Download as pdf or txt
Download as pdf or txt
You are on page 1of 23

This article was downloaded by: 10.3.97.

143
On: 28 Oct 2023
Access details: subscription number
Publisher: CRC Press
Informa Ltd Registered in England and Wales Registered Number: 1072954 Registered office: 5 Howick Place, London SW1P 1WG, UK

CRC Handbook of Thermal Engineering Second Edition

Raj P. Chhabra

Pinch Technology

Publication details
https://www.routledgehandbooks.com/doi/10.4324/9781315119717-47
Santanu Bandyopadhyay, Shankar Narasimhan
Published online on: 05 Dec 2017

How to cite :- Santanu Bandyopadhyay, Shankar Narasimhan. 05 Dec 2017, Pinch Technology
from: CRC Handbook of Thermal Engineering Second Edition CRC Press
Accessed on: 28 Oct 2023
https://www.routledgehandbooks.com/doi/10.4324/9781315119717-47

PLEASE SCROLL DOWN FOR DOCUMENT

Full terms and conditions of use: https://www.routledgehandbooks.com/legal-notices/terms

This Document PDF may be used for research, teaching and private study purposes. Any substantial or systematic reproductions,
re-distribution, re-selling, loan or sub-licensing, systematic supply or distribution in any form to anyone is expressly forbidden.

The publisher does not give any warranty express or implied or make any representation that the contents will be complete or
accurate or up to date. The publisher shall not be liable for an loss, actions, claims, proceedings, demand or costs or damages
whatsoever or howsoever caused arising directly or indirectly in connection with or arising out of the use of this material.
4.14 PINCH TECHNOLOGY
Santanu Bandyopadhyay and Shankar Narasimhan
Downloaded By: 10.3.97.143 At: 15:12 28 Oct 2023; For: 9781315119717, sec4_14, 10.4324/9781315119717-47

Introduction
The human appetite for energy has been growing rapidly, especially since the industrial revolution,
and it is now used as one of the indicators of a country’s progress and development. According to a
UNIDO working paper in 2010 (Upadhyaya, 2010), more than 80% of the world’s energy consump-
tion is met by fossil fuels. Out of this, manufacturing accounts for a quarter of the total energy con-
sumption. The increase in energy consumption has concomitantly led to a growing concern about
the impact that the burning of fossil fuels has on the environment. While the world continues to seek
ways and means of reducing our dependence on fossils fuels, by utilizing renewable energy sources,
it is also important to use advanced technological methods to improve the energy efficiency of
industrial processes. Pinch technology is a systematic approach that was developed in the mid-1980s
to analyze the energy utilization in process industries, and to design integrated energy sub-system
of a process which achieves the optimal trade-off between the cost of energy used and the capital
investment made (Linnhoff et al., 1982; Klemeš, 2013).
In any design, there exists a trade-off between the capital investment and operating cost. For
an energy sub-system, the capital cost is mainly dependent on the number and size of heat transfer
equipment (boilers, furnaces, heat exchangers), while the operating cost is primarily a function of
the consumption rate of utilities such as fuel, cooling water, refrigerants, steam, etc. Pinch technol-
ogy effectively captures the trade-off between capital and operating costs, and can be used for the
optimal design of the energy sub-system of a process. The method is now widely used to design the
heat and power sub-system of processes such as refineries, petrochemical plants, pulp and paper
plants, thermal power plants, etc. It is reported that energy savings of up to 30% have been achieved
using this technology. Although, pinch technology started by considering heat integration in pro-
cesses it has now been extended to minimize usage of other resources such as water and to mini-
mize emissions, and has also been rechristened as pinch analysis (Linnhoff, 1993). In this chapter,
the basic concepts used in pinch technology are introduced and explained. The use of these concepts
for the optimum design of a heat exchanger network is also described in detail. The integration of
the heat exchanger network with other energy consumers/producers such as distillation columns,
pumps, compressors, and turbines is also discussed. There are several books, book chapters, and
review articles that the reader can refer to get a more comprehensive understanding of pinch tech-
nology. Shenoy (1995) and Kemp (2007) have written books devoted entirely to pinch technology,
while the book by Smith (2005) contains several chapters related to pinch technology. A bibliog-
raphy compiled by Furman and Sahinidis (2002) contains references to more than 400 research
papers on HEN synthesis based on pinch technology and mathematical programming methods.
Early applications of pinch technology were mainly in refineries and petrochemical plants (Kemp,
2007), but applications to other industries such as pulp and paper (Svensson and Harvey, 2011) and
steel (Grip et al., 2013) have been recently reported.

