption subsystem (BTXCon). Further decomposition of BTXCon is done by de ning system elements that produce the intermediates and polymers respectively. Further decomposition of the BTX production system, however, is not straightforward. Usually, aromatics are rst isolated from non-aromatics. Subsequently, somewhere in the production system the mixture (BTX) is transformed to essentially pure B, T, oX and pX, while mixed-xylene (mX) is usually converted and recycled to extinction. In present state-of-the-art there exist a variety of system con gurations that depend on feedstock selection and required productmix. Functional modelling, however, allows abstraction to a single functional model that encompasses all possible realizations of the conversion of complex hydrocarbon mixtures to the required product-mix of B, T, mX, oX and pX. The de nition of the functions listed in Table 2 suf ces. FUNCTIONAL MODELLING IN A STRATEGY FOR PROCESS SYSTEM INNOVATION The aromatics example above illustrates that functional modelling provides a mechanism to arrive at some technology-free superstructure for the petrochemical industry or parts thereof. To really contribute to the effectuation of sustainable development in system concepts selected for industrial implementation, production losses must be

336

DIJKEMA et al.

minimized or even eliminated at acceptable costs. The ultimate goal is to convert the feedstock into desired products only with minimum energy loss from the production system. A strategy to arrive at a functional speci cation of technological or system content for sustainable development in the petrochemical industry can comprise the following steps: (1) development of a functional model of the existing system (technology-free!) (2) assessment of the existing system implementation and identi cation of weak elements (3) development of alternative functional system concepts by using functional superstructure (4) evaluation of the scope of improvement achievable by the process system innovations developed

Table 2. Functions in the aromatics system (after Dijkema and Grievink, 2000). 1. Produce BTX-rich stream a) Pretreat and precondition b) Separate aromatics=non-aromatics c) Polish to required speci cation 2. Produce pure B 3. Produce pure T 4. Produce mixed-X 5. Produce pure pX 6. Produce pure oX 7. Produce pure mX 8. Produce BX-mixture 9. Produce TX mixture 10. Produce BT mixture

Steps 1 and 2 have been addressed above. Step 3 is where the process system innovations must be conceptualized and speci ed. As stated, based on the results of the aromatic system assessment, an R&D portfolio is preferred that includes options for the change of individual processes, process technology, the structure of the aromatics system, and the development of new process routes or products that replace existing functional links between this system and the industry. Step 4 can be completed when the innovative process system concepts have been developed into a conceptual process design, where again one employs some systematic series of steps to complete the design. For a plant that must produce a speci ed amount of a speci c product, these included assessment of building blocks, synthesis of alternatives, analysis of dynamic behaviour and evaluation of expected performance (Biegler et al., 1997). In the search for R&D directions, it is common to look at (incremental) improvement of currently employed processes. Using functional modelling, the same can be done. In the process network for the production of mixed-xylenes, for example, the ef ciency of the stabilization of pyrolysis gasoline can be increased by replacement of the selective hydrogenation of impurities in the feedstock. It appears that these impurities can be recovered by enhanced extraction as valuable products. Whilst this involves a change in the realization of function 1a in Table 2, pre-treatment, actually a new functional element is introduced: recover impurities Z Trans IChemE, Vol 81, Part B, September 2003

