100% found this document useful (1 vote)
239 views

Exergy Analysis of Chemical Plant

Uploaded by

Imran Niaz Khan
Copyright
© © All Rights Reserved
Available Formats
Download as PDF, TXT or read online on Scribd
100% found this document useful (1 vote)
239 views

Exergy Analysis of Chemical Plant

Uploaded by

Imran Niaz Khan
Copyright
© © All Rights Reserved
Available Formats
Download as PDF, TXT or read online on Scribd
You are on page 1/ 40

468 Int. J. Exergy, Vol. 15, No.

4, 2014

An overview of exergy analysis for chemical process


industries

Samir C. Nimkar
Department of Chemical Engineering,
Bharati Vidyapeeth Institute of Technology,
Navi Mumbai 400614, Maharashtra, India
E-mail: [email protected]

Rajubhai K. Mewada*
Department of Chemical Engineering,
Institute of Technology,
Nirma University,
Ahmedabad 382481, Gujarat, India
E-mail: [email protected]
*Corresponding author

Abstract: Among all the industrial sectors, the chemical industry is one
of the energy intensive industry with 33% share in industrial energy use.
Increasing cost of raw materials, infrastructure cost and energy cost increases
burden on the profit margin of the chemical sector. Chemical industries
are looking for improved process design and efficient equipments where
energy consumption can be reduced. In such a situation, exergy analysis
can acts as an important tool to pinpoint energy wastage. It also provides
an insight for improvement in the process or equipment to enhance its energy
efficiency. In this paper, major unit operations and processes involved
in various chemical industries are discussed with recommendations for
reduction in exergy losses along with several case studies. Overview of process
improvement and environmental protection using exergy analysis is also
carried out.

Keywords: exergy analysis; exergy destruction; exergy efficiency;


chemical process industries; inorganic chemical industry; distillation column;
refinery.

Reference to this paper should be made as follows: Nimkar, S.C. and


Mewada, R.K. (2014) ‘An overview of exergy analysis for chemical process
industries’, Int. J. Exergy, Vol. 15, No. 4, pp.468–507.

Biographical notes: Samir C. Nimkar is working as lecturer at Chemical


Engineering Department, Bharati Vidyapeeth Institute of Technology, Navi
Mumbai, India. He is certified energy auditor of BEE, India. He holds BTech
and ME in Chemical Engineering. He is teaching energy management,
process technology to the undergraduate students and worked as a core
committee member for curriculum design. He is also involved in the training
for employees of the various industries.

Copyright © 2014 Inderscience Enterprises Ltd.


An overview of exergy analysis for chemical process industries 469

Rajubhai K. Mewada is working as Associate Professor at Chemical


Engineering Department, Institute of Technology, Nirma University,
Ahmedabad, India. His area of research is green process development and
design, catalysis and reaction engineering, solar energy utilisation in chemical
processes, modelling and simulation for multiphase reactions and operations.
He holds PhD in Chemical Engineering from Institute of Chemical
Technology, Mumbai, India.

1 Introduction

World’s primary energy consumption grows by 45% over the past 20 years (BP
Energy Outlook 2030, 2011). Total world energy use will rise from 580.73 × 103 PJ
to 813.02 × 103 PJ in 2035. Worldwide industrial energy consumption will grow
from 201.67 × 103 PJ in 2008 to 304.10 × 103 PJ in 2035(International Energy Outlook,
2011). World energy related carbon dioxide emission will also rise up to 43.2 billion
MT in 2035. Among all industrial sectors chemical, iron and steel, non-metallic minerals,
paper and non-ferrous metal manufacturing accounts for the majority of all industrial
energy consumption.
The chemical industry has successfully reduced its energy consumption after the oil
crisis during 1970–1980. But in last few years energy efficiency improvement is almost
steady (Chemical Bandwidth Study, 2004). The challenges ahead are increasing fuel cost
and competition in international market. The chemical industry is looking forward
for various alternatives to reduce production cost by improving process, modification
of equipments and analysing resource consumption. As shown in Figure 1, chemical
industries are operating above the practical minimum energy requirement. Exergy
analysis is one of the tools to pinpoint true energy losses at various stages and
equipments in a chemical plant. Sciubba and Wall (2007) presented a brief history
about the development in exergy from 1824 to 2004. More than 2600 references are cited
to explain the progress of exergy in every field ranging from power cycles to the social
system. Exergy analysis of the industrial sector is reviewed by Boroumandjazi et al.
(2013). The separation process is more energy intensive in the chemical industry.
The current applications of exergy analysis in distillation operation, membrane
technology and CO2 capture are reviewed by Luis (2013). Though first known work of
Rant Z. is in the year 1956, exergy gain importance in chemical engineering after 1990.
Out of total publications in the field of exergy, <15% relate to chemical engineering
(Luis, 2013). Exergy analysis of cement, petrochemical, chemical and iron production
shows significant improvement potential. In the present paper, an attempt has been made
to take an overview of exergy analysis of chemical processes. The scope of the study is
limited to units and complexes (Section 3), bulk inorganic chemicals (Section 4),
crude oil refinery with petrochemicals (Section 5) and process improvement (Section 6).
Exergy use in pollution control (Section 7) is also discussed. References used in the paper
are mainly those which are available in English from year 1985 to 2013. In the view of
simplification and uniformity, notations used in this paper are based on the nomenclature
suggested by Tsatsaronis (2007).
470 S.C. Nimkar and R.K. Mewada

Figure 1 Depiction of energy bandwidth (see online version for colours)

Source: Chemical Bandwidth Study (2004)

2 Exergy analysis

Exergy is the availability of energy in provided resource and its value is depends upon
reference environment. The exergy of the system is defined as the maximum shaft work
that could be done by composite of the system and a specific reference environment that
is assumed to be infinite, in equilibrium, and ultimately to enclose all other systems
(Dincer and Cengel, 2001).

2.1 Flow and non-flow exergy


Exergy of a closed system can be expressed as (Tsatsaronis, 2007)
PH
Esys = Esys + E KN + E PT + E CH . (1)

Physical exergy is the work obtained when the system is brought from its original state to
environmental state and Chemical exergy is the work obtained when the system is
brought from its environmental state to the dead state.
PH
Esys = (U − U 0 ) + P0 (V − V0 ) – T0 ( S − S0 ) , (2)

E CH = ∑ ( µi 0 – µi 00 ) ni , (3)
i

E PT = mgz, (4)
An overview of exergy analysis for chemical process industries 471

1
E KN = mν 2 . (5)
2
The exergy of flowing stream is written as (Dincer and Rosen, 2007)
E = Esys + ( P – P0 ) V . (6)

Term (P – P0)V is exergy associated with flow work of stream. From equation (2)
and (6)
E PH = (U − U 0 ) + P0 (V − V0 ) – T0 ( S − S0 ) + ( P – P0 ) V . (7)

Rearranging equation and putting H = U + PV we obtain


E = ( H − H 0 ) − T0 ( S − S0 ) + E KN + E PT + E CH . (8)

Quality of hot stream is depends upon the work obtained from it. According to second
law, it is not possible to convert all heat into work though it contains lots of energy.
Exergy of thermal energy is denoted by
E Q = (1 − T0 / T ) Q = τ Q, (9)

where τ = (1 – T0/T) is called the exergetic temperature factor.

2.2 Chemical exergy and reference environment


Chemical exergy can be defined as maximum work that can be obtained
when a substance is brought to state of reference substances from its present state. The
standard chemical exergy of any chemical compound can be calculated by (Szargut et al.,
2005)
EnCH = ∆G f + ∑ ne E CH ne , (10)

where
∆Gf: Formation Gibbs energy
n e: Amount of kmol of element e
CH
E ne: Standard chemical exergy of element e.
For the calculation of the chemical exergy of the reference gases
EnCH = − RT0 ln P0 n / Pn = − RT0 ln z0 , (11)

where
P0n: Conventional mean ideal gas partial pressure in the atmosphere
Pn: Standard pressure
z0: Conventional standard molar fraction in the environment.
Chemical exergy of substance depending upon reference environment, as natural
environment is not in equilibrium. Different reference environment models are proposed
472 S.C. Nimkar and R.K. Mewada

by various researchers (Gaggioli and Petit, 1977; Sussman, 1980; Ahrendts, 1980),
but Szargut’s reference environment (Szargut et al., 1988) is most popular amongst
engineers. His model consists of a reference substance based on its availability
in environment abundantly. Nitrogen, Oxygen, carbon dioxide, water, noble gases,
gypsum and limestone are part of this environment. Comparison of various reference
environments is carried out by Valero (2008) and Gaudreau et al. (2012). One reference
model cannot be accepted internationally because all models are having its own
advantages and disadvantages. Thus the international reference environment model
is proposed by Szargut et al. (2005) by analysing various models that are previously
proposed.

2.3 Exergy efficiency


There are several ways to define exergy efficiency. The common way is the ratio of
output exergy to input exergy. But output exergy consists of both useful exergy and
exergy in waste. Useful exergy which can be utilised may be incorporated to get a new
type of exergy efficiency. Part of exergy that leaves the system unchanged with utilisable
exergy is called transiting exergy, Etr (Brodyansky et al., 1994). It is a part of exergy
associated with feed or inerts.
ηe = ( Eout – Etr ) / ( Eout – Etr ) . (12)

Rational efficiency is the ratio of the desired exergy output to the used exergy (Kotas,
1995)
Rational efficiency = Total exergy output/Total exergy input
= 1 – Exergy consumption/Total exergy input.
Rational efficiencies of various operations are listed in Table 1. Exergy efficiency helps
to improve the process by pinpointing the losses. The Exergy efficiency of the whole
chemical industrial sector in the USA is 32%. Inorganic industries are having 29%
efficiency and organic industries are having 35% efficiency (Ayres et al., 2011). Analysis
is carried for the sector by dividing chemicals into two groups:
• basic chemicals
• intermediate chemicals.
Basic chemicals are manufactured by using different elements and intermediate
chemicals are formed by using basic chemicals. For inorganic industry NaCl, S,
CH4, O2, N2 and H2 are mass inputs. Methane is considered as one of the input because
it is used as raw material for ammonia production. Same way sulphuric acid is used in
petroleum refining and petrochemical industries. Inputs for organic industries
are hydrocarbon feed (HC), Cl2, H2SO4, NH3, NaOH. Figure 2 shows 60% mass
is lost in the organic chemical industry and 8.6% mass is lost in inorganic industry
in the USA. More exergy input is observed in the organic chemical industry due to
hydrocarbon feed.
An overview of exergy analysis for chemical process industries 473

Table 1 Rational efficiency of commonly used unit operations

Equipment Function Exergy efficiency

Steam turbine: Shaft work EShaftwork


Expander
generation in power plant EIN − EOUT

EOUT − EIN
Compressor Adiabatic: Increase of exergy
EShaftwork

EOUT − EIN
Pump Increase of pressure
EShaftwork

EOUT − EIN
Heater or cooler Heating or cooling by utilities Q
EUtility

EC 2 − EC1
Heating
E H 1 − EH 2
Heat exchanger
E H 1 − EH 2
Cooling
EC 2 − EC1

Separation of product 1 and 2 from Eproduct1 + Eproduct 2 − EFeed + EUseful heat


Separator
feed EUseful heat − EShaftwork
CH CH
Endothermic Eproduct − Efeed
Formation of product from feed
reactor Q
EEndothermic

Source: Ghannadzadeh et al. (2012)

Figure 2 Mass (MMT) and exergy (PJ) balance of the production of inorganic and organic
chemical in USA in 1991 (see online version for colours)

Source: Ayres et al. (2011)