Motivating Example
Before describing the essential concepts of pinch technology and how it can be used to design a heat
exchanger network (HEN), a simple example is provided to illustrate the significant role that a HEN
structure has on the energy requirements. A simplified example abstracted from a thermal power
plant is used to provide the motivation. In a thermal power plant, fuel is burnt in a boiler to generate
high-pressure steam, which is subsequently used to generate power. The flue gas from the boiler
contains sufficient energy, which can be recovered. The recovered energy can be used to preheat the
air used for combustion, or to preheat the water fed to the boiler for generating steam. Figure 4.14.1

875
876 CRC Handbook of Thermal Engineering

25°C Air 104°C Water

390°C 230°C 142°C


1225 m2 1218 m2
Flue gas
Downloaded By: 10.3.97.143 At: 15:12 28 Oct 2023; For: 9781315119717, sec4_14, 10.4324/9781315119717-47

160°C 127°C

FIGURE 4.14.1 Existing HEN structure.

TABLE 4.14.1
Streams and Their Properties
Stream Specific Heat Capacity (kJ/kg °C) Flow Rate (kg/h)
Flue gas 1.1 39,032
Air 1.01 40,072.8
Water (10.5 bar) 4.41 31,824.0

TABLE 4.14.2
Exchangers and Their Properties for Original Configuration
Heat Lost by Hot Heat Gained by Cold Heat Lost to
Exchanger Area (m2) Stream (kW) Stream (kW) Ambient (kW) U (W/m2 K)
HE1 1225.0 1908.23 1517.76 390.47 5.7
HE2 1218.1 1049.53 897.25 152.28 11.3

shows an existing network of two heat exchangers in series used to recover energy from the flue gas.
The flue gas is first used to preheat the air in the first exchanger (HE1) and subsequently to preheat
pressurized water in the second exchanger (HE2). The flow rate and specific heat capacity of the
different streams are given in Table 4.14.1, while the inlet and outlet temperatures of streams from
each exchanger are shown in Figure 4.14.1. The heat exchanger areas, heat loads, and heat lost to
ambient are given in Table 4.14.2.
From the heat loads of exchangers, it is observed that 2957.76 kW of energy is recovered from the
flue gas out of which 542.75 kW is lost (about 20%) to the ambient. From Figure 4.14.1, it is observed
that the final exit temperature of the flue gas is 142°C. The question is whether it is possible to
extract more energy from the flue gas using the same exchangers. Figure 4.14.2 shows an alternative
configuration where the flue gas is first used to preheat water in exchanger HE1 followed by preheat-
ing air in exchanger HE2. It is assumed that the overall heat transfer coefficient depends only on
the streams exchanging energy. Furthermore, in order to make a fair comparison, it is assumed that
heat losses in the exchangers are same as before. Table 4.14.3 shows the heat loads of exchangers
for the new configuration. It is observed that the total energy recovered from flue gas is 3375.5 kW
out of which the loss to the ambient is almost same as before (542.4 kW). The final exit tempera-
ture of the flue gas is 107°C. If this temperature is above the acid dew point, then technically it is
possible to extract 10% more energy from the flue gas just by a simple change in the structure of
the heat exchanger network (interchange the order in which flue gas exchanges energy with air and
water). Although it is not shown here, it is also possible to recover 10% more energy using the same
exchangers in a parallel configuration with appropriate apportioning of the flue gas flow rate to
the exchangers. This example clearly demonstrates that the heat exchanger network configuration
Applications 877

104°C Water 25°C Air

390°C 173.8°C 107°C


1225 m2 1218 m2
Flue gas
Downloaded By: 10.3.97.143 At: 15:12 28 Oct 2023; For: 9781315119717, sec4_14, 10.4324/9781315119717-47

160°C 82.5°C

FIGURE 4.14.2 Modified HEN structure.