FUNCTIONAL MODELLING FOR A SUSTAINABLE PETROCHEMICAL INDUSTRY as products, which will have an effect beyond the particular process studied. In addition to intra-process innovations, we can search for process system innovations using the abstraction by functional modelling. One might argue, for example, that whilst production of B can be done via hydrodealkylation, the primary function of that process really is to upgrade T that would otherwise be sent to the gasoline-pool. Since this is a product destination that is limited today and may be blocked under the new Auto Oil regulations in the not too distant future, and because the hydrodealkylationprocess involves a relatively high energy-loss, would it be possible to conceptualize an aromatics system where this function is not employed? This can involve, for example, the development of applications of T and X that replace the use of B, which in fact implies the introduction of new intermediate functions in BTXCon. The conversion of T into para-methyl-styrene (PMS) represents such a new function (Keading et al., 1982). The polymer, poly-para-methyl-styrene (PPMS) can replace ordinary polystyrene in a range of applications, i.e., its functions. As a consequence, the actual realization of BTXProd will be affected, because the operation of the dealkylation process can be avoided, as well as its associated system inef ciencies. Its functional model, however, remains the same because dealkylation is only one option for realization of function 4, produce pure B. TOWARDS TRIGENERATION COMPLEXES? Introducti on The transformation of chemical plants to trigeneration systems for the simultaneous production of chemicals, electric power and heat is an example of generalization by functional modelling. In petrochemical plants, traditionally, heat is supplied in industrial furnaces and reaction furnaces, for example, the naphthacrackers or EDC cracking furnaces. In a study on the heat demand of the Dutch industry it was revealed that the chemical industry largely requires heat at a temperature level between 300 C and 800 C (Boot and Van Wees, 1982), which led the authors to remark that ample opportunity existed for cogeneration in this industry. Indeed most petrochemical sites in the Netherlands either already included cogeneration at the time or such facilities have been erected since. Most of these appear to have been designed for industrial processheat demand, the surplus power being transported to the grid. Functional Modelling for the Formulation of System Concepts Where the combination of chemical reaction and energy conversion is fundamental to chemical engineering and crucial to achieve proper plant design, surprisingly trigeneration is not. Trigenerate unit operations, however, have emerged recently under the label of multifunctional reactors, notably fuel cell reactors that incorporate some membrane. In contrast with such science-based innovation of unit operations within chemical plants, the concept of trigeneration suggests innovation beyond the boundary of individual plants, complexes or even the industry by expanding its set of functions (Dijkema, Grievink and Weijnen, 2001). Trans IChemE, Vol 81, Part B, September 2003

337

Thus, trigeneration represents a novel, functional view of the large-scale petrochemical industry: it not only serves to provide petrochemicals, but also acts as a large heat-sink that enables a net global CO2 emission reduction by trigeneration. The realization of trigeneration in the (petro)chemical industry thus involves a paradigm-shift. Rather than considering the industry a possible and threatening source of pollution, safety and health risks, it must be considered an enabler of a reduction in global CO2 emission by increased fossil resource utilization by viewing it a co-producer of chemicals, electric power and heat. In system design, the next step is to consider other options to bring about the required functions, in this case power generation from natural gas. Trigeneration system development implies that there must be an initial focus on what possible alternative realization of a particular system may exist in case the system performs 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 the example we have added to a hypothetical petrochemical complex the possibility of inclusion of additional system elements: stand-alone power generation (power plant) cogeneration type 3, heat-to-power ratio 3:1 cogeneration type 4, heat-to-power ratio 4:1 steam turbine operating between 600 C and 200 C steam generator, 600 C steam generator, 200 C These suf ce to give a rst impression of trigenerate petrochemical complexes. The list can be expanded, however, for a more sophisticated analysis, for example, more temperature levels. Notably the 600 C level produced by cogen-3 or cogen-4 can also be perceived as the temperature level of the heat-sink presented by industrial furnaces or industrial reaction furnaces. Thus, the analysis also leads to insight into the merits of trigenerate petrochemical plants by coupling of gas turbines to industrial furnaces.

Selection of Trigeneration System Concept Our objective was to evaluate the impact of converting a petrochemical complex to a trigenerate petrochemical complex on total CO2 emission. Thus, we compared the base case, optimal system design to meet internal power and heat exactly, with alternatives where a net exporter of electric power to the grid is anticipated. A straightforward linear programming (LP) model was constructed to select the optimal combination of system elements for heat and power generation. Since this illustration comprises a smallscale problem, it could easily be implemented in MS Excel. The optimal solutions were calculated using MS Excel Solver functionality. The model representation is (after Stadtherr and Rudd, 1976, 1978): Ac c r p 0

338

DIJKEMA et al.
Table 3. Coef cients used in the A-matrix. Includes net loss per element (MW). System elements Cogen-3 Energy carriers Fuel Electricity Heat600 C Heat200 C Energy loss 1 0.24 0.76 0 0.05 Cogen-4 1 0.19 0.76 0 0.05 Power plant 1 0.5 0 0 0.5 StmGen-1 1 0.02 0.85 0 0.17 StmGen-2 1 0.02 0 0.85 0.17 Stmtrbine 0 0.14 1 0.76 0.1

The elements of matrix Ac are the system elements inputoutput coef cients; c is the vector of capacities; r is the vector of feedstocks; r s (upply); p is the vector of products; p d (emand); An LP-model in standard form (Ax 0) is obtained by using A [AcjI j I ] x [c r p]T c 0;