2.4 Exergy destruction


In the real system input exergy is not equal to output exergy. Some part of input exergy is
destroyed within the system. An ideal process is perfectly reversible and there is no
exergy destruction within the system. In the real processes exergy destruction takes place
due to thermodynamic irreversibility. Irreversibility (I) can be written as
474 S.C. Nimkar and R.K. Mewada

I = T0 Sgen . (13)

Equation (13) is known as Gouy–Stodola equation (Dincer and Rosen, 2007). Calculation
of exergy destruction will help to locate targets for improvement. A general expression
for lost work or exergy destruction of a control volume over a period of time ∆t is (Seider
et al., 2009)
ED = −∑ Wi − D[m ( H – T0 S )]flowing stream
i
(14)
+ ∑ (1 − T0 /Ti ) Q − ∆  m ( H − T0 S − PV )  sys / ∆t.
i

The above equation can be reduced to calculate exergy destruction of two stage
compressor
ED = [m ( H – T0 S )]stage1 + m ( H – T0 S )]stage2 + Wele,in – Exergy lost with CW , (15)

Wele,in is electrical power given to the compressor. Overall exergy destruction in


refrigeration cycle is derived from equation (14)

ED = Win + (1 − T0 /Tevaporator ) Qin – (1 − T0 /Tcondenser ) Qout . (16)

Evaporator and condenser are the important parts of refrigeration cycle where exergy
destruction takes place. External energy for the compressor is in the form of electricity.
Properly designed heat exchanger can transfer almost 100% energy but not exergy.
Exergy loss in hot and cold fluid in the heat exchanger is shown by equations (17)
and (18) (Wall, 2009).
∆EH = EH 1 − EH 2 = mH CP , H (TH 1 − TH 2 − T0 ln TH 1 /TH 2 ) , (17)

∆EC = EC 2 – EC1 = mC CP ,C (TC 2 – TC1 − T0 ln TC 2 /TC1 ) . (18)

The total loss of exergy in the heat exchanger is


∆E = ∆EH – ∆EC . (19)

3 Exergy analysis of process equipments

3.1 Exergy analysis of distillation column


Distillation is a very important unit operation in chemical process industries which
comprises 95% of the all separation processes. Distillation accounts for more than 50% of
the plant operating cost. Energy consumption can be reduced by optimising operating
parameters in the column. Irreversibility produced by mixing streams of various
concentrations and temperatures causes exergy loss in the column. Distillation column’s
exergy performance is sensitive to temperature pinch (Mustapha et al., 2007). Owing to
requirement of large driving force it exchanges more energy in form of heating and
cooling. Equation (20) shows loss of exergy in the column including condenser, reboiler
and pumps (Seider et al., 2009).
An overview of exergy analysis for chemical process industries 475

ED = −∑ Wele + [m F ( H F – T0 S F ) − m D ( H D – T0 S D ) − m B ( H B – T0 S B )]
i (20)
+ (1 − T0 / TH ) QRB − (1 − T0 /TC ) QCON .

Driving force distribution in column is described by Zemp et al. (1997) using the exergy
loss profile. The minimum energy required by using geometry and thermodynamic
condition is studied by Ayotte-Sauvé and Sorin (2010). A reversible column profile and
pinch point curve is also used for characterisation of minimum energy requirement.

3.1.1 Adiabatic distillation column


Two level idealisation concept is introduced by Chang and Chuang (2003), and Chang
and Li (2005) to reduce exergy losses in column. The first level idealisation
is the reversible operation and the second level idealisation is the thermodynamic
equilibrium operation. Intrinsic exergy losses caused by configuration constraints are
arising from first level idealisation and extrinsic exergy losses caused by transport rate
limitation are arising from second level idealisation. Intrinsic loss due to changes in
concentration is

Eint, ∆C = −T0 ( SV ,in + S L ,in – SV ,out,lim – S L ,out,lim ) . (21)

Limitation of two level idealisation method regarding the trend of optimum values is
rectified by four level idealisation concept. It is used to decide operating conditions of
column to reduce exergy losses. Khoa et al. (2010) used three-dimensional graphical
exergy analysis curves to visualise correlations between exergy lost, design and operating
parameters of a distillation column. Three dimensional exergy analysis curve is also
useful for exergy analysis of absorption column (Khoa et al., 2012). Exergy loss as a
function of L/G ratio is plotted on the curve and it will be helpful to determine design
parameters of absorption column.
In four level idealisation concepts, each level corresponds to one operating state of
the column. First level is reversible operation, second is an ideal adiabatic column, third
is thermodynamic equilibrium operation with infinite transport rate and finite
configuration and fourth is infinite configuration and finite transport rate. Exergy lost in
column can be reduced by operating from its current level to third level by transport rate
improvement and to fourth level by improving configuration such as increasing numbers
of trays.
Major exergy loss located at stages below the feed stage in deethanizer column is
reduced by using two level idealisation concept. An inter reboiler is introduced at a stage
of peak exergy loss reduced intrinsic exergy losses by 11.42%. Extrinsic exergy loss
reduction needs dimensional changes of the column. A maximum loss can be reduced by
increasing weir height, length and decreasing reflux ratio. Three-dimensional exergy
analysis curve is applied to xylene column having furnace heat duty of 99.132 MW
(Figure 3). Exergy loss in the column is 26.482 MW (i.e. 26.71% of input furnace
heat duty). The optimum value of the reflux ratio by considering product purity is
2.26 compared with initial values of 2.63. Column exergy loss is reduced by 15.5% using
new reflux ratio and preheating the feed. In both the concepts, reduction of the reflux
ratio will increase exergy efficiency. Exergy loss can be reduced by 12–21% if the
476 S.C. Nimkar and R.K. Mewada

single feed is split into two and fed at different locations by preheating them (Olakunle
et al., 2011). Changes in the geometry of column involve capital cost. Economic viability
is the key factor for such modifications.

Figure 3 The three dimensional exergy analysis curve for xylene column (see online version
for colours)

Source: Khoa et al. (2010)

Ishida and Nakagawa (1985) proposed energy utilisation diagram (EUD) for the study of
energy transformation and exergy analysis. Exergy loss in distillation column can be
analysed by using EUD. It is used to show six types of exergy losses taking place on
column tray. These are
• cooling process of the vapour phase
• the heating process of the liquid phase
• the mixing process of the vapour phase
• the mixing process of the liquid phase
• the condensation of heavier component
• the evaporation of the lighter component (Taprap and Ishida, 1996).
An overview of exergy analysis for chemical process industries 477

Column grand composition curve (CGCC) and temperature composition profile will
help in assessing the side heating and cooling requirements to reduce exergy losses.
Both diagrams can be generated by using popular simulation engines. The method
is simple and easy compared with two level and four level idealisation concepts. This
method is able to reduce exergy loss to more than 20% by installing side condensers and
side reboilers (Faria and Zemp, 2005; Demirel, 2006). Similarly exergy grand composite
curve (EGCC) is also used to find out location and duty of side exchangers (Wei et al.,
2012).

3.1.2 Diabatic distillation column


Distillation Columns used in the industry are adiabatic columns with the condenser
at the top and reboiler at bottom. Maximum exergy loss is observed on the tray having
largest composition difference. Exergy loss in the column can be reduced by adding
heat exchangers on each tray. This column is called diabatic column. In diabatic column,
heat is exchanged depending upon the exergy loss profile of the column. Figure 4 shows
exergy loss profiles with equal heat duty on each tray for diabatic column (Rivero, 2001).
The second law efficiency of the column is (Koeijer and Rivero, 2003)

ε = W min / (W min + Ex loss ) , (22)

ED = T0 ds irr /dt. (23)

Figure 4 Distribution of exergy losses in diabatic column

Source: Rivero (2001)


478 S.C. Nimkar and R.K. Mewada

The entropy production rate due to interface of mass and heat transfer on the nth tray is
dsnirr /dt = J q , n X q , n + J w, n X w, n + J e , n X e , n . (24)

The average force for heat transfer on tray n is


1
X q , n = ∆ (1/ T ) = (1/ Tn +1 ) – (1/ Tn −1 )  . (25)
2
The total entropy production rate for tray n with a heat exchanger becomes

dsnirr /dt = J q , n X q , n + J w, n X w, n + J e , n X e , n + ∆H vap ( J w, n + J e , n ) X nHX . (26)

Figure 5 shows the difference between adiabatic and diabatic column exergy analysis
for the ethanol – water system. Results of exergy analysis of various diabatic columns are
shown in Table 2. By increasing the number of trays in diabatic column, exergy saving
also increases due to approaching reversible separation for the column (Jimenez et al.,
2004). Exergy improvement is also possible by integrating heat exchanger at the top
and bottom instead of using separate condenser and reboiler (Le Goff et al., 1996).
Heat and mass transfer effects in this column can be seen simultaneously by plotting
specific enthalpy verses Carnot factor. Use of double diabatic column in place of double
adiabatic column in cryogenic air separation reduces 23% exergy loss (Rizk et al., 2012).
In naphtha cracking process also double diabatic column is considered superior to the
adiabatic distillation system (Hirata, 2009).

Table 2 Exergy analysis of various diabetic columns

Reduction of exergy loss after


converting the adiabatic column
System to diabatic column (%) Workers
Ethanol–water 35.92 Rivero (2001)
(45 mole % ethanol in feed)
Ethanol–water 39.00 Koeijer and Rivero (2003)
(6.75 mole % ethanol in feed)
Depentenizer 35.33 Rivero et al. (2004a)
Propylene–Propane 05.58 Røsjorde and Kjelstrup
(2005)
Benzene–Toluene 35.00 Jimenez et al. (2004)
Cryogenic air separation 23.00 Rizk et al. (2012)

3.1.3 Dividing wall column


In dividing wall column (DWC) working of two columns takes place in the single
shell using a wall. It helps to save both energy and capital. The feed side of the two
compartments acts as the prefractionator and the product side as the main column as
shown in Figure 6. In the case of a sharp split, a DWC can be used to produce three
pure products from a single tower. In recent time DWC are installed successfully
by BASF, ExxonMobil and Dow at various locations (Pendergast et al., 2008).
An overview of exergy analysis for chemical process industries 479

Energy saving achieved for various processes is in the range of 25–50% (Slade et al.,
2006; Isopescu et al., 2008; Dejanovića et al., 2010; Kiss and Rewagad, 2011; Sumiju
Plant Engineering Co., Ltd, 2010). DWC operates as a single unit by replacing two or
three columns hence it is better from the first law prospective. Suphanit et al. (2007)
applied exergy analysis to DWC to determine addition or rejection of heat at a particular
stage of BTX column. Minimum driving force profile is used to identify suitable
locations by dividing the column. Exergy loss in the column can be reduced up to 8.2%
by allowing heat transfer across the wall of the column.

Figure 5 Grassmann diagram of adiabatic and diabatic column

Source: Koeijer and Rivero (2003)

3.2 Exergy analysis of reactor


Chemical changes in the process industry are due to chemical reactions. Available
reaction energy is depending upon the standard Gibbs energy of the overall reaction.
It helps to understand exergy losses associated with a chemical reaction. Physical exergy
obtained in the reaction is a part of chemical exergy (Hinderink et al., 1996a). This
conversion cannot be completed without exergy loss as shown in Figure 7. The equation
for exergy destruction in a plug flow reactor is developed by Ruyck (1998).

dI = δ q chem 1 − (Tsurr / T )  – δ q chem 1 − (Tsurr / Trev )  . (27)

The entropy production rate of exothermic ammonia reaction in a tubular fixed bed
reactor can be varied by changing the inlet condition and coolant temperature (Nummedal
et al., 2003). The change in entropy production rate is observed with different
heat transfer coefficient. The value of the overall heat transfer coefficient plays major
role in entropy production. Equation (28) shows local entropy production rate in plug
flow reactor (Johannessen and Kjelstrup, 2004).

σ = Ωρ B ∑  rj ( −∆ r G j / T )  + π DJ q ∆ (1 / T ) + Ων  − (1 / T )( dP /dz )  . (28)
j

The total entropy production rate is


L
( dS /dt )irr = ∫0 σ dz. (29)
480 S.C. Nimkar and R.K. Mewada

Figure 6 Sections in dividing wall column (see online version for colours)

Source: Sumiju Plant Engineering Co., Ltd (2010)

Overall exergy balance in the reactor is shown in Figure 8. Second law efficiency
represents internal exergy losses and intrinsic Exergy efficiency depends on exergy losses
per unit of produced exergy. Hence improvement of the system is possible by minimising
the exergy losses per unit of produced exergy (Sorin et al., 1998b).
The entropy production rate is maximum at inlet due to high reaction rate and high
heat flux in the tubular fixed bed reactor used for the oxidation of SO2. Entropy
production can be minimised to 10.47% by external cooling. If rector length is reduced
by 19% from original, reduction in entropy is 24.7% (Johannessen and Kjelstrup, 2004).
Irreversible thermodynamics can find out the location and cause of exergy loss during
diffusion and chemical reaction (Kjelstrup and Arons, 1995). It will be helpful for the
analysis of reactor and to improve its efficiency. Exergy losses are more in the reactor
in the production of methanol by partial oxidation of methane (Smith et al., 2002).
Oxidation of solid carbon at atmospheric pressure with air is analysed by Prins and
Ptasinski (2005). The Exergy efficiency of air blown gasification is more than oxygen
An overview of exergy analysis for chemical process industries 481

blown gasification. Though total relative losses are more in the air blown gasification,
exergy loss for separating oxygen from air caused less exergy efficiency. If gasifier
operated at 927°C to 1127°C at atmospheric pressure, 75% exergy of fuel can be
converted into chemical exergy of CO and H2. Membrane reactors are becoming more
promising compared with traditional reactors on the basis of exergy efficiency and
also easier temperature control, improved safety and more utilisation of raw material
(Bernardo et al., 2006). Exergy efficiency of Membrane reactor used for ethylene
production is 1.3 times higher (80.9%) than traditional reactor (69.9%).

Figure 7 Changes in chemical and physical exergy during an overall chemical reaction

Source: Hinderink et al. (1996a)

Figure 8 Graphical presentation of exergy balance in reactor

Source: Sorin et al. (1998a)


482 S.C. Nimkar and R.K. Mewada

4 Exergy analysis of inorganic chemical industry

4.1 Ammonia production


Ammonia is produced by reacting nitrogen and hydrogen using Haber and Bosch process.
The basic components are ammonia reactor, compressors, turbines and ammonia
condensation system. Degree of perfection in ammonia plant is 56.11% reported by
Szargut et al. (1988) and 68.83% reported by Radgen (1997). Exergy analysis of different
ammonia synthesis loops is carried out by Kirova-Yordanova (2004). The various
ammonia synthesis loop shows an exergy degree of perfection in the range
of 0.901 to 0.931. Major losses are in the synthesis reaction and heat exchanger.
Exergy losses in synthesis reactor are due to irreversibilities in reaction as well as the
heat exchange in the reactor itself. Intercooler after first compression stage contributes
major exergy loss in compression stage. The hydrogen used in ammonia production by
reforming process is also having exergy loss due to chemical reaction in reformer
(Boyano et al., 2011; Simpson and Lutz, 2007; Cruz and Oliveira, 2008), which reduces
overall exergy efficiency of the plant. Exergy losses can be reduced by using reaction
heat for HP steam generation, multistage ammonia condensation. In refrigeration system
losses can be reduced by optimising temperature and pressure in the cycle. 15% reduction
in exergy loss is possible by slightly increasing temperature and pressure in refrigeration
cycle (Panjeshahi et al., 2008).

4.2 Nitric acid production


Nitric acid is produced by oxidation of ammonia. This plant is a net exporter of energy.
The heat produced during oxidation is recovered by using expander and turbine to run
turbocompressor. The overall exergy efficiency of various nitric acid plants are between
17 and 33% (Szargut et al., 1988; Kirova-Yordanova et al., 1994; Gaggioli et al., 1991).
Most of the exergy loss is in ammonia oxidiser followed by waste heat boiler. Exergy
analysis of nitric acid plant is shown in Table 3. Exergy efficiency of heat exchanger
network of nitric acid plant is 48.43% (Belghaieb et al., 2010). High temperature
gas/gas exchange exergy loss is 45% and low temperature gas/gas exchange is 24%.
Exergy destruction in various equipments of two dual pressure nitric acid plants with
different NOx treatment methods is shown in Table 4. Cooling tower takes out the
heat from cooler condenser and absorber. If this heat is recovered, exergy efficiency
can be increased. Power output in plant can be increased by 34% by applying Sama’s
13 common sense guidelines to detect second law errors suggested (Živković et al., 2004;
Laković et al., 2005). Input chemical exergy in ammonia SOFC cell can be used to
produce nitric oxide and electricity simultaneously (Farr, 1979).

4.3 Sulphuric acid production


Exergy efficiencies of sulphuric acid production reported by various authors by using
different processes are
• contact process – 40.44% (Szargut et al., 1988), 55.15% (Rasheva and Atanasova,
2002), 47.20% (Kotas, 1995) and 59.4% (Chouaibi et al., 2013)
An overview of exergy analysis for chemical process industries 483

• wet catalyst – 36.7% (Almirall, 2009)


• pyrometallurgical process – 36% (Patronov and Magaeva, 2007; Magaeva et al.,
2000).
A large amount of heat is produced in the sulphuric acid manufacturing process. Exergy
efficiency depends upon effective heat recovery. Combustion and sulphur trioxide
absorption are the sections where exergy losses are more. The losses are due to
irreversibilities in chemical reactions, heating and mixing different temperature,
composition streams. The exergy balance of the typical DCDA process is tabulated in
Table 3.
Exergy losses can be reduced by increasing the temperature of incoming air in burner,
increasing steam temperature and increasing pressure in the burner (Rasheva and
Atanasova, 2002). Exergy losses in the absorbers can be decreased through the
use of a hot absorption system and the devices for cooling of the acid could be used
for heating of the supplied water. But optimum temperature is required as absorption
efficiency will decrease with increasing temperature. An Organic Rankine Cycle can be
used to recover low temperature heat, if it finds economic feasibility (Szargut et al.,
1988).

Table 3 Exergy balance of inorganic acid production

Exergy
efficiency
Product Input exergy (%) Output exergy (%) (%) Reference
Sulphuric acid Sulphur (97.02) Sulphuric acid (25.97) 40.44 Szargut et al.
from sulphur Electric power (2.98) Export steam (14.55) (1988)

Stack gas (0.55)


Internal exergy loss
(58.93)
Sulphuric acid Sulphur (93.24) Sulphuric acid (26.58) 55.15 Rasheva and
from liquid Chemically purified Oleum (7.17) Atanasova (2002)
sulphur water (2.82)
Electric power (3.94) Power from steam
(21.40)
External exergy loss
(14.20)
Internal exergy loss
(30.65)
Sulphuric acid Sulphur dioxide Sulfuric Acid (49.29) 49.29 Atanasova (2010)
from SO2 (90.83)
Chemically purified External exergy loss
water (0.9288) (13.06)
Electric power (8.24) Internal exergy loss
(37.65)
484 S.C. Nimkar and R.K. Mewada

Table 3 Exergy balance of inorganic acid production (continued)

Exergy
efficiency
Product Input exergy (%) Output exergy (%) (%) Reference
Phosphoric acid Phosphate rock Phosphoric acid (17.34) 32.43 Atanasova (2010)
(wet process) (38.45)
Sulphuric acid (39.66) Phosphogypsum (15.08)
Fresh water (7.76) External exergy loss
(10.68)
Steam (9.44) Internal exergy loss
(56.9)
Electric power (4.69)
Nitric acid (dual Ammonia (87.43) Nitric acid (7.66) 30.53 Szargut et al.,
pressure with Water (0.11) Steam (22.87) (1988)
extended
absorption) Electricity (12.46) External exergy loss
(55.82)
Internal exergy loss
(13.65)
Nitric acid (dual Ammonia (100) Nitric acid (13.7) 19.66 Gaggioli et al.
pressure with Electricity (5.96) (1991)
extended
absorption) Exergy loss (80.34)
Nitric acid (dual Ammonia (95.36) Nitric acid (12.92) 17.46 Kirova-Yordanova
pressure with Water (1.62) Steam (4.54) et al. (1994)
extended
absorption) Air (0.13) External exergy loss (72)
Steam (2.17) Internal exergy loss
(10.54)
Electricity (0.72)
Nitric acid (dual Ammonia (61.94) Nitric acid (8.16) 28.30 Kirova-Yordanova
pressure with Water (4.51) Steam (20.14) et al. (1994)
catalytic high
temp reduction Air (0.05) External exergy loss
of NOx) (60.93)
Natural gas (33.50) Internal exergy loss
(10.77)
Electricity (0.72)

4.4 Phosphoric acid production


Phosphoric acid is produced by wet process using phosphate rock and sulphuric acid as a
raw material. Phosphogypsum is obtained as a by-product along with phosphoric acid.
Phosphogypsum dumping is causing serious ecological problem. But decomposition of
phosphogypsum into SO2 and lime can takes place in an eco-friendly way. The main
reaction of thermochemical decomposition of phosphogypsum is
0.25CH4 + CaSO4 → CaO + SO2+ 0.25CO2 + 0.5H2O
An overview of exergy analysis for chemical process industries 485

Table 4 Exergy destruction in nitric acid plant

% Exergy destruction
Dual pressure process
Dual pressure process (Catalytic high temp
Equipment (Extended absorption) reduction of NOx)
Reactor 37.41 31.00
Compressor and turbine 19.71 10.31
Absorber and bleacher 4.38 1.57
Adiabatic reactor 4.06 NA
Air–Ammonia mixer 4.01 5.49
Separator (gas and acid) 1.28 –
Boiler 13.54 –
Heat exchangers 13.26 18.66
Tail gas treatment – 8.46
Furnace and heat exchanger – 16.45
Other (pump, de aerator etc.) 2.35 8.06
Total 100 100
Source: Kirova et al. (1994)

Sulphuric acid produced from SO2 can be used in the process again. Exergy analysis of
the whole system is studied by Atanasova (2010). Exergy efficiency of phosphoric acid
production is found to be 32.43% without considering decomposition of phosphogypsum.
For thermochemical decomposition natural gas is burned to provide heat for the reaction
with 49% exergy efficiency (Atanasova, 2002). Exergy efficiency in increased to 55.6%
by increasing O2 enriched air for oxidation and by using new heat exchanger system for
better heat utilisation. By increasing O2 content to 95%, exergy efficiency of oven
decrease to 55.3% due to extra energy required for the production of oxygen. Reduction
gas is produced by reacting steam with methane and then sending it oven for
decomposition of phosphogypsum. Exergy efficiency is increased to 59% and natural gas
consumption is decreased by 2.8 times. Reactions are
CH4 + H2O → CO + 3H2,
CaSO4 + H2 → CaO + SO2 + H2O,
CaSO4 + CO → CaO + SO2 + CO2.
Sulphuric acid production by decomposition of gypsum reduces exergy efficiency of wet
phosphoric acid production process to 20.86%. Table 3 shows the exergy balance of wet
phosphoric acid production with decomposition of gypsum. The reason for the decrease
in efficiency is the use of natural gas and air separation process. Production of sulphuric
acid by using sulphur is more efficient than the decomposition of phosphogypsum to
produce sulphuric acid (Atanasova, 2002).
486 S.C. Nimkar and R.K. Mewada