These results demonstrate that trigeneration using a petrochemical complex as high-level heat-sink indeed leads to the minimum fuel consumption (MW): at no export of electricity HT and LT steam are produced as stand-alone. Any internal power demand can be supplied by the grid, or, if substantial, can be supplied by a Cogen-4. at limited electricity export, rst Cogen-4 capacity is realized or expanded, subsequently steam turbine capacity is added increasing the allowed electricity export to the grid brings Cogen-3 into being, thereby providing the optimal match between electric power and heat required nally, above 6 MW E-export (per 12 MW of heat HT:LT 4:8), only Cogen-3, a steam turbine and stand-alone power plant capacity are selected for optimal trigeneration The results show the optimal selection of system elements picked to satisfy internal heat demand with the amount of electric power that can be exported from an industrial complex. Secondly, the minimum fuel consumption calculated for trigeneration demonstrates that an industrial complex can be used as a heat-sink for electric power production. Finally, the analysis shows that the transformation to trigeneration leads to avoidance of CO2 emission compared with disjoint chemicals and electric power production. This has been illustrated in Table 4, where, based on the results of system element selected (Figures 4 and 5) the net avoided amount of CO2 has been calculated for each extra MW of electricity exported on top of internally produced 4:8 HT:LT heat. The numbers are valid assuming fuel is natural gas (methane) with a lower heating valve (LHV) of 50 (MJ kg1).

The optimal network design variables x0 are found by optimization. minx F subject to Ax 0; F(x) f T x

F is a vector of the coef cients of the linear objective function. The coef cients of the A-matrix are listed in Table 3. These have been expressed in megawatts (MW). A negative sign indicates an input, a positive sign an output. In this case, F is set such that r, which represents total fuel input is minimized. Thus, under the assumption of fossil fuel combustion to drive the energy conversion, in this optimization the net CO2 emission will also be minimized. Since the actual emission depends on the fuel-mix employed, a rst indication is obtained considering natural gas (CH4) as the fuel of choice. Figures 4 and 5 illustrate the optimization results for a petrochemical complex that is characterized by an export of electric power (E) and heat demand at 600 C (HT) and 200 C (LT) respectively in a ratio of HT:LT:E 4:8:E.

Figure 4. Total trigenerate power generation and power generation distribution as a function of power export to the grid.

Figure 5. Total trigenerate fuel consumption and fuel distribution as a function of electricity export to the grid.

Trans IChemE, Vol 81, Part B, September 2003

FUNCTIONAL MODELLING FOR A SUSTAINABLE PETROCHEMICAL INDUSTRY


Table 4. Avoided amount of CO2 (kg MWh 1) per interval of extra trigenerated electricity exported (MW). Electricity export (MW) 1st 2nd 3rd 4th 5th 6th 7th 8th12th CO2 reduction [kg=MWh] 276 191 191 191 191 190 73 0