5 Exergy analysis of organic Chemical industry

5.1 Exergy analysis of Refinery


A crude oil is a complex mixture of identified and undefined substances. The chemical
exergy of pseudocomponents can be obtained from its elemental composition and heating
value (equation (30)) (Rivero et al., 1999)
eqiCH = NHVi β i + ∑ z j eqjCH , (30)

where zj is the mass fractions of metals, Fe, Ni, V, and water in the pseudo-component
and eqj is their corresponding specific standard chemical exergies, obtained from the
standard chemical exergy values. NHVj is the net heating value of the pseudo-component
j and βj is the chemical exergy correction factor as a function of its C, H2, O2, S and N2,
mass fractions. A correlation for calculation of chemical exergy of petroleum fractions
based on enthalpies of combustion is developed by Govin et al. (2000). Zhu (2008)
derived equation for calculation of cumulative exergy rate on the basis of the minimum
separation power of product (MSPP). The MSPP of the kth stream (Gk) based on the
concept of exergy is given by
l
MSPPK = ∑ n Fj e Fj – nkP ekP – nkPP ekPP , (31)
j =1

l
MSPPK = ∑n (H
j =1
F
j
F
j − T0 S Fj ) – nkP ( H kP − T0 SkP ) – nkPP ( H kPP − T0 SkPP ) . (32)

Value of allocation of annual total cumulative exergy rate of kth product is


m
E A, k = Gk ∑ G (∑ E
i =1
k
cr
C + ECop ) . (33)

Crude oil is taken out from the oil well using platform before sending it to refinery.
Exergy analysis of two different oil platforms shows exergy efficiency of 9.7%
(De Oliveira and Van Hombeeck, 1997) and 13% (Voldsund et al., 2013). As both
platforms are not identical, change in some processes like heating, separation,
compression causes a change in exergy efficiency. Recompression and reinjection of gas
stream are the processes causing most of the exergy destruction followed by production
manifold. Recompression and reinjection of gas is required to maintain pressure in
reservoir. Available gas from the well is used to fulfil the energy requirement of all the
operations used in the process. About 21–27% of input exergy is converted into useful
power on the platform for production process (Nguyen et al., 2012).
Crude oil from the platform is taken to oil refineries where it is fractionated
using distillation columns to get different products. Rivero (2002) conducted an exergy
analysis of refineries in IMP Mexico and implemented various projects to reduce exergy
losses. The Exergy diagram for the combined distillation unit of the refinery is shown in
Figure 9. Effluent exergy losses are 11.32% and irreversible exergy losses are 88.68%
(Rivero et al., 2004b).
An overview of exergy analysis for chemical process industries 487

Figure 9 Grassman diagram of the combined distillation unit in refinery

Source: Rivero et al. (2004)

Al-Muslim et al. (2005) compared single stage and two stage crude oil distillation column
for minimisation of exergy loss. Increasing reference temperature, decreases exergy
efficiency and increases exergy loss in both one and two stage crude oil distillation.
Exergy losses observed in the furnace are more than in column. In two stage distillation,
exergy efficiency is drastically increased from 14% to 31.5%. In another method single
flash is installed before distillation column to separate lighter components before the feed
enters the furnace (Benali et al., 2012). The unnecessary presence of light components
causes their overheating; hence they are directly fed to the top of the column. Exergy
destruction is reduced by 14% after installation of flash. Divided wall column (DWC)
will be a good option to replace these systems and reduce energy and capital cost. Two
distillation columns can be combined in one to avoid high temperature difference at the
top and bottom. A number of crude distillation units for optimum exergy utilisation will
depend upon the method used for heat integration. The addition of one or more vacuum
distillation unit to a crude distillation unit may save more energy (Li et al., 2013).
Fluidised catalytic cracking (FCC) is used to convert heavy petroleum components
into light components, which are commercially important. FCC unit is optimised by using
the exergy production rate as the objective function (Souza et al., 2011). Geometric
parameters of riser and operating conditions of the unit can be used for maximisation of
thermodynamic performance. Exergy destruction observed in FCC unit is 1255.71 MJ/ton
reported by Song et al. (1999). Energy saving in FCC unit can be achieved by improving
the recovery and utility ratio of the flue gas expander and reducing the exhaust
temperature of the waste heat boiler. Catalyst to oil ratio (COR) is also an important
parameter to be considered while optimisation.
488 S.C. Nimkar and R.K. Mewada

5.2 Exergy analysis of Petrochemical process


Petrochemicals are basic building blocks for most of the products in today’s
world. Fraction obtained from the refinery is used as raw material for petrochemical
complex. Vinyl chloride is produced using ethylene and chlorine. In the first
step 1–2 dichloroethane (EDC) is produced by reacting both reactants. Exergy loss in the
VC plant is divided into five types (Graveland and Gisolf, 1998):
• internal losses (units and equipments).
• external losses (emission to environment).
• utility generation losses.
• primary energy losses (primary fuel consumption).
• cumulative exergy losses (production process leading from natural resources to final
product).
Figure 10 shows exergy balance of vinyl chloride monomer plant. Only 15% of the input
exergy is transformed into work. Exergy destruction can be reduced by replacing the
quench tower by an Organic Rankine cycle and the production of electricity by using heat
of cracking reaction gas. Improving sequence of distillation and direct compression of
EDC vapours instead of condensation before sending to cracking furnace will also
improve exergy efficiency. Adiabatic compressor can be used to compress vapours
leaving the top of the EDC purifying column and vapours are used to heat up the bottom
product in the reboiler (Araújo et al., 2007). The irreversibility rate distribution for the
original and modified column with compressor at different temperature differences is
shown in Figure 11. Reboiler irreversibility increases marginally for temperature
difference more than 10°C and becomes almost double at a difference of 30°C.
Irreversibility in the condenser is reduced substantially in the new configuration than
original configuration and it improves the thermodynamic efficiency of the column.

Figure 10 Exergy balance of vinyl chloride monomer (see online version for colours)

Source: Chemical Bandwidth Study (2004)


An overview of exergy analysis for chemical process industries 489

Figure 11 Irreversibility rate distribution in VCM unit

Source: Araújo et al. (2007)

Exergy analysis of methanol production shows that convective reforming process is 40%
more exergy efficient process than the steam reforming process (Rosen and Scott, 1988).
The refrigeration cycle in petrochemical plant that uses ethylene and propylene as
refrigerant shows more losses in compressor and evaporator (Fábrega et al., 2010).
Maximum loss of ethylene cycle is in mixer due to uncontrolled mixing of streams. In the
propylene cycle major exergy destruction is in the compressor and then in the evaporator.
Almost 13% of exergy loss can be reduced by reducing differences in temperature and
pressure of different streams. Exergetic efficiency of propane refrigerant system used in
the C2+ recovery plant of Marun Petrochemical Company, Iran is found to be 43.45%.
Analysis indicates a potential for reduction of exergy loss in heat exchanger and the
expansion section (Tirandazi et al., 2011). Energy saving in olefin plant is possible by
using various retrofits combinations. Ethylene separation at low temperature in this plant
needs different refrigeration levels. Three modification schemes are studied based on
exergy analysis to find optimum scheme. The first scheme involves modification of
column parameters, in second only refrigeration cycles are optimised and in third scheme
all parameters in plant are optimised using a genetic algorithm (Tahouni et al., 2013).
Highest energy saving is achieved in third scheme with the energy saving cost of a
4M$/y.
Acetylene is produced by partial oxidation of natural gas. Acetylene reactor and
quenching sections are major exergy destructors. Exergy analysis of reactor shows that
when oxygen to methane ratio raises exergy loss increases. The Exergy efficiency of the
reactor can be increased by adjusting the molar ratio of oxygen to methane and carbon
monoxide to the methane (Zhifang and Danxing, 2008). A quenching oil auxiliary system
to recover the reaction heat of product gas, membrane separation technology to separate
H2 and recycling some CO to the feedstock can reduce exergy losses by 19% and
increases the C2H2 yield. Polygeneration system which produces acetylene and electricity
simultaneously will also improve exergy efficiency (Wang et al., 2009). It consists of
490 S.C. Nimkar and R.K. Mewada

partial oxidation/combustion unit, an acetylene separation unit, a water gas shift unit and
H2/O2 cycle unit. Hydrogen produced in reforming helps to improve exergy efficiency.
In this system exergy loss is only half of partial oxidation/combustion system.
Propane dehydrogenation process mainly consists of a reactor, condenser-separator
and series of distillation columns along with mixer, heat exchangers and compressor
expanders. Cryogenic conditions are required for all condensers, i.e. for separation of
reactor product in to two phases and at the distillation column top. Maximum entropy
production takes place in a reactor and two columns (Røsjorde et al., 2007). Entropy
generation can be minimised by increasing the amount of recycle in rector, selectivity and
pressure in separator.
The entropy production rate of separation function is found to be more than reaction
function in MTBE production unit (Almeida-Rivera and Grievink, 2007). A typical
MTBE column in FCC unit shows exergy losses of 63%. By changing feed location,
exergy losses can be reduced which results in saving reboiler duty (Rivero and Garcia,
2001). Cracking, quenching and separation are the main sections for the exergy loss in
various petrochemical plants. Reaction temperature and feed composition affect exergy
loss. Increase in temperature causes more exergy loss but it is necessary to keep the high
temperature for cracking reaction in reforming process (Portha et al., 2010). Exergy
diagram of the naphtha reforming process is shown in Figure 12. Table 5 shows the
exergy loss in various sections of the major petrochemical processes.

Table 5 Exergy loss (%) in major petrochemical processes

Major candidates for exergy loss External exergy Internal exergy


Product (% share in total loss) loss (%) loss (%)
Ethylene from propane Demethanisation (32) 78.00 22.00
Cracking and quenching (21) 3.50 96.50
Heat recovery and refrigeration (16) 64.35 35.65
Ethylene from naphtha Cracking and quenching (39) 9.83 90.17
Demethanisation (20) 63.03 36.97
Compression and deacidification (16) 27.80 72.20
Ammonia from natural Preheating and reforming (38) 13.93 86.07
gas Ammonia synthesis (35) 19.92 80.08
Gas Upgrading (20) 77.52 22.48
Ethylene oxide by Shell Reactor (33) 0 100
process EO stripper (32) 6.95 93.05
EO purification (19) 75.00 25.00
Propylene from naphtha Reactor (72) 1.36 98.64
Product separation (23) 39.62 60.38
Terephthalic acid from Oxidation (73) 15.30 84.70
Paraxylene Crystalliser (10) 51.33 48.67
Purification (10) 82.31 17.69
An overview of exergy analysis for chemical process industries 491

Table 5 Exergy loss (%) in major petrochemical processes (continued)

Major candidates for exergy loss External exergy Internal exergy


Product (% share in total loss) loss (%) loss (%)
Methyl tertiarybutyl ether MTBE recovery (93) 54.33 45.67
(MTBE)
Formaldehyde Reactor (88.9) 16.50 83.50
Absorber (7.07) 0 100
Methanol from natural gas Reforming (52) 0.28 99.72
(ICI LP) Refining (38) 60.00 40.00
Methanol from natural gas Reforming (55) 22.19 77.81
(Lurgi) Heat recovery (20) 90.00 10.00
Synthesis (19) 49.06 50.94
Acrylonitrile by Propylene Propyleneammoxidation (51) 7.70 92.30
ammoxidation Heat recovery and refrigeration 99.79 0.21
(27)
Acrylonitrile Separation (22) 20.16 79.84
Acrylonitrile by Propane Propane ammoxidation (70) 3.30 96.70
ammoxidation Heat recovery and refrigeration 99.57 0.43
(16)
Acrylonitrile Separation (14) 21.90 78.10
Styrene by UOP Process Air Cooler (39) 74.18 25.82
Feed preheater and reactor (27) 0 100
Steam super heater (14) 0 100
Styrene by Fina Process Feed preheater and reactor (51) 0 100
EB/Styrene Stripper (20) 64.86 35.14
Air Cooler (20) 76.70 23.30
Ethylbenzene by Mobil Benzene Fractionator (80) 73.92 26.08
Process
Ethylbenzene by Lummus Benzene Fractionator (46) 79.14 20.86
Process Primary reactor (40) 0 100
Nitric Acid from Ammonia Reaction (59) 7.74 92.26
oxidation Heat recovery (27) 95.11 4.89
Paraxyle from C8 isomers Isomerisation (70) 22.55 77.45
Fractionation (30) 53.40 46.60
CO2 recovery with MEA CO2 absorber (52) 1.65 98.35
CO2 stripper (23) 80.28 19.72
Vinyl Chloride via gas phase Dehydrogenation reaction (43) 0 100
pyrolysis Pre heater (17) 0 100
EDC recovery (16) 76.70 23.30
Acetic Acid from Methanol Carbonylation (57) 2.58 97.42
Acetic acid refining (43) 25.57 74.43
492 S.C. Nimkar and R.K. Mewada

Table 5 Exergy loss (%) in major petrochemical processes (continued)

Major candidates for exergy loss External exergy Internal exergy


Product (% share in total loss) loss (%) loss (%)
Butadiene form C4streams Extractive distillation (59) 5.34 94.66
Feed vapouriser and DMF 10.22 89.78
cooling (21)
Conventional Distillation (20) 25.60 74.40
Cumene via Propylene Cumene Recovery (71) 46.15 53.85
alkylation (SPAcatalysed) Alkylation (27) 0 100
Carmen via Propylene Cumene Recovery (60) 64.34 35.66
alkylation (Zeolite catalysed) Alkylation (40) 25.60 74.40
Source: Chemical Bandwidth Study (2004)

Figure 12 Exergy diagram of naphtha reforming process (see online version for colours)

Source: Portha et al. (2010)

6 Exergy analysis for process improvement

Various efforts are carried out to perform exergy analysis by developing new
software and also by using existing simulation packages (Querol et al., 2011; Hinderink
et al., 1996b; Abdollahi-Demneh et al., 2011; Graveland, 1999; Ghannadzadeh et al.,
2012). Exergy analysis is also helpful while designing a new process. It is not only to
reduce energy losses but also to reduce capital cost. Different ways for the reduction of
irreversibilities in various process equipments are shown in Table 6. If a cold stream is
heated by high pressure steam, then its availability will cause ∆T hence reduces surface
area. It will be saving in capital cost of heat exchanger. Thirteen common sense second
law guidelines suggested by Sama et al. (1989) will be helpful to reduce second law error
which is the main reason for the failure of exergy analysis. Designer’s skill is more
important to get exergy efficient process rather than relying only on software (Sama,
1995b). The cost associated with the equipment in the process is fixed cost and running
cost. The exergy destruction cost is the major running cost of equipment. Equations were
developed for costing and performance by El-Sayed (2002).
An overview of exergy analysis for chemical process industries 493

Table 6 Internal exergy loss in unit operations and its sources

Unit operation Sources of irreversibility Improvement ways


Reactor Low conversion Recycle the non-converted feed
Exothermic reaction Raise the temperature
Endothermic reaction Reduce the temperature
Temperature difference of cold feed and Pre heating of feed
hot reaction medium
Concentration gradients Increase reaction stages as much as
possible
Mixing of streams Mixing as uniform as possible
Distillation column Concentration gradients Use intermediate reboiler
Equal partition of driving force
Improper separation sequence Optimise distillation sequencing
Pressure drop and mechanical friction Optimise hydraulics of column
Bubble-liquid mass transfer on the tray Optimise hydraulics of column
Thermal gradients Introduce feed in proper tray
Heat exchanger Temperature difference Use as low as possible driving
force
Non uniform gradient Use as uniform gradients
Pressure drop Reduce number of baffles
Low heat transfer Optimise flow velocity
Cold utility Refrigeration Minimise use of sub ambient
system and replace it with cooling
water
Thermal difference Use high level as possible
Use of external utilities Maximise process steam
generation
Throttling valve Pressure drop Replace by steam turbine (for
temperature greater than ambient)
Steam boiler A chemical reaction for oxidation of the Preheating of combustion air
fuel
An internal heat transfer between high Use as low driving force as
temperature product and unburned possible
reactant
A physical mixing process Mix it as uniform as possible
A diffusion process where the fuel and Make it as gradually as possible
oxygen molecules are drawn together
High heat capacity of combustion Oxygen enrichment
products
Isobar combustion Isochoric combustion
Compressor Hot inlet streams Temperature reduction of inlet
streams or between the stages by
intercooler
494 S.C. Nimkar and R.K. Mewada

Table 6 Internal exergy loss in unit operations and its sources (continued)

Unit operation Sources of irreversibility Improvement ways


Steam turbine Low temperature of steam Use inter-heater (e.g. super-heater)
between the stages
Pump Hydraulic friction Optimise the hydraulic of system
Mixer Temperature difference Isothermal mixing
Pressure difference Isobar mixing
Composition difference Mixing as close as possible
composition
Source: Ghannadzadeh et al. (2012)

6.1 Pinch technology and exergy


Exergy analysis becomes important tool for process optimisation and synthesis. The
combination of pinch technology and exergy analysis helps for improvement in process
(Rücker and Gruhn, 1999; Feng and Zhu, 1997; Sorin and Paris, 1997). A comparative
study of both methods is carried out by Wall and Ging (1996). The composite curves in
pinch analysis (T-H diagram) for a heat transfer system can be converted into the exergy
composite curves and the grand composite curve as shown in Figure 13. Exergy losses
associated with the process are shown by the shaded area. Heat exchanger network in the
aromatic complex is revamped by Manan and Hain (2000) using exergy and pinch
analysis. Energy improvement potential for hot and cold utility is found to be 60% and
20%, respectively. Five column methanol distillation scheme reduces exergy loss by
21.5% compared with traditional four column scheme by doing heat integration with the
aid of exergy and pinch analysis (Sun et al., 2012). The Exergy load distribution method
developed by Sorin and Paris (1999) improves nearly optimised process by changing few
parameters.
In ammonia production process, pinch analysis (Radgen, 1996) and exergy analysis
(Radgen, 1997) shows almost same optimisation potential. Sama (1995b) has claimed
that exergy analysis can easily identify and correct the design errors in heat exchanger
network compared with pinch technology. It also provides an understanding of the
process which is absent in pinch technology. A challenge problem for the integration of
new process into existing site using exergy analysis (Gaggioli et al., 1991) vs. pinch
analysis (Linnhoff and Alanis, 1991) shows three times more energy saving in exergy
analysis than pinch technology (Sama,1995a).

6.2 Process synthesis


Energy efficiency, waste minimisation and effective raw material use can be measured
and optimised by using utilisable exergy coefficient in chemical process design. Exergy
load distribution analysis carried out for final optimal flowsheet topology shows increase
in exergy efficiency by 10% (Sorin et al., 2000a, 2000b). Second law concept is used to
reduce energy consumption in the chemical industry by following some consequences
suggested by Leites et al. (2003).
The counteraction principle shown in Figure 14 is the Second Law requirement to
make maximum use of the Gibbs energy of chemical reactions. The driving force method
An overview of exergy analysis for chemical process industries 495

is used to reduce thermodynamic irreversibility by developing engineering methods to


reduce driving force. The quasi-static method examines the process conditions of a
theoretical limiting quasi-static process that could be run with zero driving force along
the equilibrium line. Some of the commandments suggested based on second law of
thermodynamics are as follows (Leites et al., 2003):
• Retard the process at the start and then gradually remove retarding function.
• For an exothermic reaction increase the temperature and for an endothermic reaction
decrease the temperature. It is better to conduct the exothermic processes in a low
flow-heat-capacity medium.
• If volume increases in gas phase reaction it is necessary to raise the pressure and vice
versa.
• It is not necessary to carry out chemical reactions up to their completion. It is better
to recycle the unreacted streams.
• Do not mix streams of different temperatures, different compositions or different
pressures.
• The best chemical reactor is a counter-current one with plug flow.

Figure 13 Exergy transformation from CC to ECC and EGCC

Source: Feng and Zhu (1997)

6.3 Six Sigma and exergy analysis


Six Sigma is used to improve quality of process output by measuring defects and
reducing variation. It is based on methodology of DMAIC, i.e., Define, Measure,
Analyse, Improve and Control. Energy consumption in distillation unit of the naphtha
reforming process is reduced by Six Sigma method (Falcón et al., 2012). Fourteen critical
inputs are identified and analysed to get optimised parameters to save energy worth
150,000 Euro per year. Six Sigma methodology is used for re-evaluation of existing
process at BASF’s process development on the basis of exergy analysis (Asprion et al.,
2011). Process parameters are measured and then analysed by using pinch and exergy.
After implementing the measures expected improvement is achieved (Figure 15) and
~7 million euros are saved.
496 S.C. Nimkar and R.K. Mewada

Figure 14 Some consequences of the second law and methods of analysis based on them

Source: Leites et al. (2003)

Figure 15 Work flow for the re-evaluation of existing processes (see online version for colours)

Source: Asprion et al. (2011)

7 Exergy and pollution control

Exergy analysis can be used to design an industrial system based on ecology where one
industry’s waste can be input for other to design zero discharge processes (Granovskii
et al., 2008; Bakshi and Fiksel, 2003; Connelly and Koshland, 2001). A thermodynamic
framework for ecologically conscious process helps to reduce environmental impact
(Bakshi, 2002). Exergy losses should be minimised to obtain sustainable development
An overview of exergy analysis for chemical process industries 497

(Stougie and van der Kooi, 2009). A review of the exergy application to environmental
system is provided by Dewulf et al. (2008). Pollution in the atmosphere is due the
concentration of chemical species present and their chemical exergy. Rosen and Dincer
(1999) presented environment pollution cost and removal pollution cost based on exergy
analysis for selected fossil fuels.
Ammonia and ammonium nitrate are present in the effluents of ammonium nitrate
production plant. Recycling of both is important to reduce their impact on the
environment. Exergy analysis of pollution control method shows that the two stage
evaporation system reduces exergy consumption of the process by 52.291 MJ/t
(Kirova-Yordanova, 2010). Tail gas treatment in nitric acid production is done by either
extended absorption or reduction of NOx using ammonia (SCR) or reduction of NOx
using natural gas (NSCR). Exergy input is higher when natural gas is used. Nitric acid
plant with NSCR method is having higher exergy efficiency due to more availability of
steam from natural gas. Exergy-based indices studied by Kirova-Yordanova (2011)
shows that extended absorption is an environmental friendly method for tail gas
treatment. Another study of waste treatment in nitrogenous fertiliser factory shows
combined physiochemical and biological treatment not only increase exergy efficiency
but also reduces environmental impact (Jorquera et al., 2013). Use of the chemical heat
pump and reverse osmosis along with anaerobic ammonium oxidation process shows
67.20% exergy efficiency. The chemical heat pump can recover heat that may be used to
generate medium pressure steam and concentrate from reverse osmosis can be directly
used as fertiliser.