339

CO2 credits. Thus, a paradigm-shift in perception of the petrochemical industry may result: instead of a threatening source of pollution, safety and health risks, it becomes an enabler of global CO2 emission reduction. Functional modelling provides a structured method for abstraction, that allows one to freely consider novel realizations of existing functions of the chemical industry at every system level of interest, and thereby address the open problem of process system innovation in a systematic, topdown approach. Only after new concepts have developed, must the hard work of assessment of their technological and economic impact begin. REFERENCES
Asbjrnsen, O.A., 1992, System Engineering Principles (Skarpodd, Houston, USA). Azapagic, A., 1999, Life cycle assessment and its application to process selection, design and optimisation, Chem Eng J, 73(1): 121. Bakshi, B.R., 2002, A thermodynamic framework for ecologically conscious process systems engineering, Comput Chem Eng, 26(2): 269282. Baylin, E.N., 1990, Functional Modeling of Systems (Gordon and Breach Science Publishers, New York, USA). Biegler, L.T., Grossmann, I.E. and Westerberg, A.W., 1997, Systematic Methods of Chemical Process Design (Prentice Hall PTR, Upper Saddle River, NJ, USA). Boot, A.H. and Van Wees, F.G.H., 1982, Industrial Process Heat in Relation to the Temperature Level (Energy Research Centre of the Netherlands (ECN), Petten). Burgess, A.A. and Brennan, D.J., 2001, Application of life cycle assessment to chemical processes, Chem Eng Sci, 56(8): 25892604. Buxton, A., Livingston, A.G. and Pistikopoulos, E.N., 1997, Reaction path synthesis for environmental impact minimization, Comput Chem Eng, 21(1): S959S964. Charpentier, J.-C., 2002, The triplet molecular processes-product-process engineering: the future of chemical engineering?, Chem Eng Sci, 57(2223): 46674690. Chauvel, A. and Lefebvre, G., 1989, Petrochemical Processes. Technical and Economic Characteristics. 1. Synthesis-Gas Derivatives and Major Hydrocarbons (Gulf Publishing Company, Houston, London, Paris, Tokyo). Dijkema, G.P.J., 1997, Method for the manufacture of ammonia. WO 97=33828, Technische Universiteit Delft [NL=NL]; Julianalaan 134, NL-2628 BL Delft (NL): 110. Dijkema, G.P.J. and Reuter, M.A., 1999, Dealing with complexity in material cycle simulation and design, Comput Chem Eng, 23: S795S798. Dijkema, G.P.J. and Grievink, J., 2000, Technological change by system design-the industrial production of aromatics, European Symposium on Computer Aided Process Engineering, Escape-10, Florence, Italy, in Computer-Aided Chemical Engineering, Vol. 8, Pierucci, S. (ed) (Elsevier Science). Dijkema, G.P.J., 2001, The development of trigeneration system concepts, 6th World Congress of Chemical Engineering, Melbourne, Australia, CD-ROM paper 620. Dijkema, G.P.J. and Mayer, I.S., 2001, Corporate strategy and productoriented environmental policy, 6th World Congress of Chemical Engineering, Melbourne, Australia, CD-ROM paper 642. Dijkema, G.P.J., Grievink, J. and Weijnen, M.P.C., 2001. Process system innovation for a sustainable petrochemical industry, 6th World Congress of Chemical Engineering, Melbourne, Australia, CD-ROM paper 654. Glasser, D. and Hildebrandt, D., 1997, Reactor and process synthesis, Comput Chem Eng, 21: S775S783. Grossmann, I.E. and Westerberg, A.W., 2000, Research challenges in process system engineering, AIChE J, 46(9): 17001703. Hemmes, K. and Dijkema, G.P.J., 1998, Method of operating a molten carbonate fuel cell, a fuel cell, a fuel cell stack and an apparatus provided therewith WO 98=08264, Technische Universiteit Delft [NL=NL]; Julianalaan 134, NL-2628 BL Delft (NL). 121. Herder, P.M., 1999, Process Design in a Changing Environment: Identi cation of Quality Demands Governing the Design Process (Delft University Press, Delft, The Netherlands). Ismail, S.R., Proios, P. and Pistikopoulos, E.N., 2001, Modular synthesis framework for combined separation=reaction systems, AIChE J, 47: 629649.

DISCUSSION AND CONCLUSIONS The need for sustainable development provides a major incentive to elucidate the question of process system innovation. The demand for sustainability has emerged, however, at a time where the petrochemical industry is largely in a phase of maturation. Combined with the demands of shareholders, this has lead to a wave of restructuring of the industry to increase economic performance. Measures employed include mergers and acquisitions, productportfolio shift, and restructuring of organizations. Whitehead, however, signals a lack of strategic initiatives to improve the economic performance of the industry (Whitehead, 2000). We believe that process system innovation and the methods developed may contribute to a sound strategy for sustainability that includes improved ecological, economic and societal performance. Sustainable development requires innovation at all system levels, from reaction path to process system, from chemical plant to petrochemical complex or industry. Any new approach to innovation in the chemical industry however, must have sound technological and system content in order to have a chance to achieve some impact in the industry. We believe that functional modelling combined with evaluation as presented in this paper does meet this criterion, and therefore may help industry in its struggle with the concept of sustainability, and its perceived lack-of-content. The examples given for industrial aromatics production and trigenerate industrial complexes demonstrate the usefulness of functional modelling for the identi cation of novel opportunities for technological and system change. The aromatics example emphasizes the link between product development and process system innovation. The structure of trigenerate petrochemical complex varies subject to the technologies available and the amount of electric power that can be exported to the grid. By employing a straightforward LP-model, it was demonstrated that the large-scale petrochemical industry can enable a net reduction of global CO2 emission by trigeneration, where its production systems act as a large heat-sink. In bringing trigeneration systems into being, one must face uncertainties with respect to initial capital expenditure and the margins that depend on electricity, heat and chemicals revenue. Trigeneration, however, appears to be a promising option for base-load, which matches the petrochemical industrys requirement of continuous capacity utilization. Moreover, in a liberalized electricity market, trigeneration may represent one of the hidden or emergent innovation options. The market will simply set a price for the base-load electric power from the trigenerate petrochemical industry, as well as for any alleged Trans IChemE, Vol 81, Part B, September 2003