8 Obstacles for exergy analysis

Fundamentals of the second law of thermodynamics are vital to understand exergy


analysis. It is comparatively difficult than energy analysis. The reference environment in
exergy analysis considered by various authors to render the technique is too challenging
(Rosen, 2001, 2002). Education about exergy is not available in most of the educational
institutes and universities. Practicing engineers are not finding exergy methods
demonstrable to get results. Most of the industries are doing only energy analysis using
the first law of thermodynamics due to its practical easiness. Outcome of interview
conducted by Carl-Erik et al. (2011) shows that lack of expertise in exergy analysis, less
awareness about exergy, priority to other tasks are the obstacles for exergy analysis in the
chemical industry. Modifications suggested by exergy analysis to optimise the processes
are implemented if there is a good payback period or due to regulatory compliances.
Though more research papers are published on exergy analysis to find out irreversibilities
in process, the solution to reduce it with case studies will increase exergy use in industry.

9 Summary

From the overview of exergy analysis in chemical industries following points are
summarised:
• Exergy losses mainly are due to irreversibilities in chemical reactions, heating and
mixing different temperature, composition streams.
498 S.C. Nimkar and R.K. Mewada

• Exergy efficiency of inorganic industry is 29% and organic industry is 35% in the
USA. Equipments dealing with separation and heat transfer in chemical industries
are having lower exergy efficiency. Overall sectorial efficiency is low due to the
combination of efficiencies of all plants.
• Major losses occurred in power generation by waste heat in the plant followed by
distillation. Exothermic reactors are also a significant source of internal exergy
losses.
• Exergy losses in column can be reduced by reducing reflux ratio. Lower reflux ratio
reduces mixing of two streams with large temperature difference.
• Use of diabatic and dividing wall column over adiabatic column reduces exergy loss
by 35%.
• Entropy production in exothermic reactor is depending on the temperature and length
of the reactor. Exergy loss can be reduced by optimising both parameters.
• Crude oil distillation in refinery can become exergy efficient by installing flash
before the column. Conversion of multiple columns into DWC for crude oil
distillation is also a good option to avoid large temperature difference at the top and
bottom of the column.
• Irreversibilities in chemical reaction is a major reason for the exergy loss in sulphuric
acid and nitric acid production.
• Exergy methods can help to improve economic evaluations and environmental
assessments. Many design related applications of exergy analysis are helpful
to evaluate, compare, and improve the process. Exergy analysis shows
polygeneration systems for the production of energy and chemicals is promising in
the future (Gao et al., 2004).
• Six Sigma method is useful in the chemical industry to identify cost effective exergy
improvement projects.

References
Abdollahi-Demneh, F., Moosavian, M.A., Omidkhah, M.R. and Bahmanyar, H. (2011)
‘Calculating exergy in flowsheeting simulators: A HYSYS implementation’, Energy, Vol. 36,
pp.5320–5327.
Ahrendts, J. (1980) ‘Reference states’, Energy, Vol. 5, Nos. 8–9, pp.666–677.
Almeida-Rivera, C.P. and Grievink, J. (2007) ‘Process design approach for reactive distillation
based on economics, exergy, and responsiveness optimization’, Industrial & Engineering
Chemistry Research, Vol. 47, pp.51–65.
Almirall, B.X. (2009) Introduction to Wet Sulfuric Acid Plants Optimization Through
Exergoeconomics, Master Thesis, Technische University, Berlin, Germany.
Al-Muslim, H., Dincer, I. and Zubair, S.M. (2005) ‘Effect of reference state on exergy efficiencies
of one- and two-stage crude oil distillation plants’, International Journal of Thermal Sciences,
Vol. 44, pp.65–73.
Araújo, A.B., Brito, R.P. and Vasconcelos, L.S. (2007) ‘Exergetic analysis of distillation processes-
a case study’, Energy, Vol. 32, pp. 1185–1193.
An overview of exergy analysis for chemical process industries 499

Asprion, N., Rumpf, B. and Gritsch, A. (2011) ‘Work flow in process development for energy
efficient processes’, Applied Thermal Engineering, Vol. 31, pp.2067–2072.
Atanasova, L. (2002) ‘Exergy analysis of the process of thermal decomposition of phosphogypsum
to lime and sulfur dioxide’, International Journal of Applied Thermodynamics, Vol. 5,
pp.119–126.
Atanasova, L. (2010) ‘Exergy analysis of the process of wet-process phosphoric acid production
with full utilisation of sulphur contained in the waste phosphogypsum’, International Journal
of Exergy, Vol. 7, pp.678–692.
Ayotte-Sauvé, E. and Sorin, M. (2010) ‘Energy requirements of distillation: exergy, pinch
points, and the reversible column’, Industrial & Engineering Chemistry Research, Vol. 49,
pp.5439–5449.
Ayres, R.U., Talens Peiró, L. and Villalba Méndez, G. (2011) ‘Exergy efficiency in industry:
Where do we stand?’, Environmental Science & Technology, Vol. 45, pp.10634–10641.
Bakshi, B.R. (2002) ‘A thermodynamic framework for ecologically conscious process systems
engineering’, Computers & Chemical Engineering, Vol. 26, pp.269–282.
Bakshi, B.R. and Fiksel, J. (2003) ‘The quest for sustainability: challenges for process systems
engineering’, AIChE Journal, Vol. 49, pp.1350–1358.
Belghaieb, J., Dkhil, O., Elhajbelgacem, A., Hajji, N. and Labidi, J. (2010) ‘Energy optimization of
a network of exchangers-reactors in a nitric acid production plant’, Chemical Engineering
Transactions, Vol. 21, pp.271–276.
Benali, T., Tondeur, D. and Jaubert, J.N. (2012) ‘An improved crude oil atmospheric distillation
process for energy integration: Part I: Energy and exergy analyses of the process when a flash
is installed in the preheating train’, Applied Thermal Engineering, Vol. 32, pp.125–131.
Bernardo, P., Barbieri, G. and Drioli, E. (2006) ‘An exergetic analysis of membrane unit operations
integrated in the ethylene production cycle’, Chemical Engineering Research and Design,
Vol. 84, pp.405–411.
BoroumandJazi, G., Rismanchi, B. and Saidur, R. (2013) ‘A review on exergy analysis of industrial
sector’, Renewable and Sustainable Energy Reviews, Vol. 27, pp.198–203.
Boyano, A., Blanco-Marigorta, A.M., Morosuk, T. and Tsatsaronis, G. (2011)
‘Exergoenvironmental analysis of a steam methane reforming process for hydrogen
production’, Energy, Vol. 36, pp.2202–2214.
BP Energy Outlook 2030 (2011) http://www.bp.com/liveassets/bp_internet/globalbp/globalbp_
uk_english/reports_and_publications/statistical_energy_review_2008/STAGING/local_assets/
2010_downloads/2030_energy_outlook_booklet.pdf (Accessed on 12 February, 2012).
Brodyansky, V.M., Sorin, M.V. and Le Goff, P. (1994) The Efficiency of Industrial Processes:
Exergy Analysis and Optimization, Elsevier, Amsterdam, The Netherlands.
Carl-Erik, G., Elfgren, E., Söderström, M., Thollander, P., Berntsson, T., Åsblad, A. and Wang, C.
(2011) ‘Possibilities and problems in using exergy expressions in process integration’,
Paper presented at World Renewable Energy Congress, 8–13 May, 2011. Linkoping, Sweden.
Chang, H. and Chuang, S-C. (2003) ‘Process analysis using the concept of intrinsic and extrinsic
exergy losses’, Energy, Vol. 28, pp.1203–1228.
Chang, H. and Li, J. (2005) ‘A new exergy method for process analysis and optimization’,
Chemical Engineering Science, Vol. 60, pp.2771–2784.
Chemical Bandwidth Study (2004) Chemical Bandwidth Study, Exergy Analysis: A Powerful
Tool for Identifying Process Inefficiencies in the U.S. Chemical Industry,
http://www1.eere.energy.gov/manufacturing/resources/chemicals/pdfs/chemical_bandwidth_r
eport.pdf (Accessed on 1 February, 2012).
Chouaibia, F., Belghaiebb, J. and Hajji, N. (2013) ‘Assessment of energy performances of a
sulfuric acid production unit using exergy analysis’, International Journal of Current
Engineering and Technology, Vol. 3, No. 1, pp.149–157.
500 S.C. Nimkar and R.K. Mewada

Connelly, L. and Koshland, C.P. (2001) ‘Exergy and industrial ecology – Part 1: an exergy-based
definition of consumption and a thermodynamic interpretation of ecosystem evolution’,
Exergy, An International Journal, Vol. 1, pp.146–165.
Cruz, F.E. and Oliveira Jr., S.d. (2008) ‘Petroleum refinery hydrogen production unit: exergy and
production cost evaluation’, International Journal of Thermodynamics, Vol. 11, pp.187–193.
De Oliveira, S. and Van Hombeeck, M. (1997) ‘Exergy analysis of petroleum separation processes
in offshore platforms’, Energy Conversion and Management, Vol. 38, pp.1577–1584.
Dejanovića, I., Matijaševića, L. and Olujićb, Z. (2010) ‘Dividing wall column application for
platformate splitter – a case study’, ESCAPE 20: Proceedings of the 20th European
Symposium on Computer Aided Process Engineering, Napels Italy, pp.655–660.
Demirel, Y. (2006) ‘Retrofit of distillation columns using thermodynamic analysis’, Separation
Science and Technology, Vol. 41, pp.791–817.
Dewulf, J., Van Langenhove, H., Muys, B., Bruers, S., Bakshi, B.R., Grubb, G.F., Paulus, D.M.
and Sciubba, E. (2008) ‘Exergy: its potential and limitations in environmental science and
technology’, Environmental Science & Technology, Vol. 42, pp.2221–2232.
Dincer, I. and Cengel, Y. (2001) ‘Energy, entropy and exergy concepts and their roles in thermal
engineering’, Entropy, Vol. 3, pp.116–149.
Dincer, I. and Rosen, M.A. (2007) Exergy: Energy, Environment and Sustainable Development,
1st ed., Elsevier, Oxford.
El-Sayed, Y.M. (2002) ‘Application of exergy to design’, Energy Conversion and Management,
Vol. 43, pp.1165–1185.
Fábrega, F.M., Rossi, J.S. and d’Angelo, J.V.H. (2010) ‘Exergetic analysis of the refrigeration
system in ethylene and propylene production process’, Energy, Vol. 35, pp.1224–1231.
Falcón, R.G., Alonso, D.V., Fernández, L.M.G. and Pérez-Lombard, L. (2012) ‘Improving energy
efficiency in a naphtha reforming plant using Six Sigma methodology’, Fuel Processing
Technology, Vol. 103, pp.110–116.
Faria, S.H.B. and Zemp, R.J. (2005) ‘Using exergy loss profiles and enthalpy temperature profiles
for the evaluation of thermodynamic efficiency in distillation columns’, Engenharia Térmica
(Thermal Engineering), Vol. 4, pp.76–82.
Farr, R.D. (1979) The Higher Temperature Solid Electrolyte Ammonia Fuel Cell, MS Thesis,
Massachusetts Institute of Technology, Cambridge.
Feng, X. and Zhu, X.X. (1997) ‘Combining pinch and exergy analysis for process modifications’,
Applied Thermal Engineering, Vol. 17, pp.249–261.
Gaggioli, R.A. and Petit, P.J. (1977) ‘Use the second law first’, Chemtech, Vol. 7, pp.496–506.
Gaggioli, R.A., Sama, D.A., Qian, S. and El-Sayed, Y.M. (1991) ‘Integration of a new process into
an existing site: a case study in the application of exergy analysis’, Journal of Engineering for
Gas Turbines and Power, Vol. 113, No. 2, pp.170–183.
Gao, L., Jin, H., Liu, Z. and Zheng, D. (2004) ‘Exergy analysis of coal-based polygeneration
system for power and chemical production’, Energy, Vol. 29, pp.2359–2371.
Gaudreau, K., Fraser, R.A. and Murphy, S. (2012) ‘The characteristics of the exergy reference
environment and its implications for sustainability-based decision-making’, Energies, Vol. 5,
pp.2197–2213.
Ghannadzadeh, A., Thery-Hetreux, R., Baudouin, O., Baudet, P., Floquet, P. and Joulia, X. (2012)
‘General methodology for exergy balance in ProSimPlus® process simulator’, Energy,
Vol. 44, No. 1, pp.38–59.
Govin, O., Diky, V., Kabo, G. and Blokhin, A. (2000) ‘Evaluation of the chemical exergy of fuels
and petroleum fractions’, Journal of Thermal Analysis and Calorimetry, Vol. 62, pp.123–133.
Granovskii, M., Dincer, I. and Rosen, M.A. (2008) ‘Exergy and industrial ecology: an application
to an integrated energy system’, International Journal of Exergy, Vol. 5, pp.52–63.
An overview of exergy analysis for chemical process industries 501

Graveland, A.J.G.G. (1999) ‘Exan(TM) Pro: process visualization tool: increasing your insight into
energy conversion and large chemical processes’, Computers & Chemical Engineering,
Vol. 23, pp.S669–S672.
Graveland, A.J.G.G. and Gisolf, E. (1998) ‘Exergy analysis: an efficient tool for process
optimization and understanding. Demonstrated on the vinyl-chloride plant of Akzo Nobel’,
Computers & Chemical Engineering, Vol. 22, pp.S545–S552.
Hinderink, A.P., Kerkhof, F.P.J.M., Lie, A.B.K., De Swaan Arons, J. and Van Der Kooi, H.J.
(1996a) ‘Exergy analysis with a flowsheeting simulator-II. Application; synthesis gas
production from natural gas’, Chemical Engineering Science, Vol. 51, pp.4701–4715.
Hinderink, A.P., Kerkhof, F.P.J.M., Lie, A.B.K., De Swaan Arons, J. and Van Der Kooi, H.J.
(1996b) ‘Exergy analysis with a flowsheeting simulator-I. Theory; calculating exergies of
material streams’, Chemical Engineering Science, Vol. 51, pp.4693–4700.
Hirata, K. (2009) ‘Heat integration of distillation column’, Chemical Engineering Transactions,
Vol. 18, pp.39–44.
International Energy Outlook (2011) http://www.eia.gov/forecasts/ieo/pdf/0484(2011).pdf
(Accessed on 12 February, 2012).
Ishida, M. and Nakagawa, N. (1985) ‘Exergy analysis of a pervaporation system and its
combination with a distillation column based on an energy utilization diagram’, Journal of
Membrane Science, Vol. 24, No. 3, pp.271–283.
Isopescu, R., Woinaroschy, A. and Drãghiciu, L. (2008) ‘Energy reduction in a divided wall
distillation column’, REV CHIM, Vol. 59, pp.12–15.
Jimenez, E.S., Salamon, P., Rivero, R., Rendon, C., Hoffmann, K.H., Schaller, M. and
Andresen, B. (2004) ‘Optimization of a diabatic distillation column with sequential heat
exchangers’, Industrial & Engineering Chemistry Research, Vol. 43, pp.7566–7571.
Johannessen, E. and Kjelstrup, S. (2004) ‘Minimum entropy production rate in plug flow reactors:
an optimal control problem solved for SO2 oxidation’, Energy, Vol. 29, pp.2403–2423.
Jorquera, O., Kalid, R., Kiperstok, A., Braga, E. and Sales, E.A. (2013) ‘Effluent stream treatment
in a nitrogenous fertilizer factory: an exergy analysis for process integration’, Process Safety
and Environmental Protection, DOI :10.1016/j.psep.2013.07.003.
Khoa, T.D., Shuhaimi, M. and Nam, H.M. (2012) ‘Application of three dimensional exergy
analysis curves for absorption columns’, Energy, Vol. 37, pp.273–280.
Khoa, T.D., Shuhaimi, M., Hashim, H. and Panjeshahi, M. H. (2010) ‘Optimal design of distillation
column using three dimensional exergy analysis curves’, Energy, Vol. 35, pp.5309–5319.
Kirova-Yordanova, Z. (2004) ‘Exergy analysis of industrial ammonia synthesis’, Energy, Vol. 29,
pp.2373–2384.
Kirova-Yordanova, Z. (2010) ‘Application of the exergy method to environmental impact
estimation: the ammonium nitrate production as a case study’, Energy, Vol. 35, pp.3221–3229.
Kirova-Yordanova, Z. (2011) ‘Application of the exergy method to the environmental impact
estimation: the nitric acid production as a case study’, Energy, Vol. 36, pp.3733–3744.
Kirova-Yordanova, Z., Barakov, Y. and Koleva, D. (1994) ‘Exergy analysis of nitric acid plants: a
case study’, FLOWERS’94: Proceedings of the Florence World Energy Research Symposium,
Florence, Italy, pp.931–939.
Kiss, A.A. and Rewagad, R.R. (2011) ‘Energy efficient control of a BTX dividing-wall column’,
Computers and Chemical Engineering, Vol. 35, pp.2896–2904.
Kjelstrup, S. and Arons, R.J.D.S. (1995) ‘Denbigh revisited: reducing lost work in chemical
processes’, Chemical Engineering Science, Vol. 50, No. 10, pp.1551–1560.
Koeijer, d.G. and Rivero, R. (2003) ‘Entropy production and exergy loss in experimental
distillation columns’, Chemical Engineering Science, Vol. 58, pp.1587–1597.
Kotas, T.J. (1995) The Exergy Method of Thermal Plant Analysis, Krieger Publishing Company,
Malabar, Florida.
502 S.C. Nimkar and R.K. Mewada

Laković, M., Živković, P. and Rašković, P. (2005) ‘Exergy analyzing method in process integration
of the nitric acid production plant’, Facta Universitatis Series: Mechanical Engineering,
Vol. 13, pp.109–116.
Le Goff, P., Cachot, T. and Rivero, R. (1996) ‘Exergy analysis of distillation processes’, Chemical
Engineering Technology, Vol. 19, pp.478–485.
Leites, I.L., Sama, D.A. and Lior, N. (2003) ‘The theory and practice of energy saving in the
chemical industry: some methods for reducing thermodynamic irreversibility in chemical
technology processes’, Energy, Vol. 28, pp.55–97.
Li, Z., Lin, C., Wang, L. and Li, H. (2013) ‘Exergy analysis of multi-stage crude distillation units’,
Frontiers of Chemical Science and Engineering, Vol. 7, No. 4, pp.437–446.
Linnhoff, B. and Alanis, F.J. (1991) ‘Integration of a new process into an existing site: a case study
in the application of pinch technology’, Journal of Engineering for Gas Turbines and Power,
Vol. 113, No. 2, pp.159–168.
Luis, P. (2013) ‘Exergy as a tool for measuring process intensification in chemical engineering’,
Journal of Chemical Technology and Biotechnology, Vol. 88, No. 11, pp.1951–1958.
Magaeva, S., Patronov, G., Lenchev, A. and Grancharov, I. (2000) ‘Exergy analysis of processing
SO2 – containing gases in metallurgy into sulphuric acid and sulphur’, Journal of Mining and
Metallurgy, Vol. 36, pp.77–92.
Manan, Z.A. and Hain, F.S. (2000) ‘Combining pinch and exergy analysis for process improvement
– a case study on an aromatic complex’, Jurnal Teknologi, Vol. 33, pp.9–16.
Mustapha, D., Sabria, T. and Fatima, O. (2007) ‘Distillation of a complex mixture. Part II:
performance analysis of a distillation column using exergy’, Entropy, Vol. 9, pp.137–151.
Nguyen, T., Pierobon, L. and Elmegaard, B. (2012) ‘Exergy analysis of offshore processes on north
sea oil and gas platforms’, CPOTE 2012: Proceedings of the 3rd International Conference on
Contemporary Problems of Thermal Engineering, Gliwice, Poland.
Nummedal, L., Kjelstrup, S. and Costea, M. (2003) ‘Minimizing the entropy production rate of an
exothermic reactor with a constant heat-transfer coefficient: the ammonia reaction’, Industrial
& Engineering Chemistry Research, Vol. 42, pp.1044–1056.
Olakunle, M.S., Oluyemi, Z., Olawale, A.S. and Adefila, S.S. (2011) ‘Distillation operation
modification with exergy analysis’, Journal of Emerging Trends in Engineering and Applied
Sciences, Vol. 2, No. 1, pp.56–63.
Panjeshahi, M.H., Langeroudi, E.G. and Tahouni, N. (2008) ‘Retrofit of ammonia plant for
improving energy efficiency’, Energy, Vol. 33, pp.46–64.
Patronov, G. and Magaeva, S. (2007) ‘Exergoecological analysis of processing of SO2-containing
gases from zinc production’, Chemical Papers, Vol. 61, pp.457–463.
Pendergast, J.P., Vickery, D., Au-Yeung, P. and Anderson, J. (2008) Consider Dividing Wall
Columns, http://www.chemicalprocessing.com/articles/2008/245.html?page=full (Accessed on
10 February, 2012).
Portha, J.-F., Louret, S., Pons, M-N. and Jaubert, J-N. (2010) ‘Estimation of the environmental
impact of a petrochemical process using coupled LCA and exergy analysis’, Resources,
Conservation and Recycling, Vol. 54, pp.291–298.
Prins, M.J. and Ptasinski, K.J. (2005) ‘Energy and exergy analyses of the oxidation and gasification
of carbon’, Energy, Vol. 30, pp.982–1002.
Querol, E., Gonzalez-Regueral, B., Ramos, A. and Perez-Benedito, J.L. (2011) ‘Novel application
for exergy and thermoeconomic analysis of processes simulated with Aspen Plus®’, Energy,
Vol. 36, pp.964–974.
Radgen, P. (1996) ‘Pinch and exergy analysis of a fertilizer complex Part 1’, Nitrogen, No. 224,
pp.39–48.
Radgen, P. (1997) ‘Pinch and exergy analysis of a fertilizer complex Part 2’, Nitrogen, No. 225,
pp.27–39.
An overview of exergy analysis for chemical process industries 503