340

DIJKEMA et al.
Steffens, M.A., Fraga, E.S. and Bogle, I.D.L., 1999, Multicriteria process synthesis for generating sustainable and economic bioprocesses, Comput Chem Eng, 23(10): 14551467. Stobaugh, 1988, Innovation and Competition. The global management of petrochemical products (Harvard Business School Press, Boston, USA). Wei, J., Russell, T.W.F. and Swartzlander, M.W., 1979, The Structure of the Chemical Processing Industries, Function and Economics (McGraw-Hill, Inc., New York, USA). Wetenschappelijke Raad Voor Het Regeringsbeleid, 2003, Naar nieuwe wegen in het milieubeleid (Sdu Uitgevers, Den Haag, The Netherlands). Whitehead, 2000, Looking from the outside in: achieving value growth in asset-intensive industries, Chim Oggi, 18(5). World Commission on Economic Development, 1987, Our Common Future (Oxford University Press, Oxford, UK). Yang, Y. and Shi, L., 2000, Integrating environmental impact minimization into conceptual chemical process design-a process systems engineering review, Comput Chem Eng, 24(27): 14091419. Zevnik, F.C. and Buchanan, R.L., 1963, Chem Eng Progr, 59(2): 70.

Kalitventzeff, B., Marechal, F. and Closon, H., 2001, Better solutions for process sustainability through better insight in process energy integration, Appl Therm Eng, 21(1314): 13491368. Keading, W.W., Young, L.B. and Prapas, G., 1982, Para-methylstyrene, Chem Tech, 12: 556562. Khan, F.I., Natrajan, B.R. and Revathi, P., 2001, GreenPro: a new methodology for cleaner and greener process design, J Loss Prevent Proc, 14(4): 307328. Linke, P. and Kokossis, A., 2003, Advanced process systems design technology for pollution prevention and waste treatment, Adv Environ Res (in press, corrected proof). Marcoulaki, E.C. and Kokossis, A.C., 2000, On the development of novel chemicals using a systematic synthesis approach. Part I. Optimisation framework, Chem Eng Sci, 55(13): 25292546. Mellor, W., Wright, E., Clift, R., Azapagic, A. and Stevens, G., 2002, A mathematical model and decision-support framework for material recovery, recycling and cascaded use, Chem Eng Sci, 57(2223): 46974713. Papalexandri, K.P. and Pistikopoulos, E.N., 1996, Generalized modular representation framework for process synthesis, AIChE J, 42: 10101032. Pavitt, K., 1984, Sectoral patterns of technical change: towards a taxonomy and a theory, Res Policy, 13: 343373. Rowe, D.A., Perkins, J.D. and Walsh, S.P., 1997, Integrated design of responsive chemical manufacturing facilities, Comput Chem Eng, 21: S101S106. Sargent, R.W.H., 1998, A functional approach to process synthesis and its application to distillation systems, Comput Chem Eng, 22(12): 3145. Sharratt, P., 1999, Environmental criteria in design, Comput Chem Eng, 23(10): 14691475. Siirola, J.J., 1996, Industrial applications of chemical process synthesis, In Process Synthesis, Anderson, J.L. (ed) (Academic Press, New York, USA), 23: 261. Stadtherr, M.A. and Rudd, D.F., 1976, Systems study of the petrochemical industry, Chem Eng Sci, 31(11): 10191028. Stadtherr, M.A. and Rudd, D.F., 1978, Resource use by the petrochemical industry, Chem Eng Sci, 33(7): 923933.

ACKNOWLEDGEMENT
This paper has been based on material originally presented at the 6th World Congress of Chemical Engineering, Melbourne, Australia, September 2001 (Dijkema, Grievink and Weijnen, 2001).

ADDRESS
Correspondence concerning this paper should be addressed to Ir. G.P.J. Dijkema, Faculty of Technology, Policy and Management, Delft University of Technology, Delft, The Netherlands. E-mail: g.p.j.dijkema@tbm.tudelft.nl The manuscript was received 1 April 2003 and accepted for publication after revision 3 September 2003.

Trans IChemE, Vol 81, Part B, September 2003

You might also like