Rasheva, D.A. and Atanasova, L.G. (2002) ‘Exergy efficiency evaluation of the production of
sulfuric acid from liquid sulfur’, Exergy, An International Journal, Vol. 2, pp.51–54.
Rivero, R. (2001) ‘Exergy simulation and optimization of adiabatic and diabatic binary distillation’,
Energy, Vol. 26, pp.561–593.
Rivero, R. (2002) ‘Application of the exergy concept in the petroleum refining and petrochemical
industry’, Energy Conversion and Management, Vol. 43, pp.1199–1220.
Rivero, R. and Garcia, M. (2001) ‘Exergy analysis of a reactive distillation MTBE unit’,
International Journal of Applied Thermodynamics, Vol. 4, pp.85–92.
Rivero, R., Garcia, M. and Urquiza, J. (2004a) ‘Simulation, exergy analysis and application of
diabatic distillation to a tertiary amyl methyl ether production unit of a crude oil refinery’,
Energy, Vol. 29, pp.467–489.
Rivero, R., Rendón, C. and Gallegos, S. (2004b) ‘Exergy and exergoeconomic analysis of a crude
oil combined distillation unit’, Energy, Vol. 29, pp.1909–1927.
Rivero, R., Rendon, C. and Monroy, L. (1999) ‘The exergy of crude oil mixtures and petroleum
fractions: calculation and application’, International Journal of Applied Thermodynamics,
Vol. 2, pp.115–123.
Rizk, J., Nemer, M. and Clodic, D. (2012) ‘A real column design exergy optimization of a
cryogenic air separation unit’, Energy, Vol. 37, pp.417–429.
Rosen, M.A. (2001) ‘Editorial-exergy in industry: accepted or not?’, Exergy, An International
Journal, Vol. 1, p.67.
Rosen, M.A. (2002) ‘Does industry embrace exergy?’, Exergy, An International Journal, Vol. 2,
pp.221–223.
Rosen, M.A. and Dincer, I. (1999) ‘Exergy analysis of waste emissions’, International Journal of
Energy Research, Vol. 23, pp.1153–1163.
Rosen, M.A. and Scott, D.S. (1988) ‘Energy and exergy analyses of a production process for
methanol from natural gas’, International Journal of Hydrogen Energy, Vol. 13, pp.617–623.
Røsjorde, A. and Kjelstrup, S. (2005) ‘The second law optimal state of a diabatic binary tray
distillation column’, Chemical Engineering Science, Vol. 60, pp.1199–1210.
Røsjorde, A., Kjelstrupa, S., Johannessena, E. and Hansen, R. (2007) ‘Minimizing the
entropy production in a chemical process for dehydrogenation of propane’, Energy, Vol. 32,
pp.335–343.
Rücker, A. and Gruhn, G. (1999) ‘Exergetic criteria in process optimisation and process synthesis:
opportunities and limitations’, Computers & Chemical Engineering, Vol. 23, pp.S109–S112.
Ruyck, J.D. (1998) ‘Composite curve theory with inclusion of chemical reactions’, Energy
Conversion and Management, Vol. 39, pp.1729–1734.
Sama, D.A. (1995a) ‘Differences between second law analysis and pinch technology’, Journal of
Energy Resources Technology, Vol. 117, pp.186–191.
Sama, D.A. (1995b) ‘Second law insight analysis compared with pinch analysis as a design
method’, Presented at the International Conference on Second Law Analysis of Energy
Systems: Towards the 21st Century, 5–7 July, 1995, University of Rome, Italy.
Sama, D.A., Qian, S. and Gaggioli, R.A. (1989) ‘A common-sense second law approach for
improving process efficiencies’, TAIES’89: Proceedings of International Symposium on
Thermodynamic Analysis and Improvement of Energy Systems, Beijing, China, pp.520–531.
Sciubba, E. and Wall, G. (2007) ‘A brief commented history of exergy from the beginnings to
2004’, International Journal of Thermodynamics, Vol. 10, No. 1, pp.1–26.
Seider, W.D., Seader, J.D. and Lewin, D.R. (2009) Product and Processes Design Principles,
2nd ed., Wiley, New Delhi.
Simpson, A.P. and Lutz, A.E. (2007) ‘Exergy analysis of hydrogen production via steam methane
reforming’, International Journal of Hydrogen Energy, Vol. 32, pp.4811–4820.
Slade, B., Stober, B. and Simpson, D. (2006) ‘Dividing wall column revamp optimises mixed
xylenes production’, AICHE The 2006 Spring National Meeting, Orlando, USA.
504 S.C. Nimkar and R.K. Mewada

Smith Jr., R.L., Adschiri, T. and Arai, K. (2002) ‘Energy integration of methane’s partial-oxidation
in supercritical water and exergy analysis’, Applied Energy, Vol. 71, pp.205–214.
Song, C.M., Yan, Z.F. and Tu, Y.S. (1999) ‘Energy and exergy analysis of FCC unit’, Abstr. Pap.
Am. Chem. Soc., Vol. 217, p.U805.
Sorin, M. and Paris, J. (1997) ‘Combined exergy and pinch approach to process analysis’,
Computers & Chemical Engineering, Vol. 21, pp.S23–S28.
Sorin, M. and Paris, J. (1999) ‘Integrated exergy load distribution method and pinch analysis’,
Computers & Chemical Engineering, Vol. 23, pp.497–507.
Sorin, M., Bonhivers, J.C. and Paris, J. (1998b) ‘Exergy efficiency and conversion of chemical
reactions’, Energy Conversion and Management, Vol. 39, pp.1863–1868.
Sorin, M., Hammache, A. and Diallo, O. (2000a) ‘Exergy based approach for process synthesis’,
Energy, Vol. 25, pp.105–129.
Sorin, M., Hammache, A. and Diallo, O. (2000b) ‘Exergy load distribution approach for multi-step
process design’, Applied Thermal Engineering, Vol. 20, pp.1365–1380.
Sorin, M., Lambert, J. and Paris, J. (1998a) ‘Exergy flows analysis in chemical reactors’, Chemical
Engineering Research and Design, Vol. 76, pp.389–395.
Souza, J.A., Vargas, J.V.C., Ordonez, J.C., Martignoni, W.P. and von Meien, O.F. (2011)
‘Thermodynamic optimization of fluidized catalytic cracking (FCC) units’, International
Journal of Heat and Mass Transfer, Vol. 54, pp.1187–1197.
Stougie, L. and van der Kooi, H.J. (2009) ‘Exergy and sustainability’, ELCAS2009: Proceedings of
the 1st International Exergy, Life Cycle Assessment and Sustainability Workshop &
Symposium, Nisyros Island, Greece, pp.364–371.
Sumiju Plant Engineering Co., Ltd (2010) Mechanism of Sumitomo’s New Distillation System
‘Divided Wall Column’, http://www.spe.shi.co.jp/en/distill1.html (Accessed on 10 July, 2012).
Sun, J., Wang, F., Ma, T., Gao, H., Wu, P. and Liu, L. (2012) ‘Energy and exergy analysis of a
five-column methanol distillation scheme’, Energy, Vol. 45, pp.696–703.
Suphanit, B., Bischert, A. and Narataruksa, P. (2007) ‘Exergy loss analysis of heat transfer across
the wall of the dividing-wall distillation column’, Energy, Vol. 32, pp.2121–2134.
Sussman, M.V. (1980) ‘Steady-flow availability and the standard chemical availability’, Energy,
Vol. 5, pp.793–802.
Szargut, J., Morris, D. and Steward, F (1988) Exergy Analysis of Thermal, Chemical, and
Metallurgical Processes, Hemisphere, New York.
Szargut, J., Valero, A., Stanek, W. and Valero, A. (2005) ‘Towards an international reference
environment of chemical exergy’, ECOS 2005: Proceedings of the 18th International
Conference on Efficiency, Cost, Optimization, Simulation and Environmental Impact of
Energy Systems, Trondheim, Norway, pp.409–417.
Tahouni, N., Bagheri, N., Towfighi, J. and Panjeshahi, M.H. (2013) ‘Improving energy
efficiency of an Olefin plant – a new approach’, Energy Conversion and Management,
Vol. 76, pp.453–462.
Taprap, R. and Ishida, M. (1996) ‘Graphic exergy analysis of processes in distillation column by
energy-utilization diagrams’, AIChE Journal, Vol. 42, No. 6, pp.1633–1641.
Tirandazi, B., Mehrpooya, M., Vatani, A. and Moosavian, S. M. A. (2011) ‘Exergy analysis of C2+
recovery plants refrigeration cycles’, Chemical Engineering Research and Design, Vol. 89,
pp.676–689.
Tsatsaronis, G. (2007) ‘Definitions and nomenclature in exergy analysis and exergoeconomics’,
Energy, Vol. 32, pp.249–253.
Valero, A. (2008) Exergy Evolution of the Mineral Capital on Earth, PhD Thesis, University of
Zaragoza, Spain.
Voldsund, M., Ertesvåg, I.S., He, W. and Kjelstrup, S. (2013) ‘Exergy analysis of the oil and gas
processing on a North Sea oil platform a real production day’, Energy, Vol. 55, pp.716–727.
Wall, G. (2009) Exergetics, www.exergy.se/ftp/exergetics.pdf (Accessed on 5 September, 2011).
An overview of exergy analysis for chemical process industries 505

Wall, G. and Gong, M. (1996) ‘Exergy analysis versus pinch technology’, ECOS’96: Proceedings
of the Efficiency, Costs, Optimization, Simulation and Environmental Aspects of Energy
Systems, Stockholm, Sweden, pp.451–455.
Wang, Z., Zheng, D. and Jin, H. (2009) ‘Energy integration of acetylene and power polygeneration
by flowrate-exergy diagram’, Applied Energy, Vol. 86, pp.372–379.
Wei, Z., Wu, S., Zhang, B. and Chen, Q.(2012) ‘An exergy grand composite curve based procedure
for arranging side exchangers on distillation columns’, Computer Aided Chemical
Engineering, Vol. 31, pp.1592–1596.
Zemp, R.J., de Faria, S.H.B. and Oliveira Maia, M.d.L. (1997) ‘Driving force distribution and
exergy loss in the thermodynamic analysis of distillation columns’, Computers & Chemical
Engineering, Vol. 21, pp.S523–S528.
Zhifang, W. and Danxing, Z. (2008) ‘Exergy analysis and retrofitting of natural gas-based
acetylene process’, Chinese Journal of Chemical Engineering, Vol. 16, pp.812–818.
Zhu, P. (2008) ‘An improved calculation of the thermodynamically based allocation of cumulative
exergy consumption in the petroleum distillation process’, Chinese Journal of Chemical
Engineering, Vol. 16, pp.104–107.
Živković, P., Laković, M. and Rašković, P. (2004) ‘Exergy analyzing method in the process
integration’, Facta Universitatis Series: Mechanical Engineering, Vol. 2, pp.135–140.

Nomenclature
Esys Total exergy of system, kJ
PH
E sys Physical exergy of system, kJ
EKN Kinetic exergy, kJ
ECH Chemical exergy, kJ
PT
E Potential exergy, kJ
Etr Transiting exergy, kJ
U Internal energy
V Volume
T Temperature
S Entropy
P Pressure
Μ Chemical potential
M Mass
G Acceleration due to gravity
Z Height relative to environment
E Exergy
H Enthalpy
EQ Exergy of thermal energy
Q Heat duty
R Rate of reaction
I Irreversibility
ED Exergy destruction
W Work
506 S.C. Nimkar and R.K. Mewada

m Mass flow rate


∆t Change in time
CP Specific heat
yi Vapour mole fraction
xi Liquid mole fraction
ex
S Excess entropy
Wmin Minimum work
J Flow through interface, mol/s or J/s
X Force, J/mol K or 1/K
L Liquid flow rate
G Gas flow rate
chem
q Chemical heat release
E Specific exergy
N Molar flow rate
cr
E c Annual cumulative exergy rate of construction resource, kJ
op
E c Annual cumulative exergy rate of operating resource, kJ
EA Value of allocation annual total cumulative exergy rate, kJ
R Universal gas constant
Greek symbols
Σ Local entropy production rate, J/ s K m
Ω Cross-sectional area, m2
ρB Apparent catalyst density, kg/m3
ν Gas velocity, m/s
ε Exergy efficiency
Subscripts
0 Environmental state
00 Chemical potential at chemical equilibrium with environment
ele Electrical
C Cold fluid
H Hot fluid
F Feed
D Distillate
B Bottoms
RB Reboiler
CON Condenser
ref Reference
V Vapour
L Liquid
∆c Change in concentration
An overview of exergy analysis for chemical process industries 507

Q Heat
W Water
E Ethanol
N Tray number
surr Surrounding
ET Effluent treatment temperature
SAC Condensation temperature of sulphuric acid
Superscripts
L Liquid
V Vapour
HX Heat exchanger
Ig Ideal gas
F Feed
P Product
PP Pseudo product
0 Reference environment

You might also like