TABLE 4.14.3
Exchangers and Their Properties for Modified Configuration
Heat Lost by Hot Heat Gained by Cold Heat Lost to
Exchanger Area (m2) Stream (kW) Stream (kW) Ambient (kW) U (W/m2 K)
HE1 1225.0 2578.2 2186.3 391.9 11.3
HE2 1218.1 797.3 645.8 151.5 5.7

plays a significant role in the overall energy consumption of a process. In an actual thermal power
plant, several heat exchangers (preheaters, superheaters, economizers, feed heaters, etc.) are used
in a series-parallel arrangement. A more complex network of exchangers is used in a refinery to
preheat the crude oil using the hot distillate streams of the atmospheric and vacuum columns. The
optimal design of the heat exchanger network that can recover the maximum energy by interchang-
ing energy between the process streams, while giving due consideration to the capital investment is
a challenging problem. Pinch technology provides an insightful and systematic approach for solving
this design problem. Although the description in the rest of the chapter pertains to the design of a
HEN for a grassroots design, the concepts of pinch technology can also be used to analyze existing
processes and retrofit these designs to recover more energy.

Pinch Technology
One of the important features of pinch technology from a design viewpoint is that it can be used to
estimate achievable targets prior to the actual design of a HEN. These estimated targets include the
following:

• Minimum external hot and cold utility required for solving a given problem without violat-
ing thermodynamic constraints
• Minimum number of heat exchangers required for heat integration
• Minimum overall area of the HEN
• Minimum overall capital cost of the HEN

While the minimum external utility requirements estimate can be shown to be theoretically rigor-
ous, all other target estimates are approximate, but are reasonably good for design purposes. The
establishment of such targets prior to design is useful for assessing how closely a proposed design
meets these targets. It also indicates the direction in which a proposed design should be modified or
evolved for meeting the targets.
Another feature of pinch technology is the extensive use of graphical diagrams such as com-
posite curves, temperature interval analysis and cascade diagram, source–sink diagram, driv-
ing force plots, etc., which provide significant insights into the energy integration possibilities.
878 CRC Handbook of Thermal Engineering

Tc,out Tc,in
Mc
Mh
Th,in Th,out

FIGURE 4.14.3 Counter-current heat exchanger.


Downloaded By: 10.3.97.143 At: 15:12 28 Oct 2023; For: 9781315119717, sec4_14, 10.4324/9781315119717-47

The computation of the different targets for a given example problem, and the graphical diagrams
used in these computations are first explained before the actual design of the HEN.
Interestingly, the mathematical background required for understanding pinch technology is min-
imal. Essentially, the following three equations that relate the different quantities associated with a
heat exchanger are used in the design process. It is assumed that all the heat exchangers are counter-
current exchangers, the schematic of one of which is shown in Figure 4.14.3.

Q = M h Cph (Th,in − Th,out ) = M c Cpc (Tc,out − Tc,in ) (4.14.1)

Q = UA∆Tln (4.14.2)


where

 ∆Tln = LMTD =
(Th,in − Tc,out ) − (Th,out − Tc,in )
 (Th,in − Tc,out ) 
ln  
 (Th,out − Tc,in ) 

1 1 1
= + (4.14.3)
U hh hc 

The first equation relates the heat load of the exchanger (Q) to the heat lost by the hot stream and the
heat gained by the cold stream, where for simplicity it is assumed that the specific heat capacities of
the streams are constant and do not change with the temperature of the fluid. The second equation
is the design equation, which relates the heat load to the overall heat transfer coefficient and area
of exchanger, and the third equation relates the overall heat transfer equation to the heat transfer
coefficients of the hot and cold streams that are exchanging heat. In general, the stream heat transfer
coefficients depend on the geometry of the heat exchanger (such as diameter of the tubes, number of
tubes, tube pitch and layout, shell diameter, baffle spacing, etc.). However, for the purpose of design-
ing the HEN, it is assumed that the stream heat transfer coefficients for a given stream are constant
and are computed for assumed typical values of the exchanger geometry.

Energy Targeting
Graphical Analysis
Pinch analysis was developed as a systematic physical-insight-based tool for energy conservation in
heat exchanger networks (HENs) (Hohmann, 1971; Linnhoff and Flower, 1978; Umeda et al., 1978).
The energy targets are obtained via developing temperature-heat duty profiles of heat availability
and heat demands of all the process streams. To illustrate the graphical procedure, an illustrative
example of two hot streams and two cold streams is considered. Supply and target temperatures of
these streams, along with their heat capacity flow rates, are given in Table 4.14.4.
The heat availability profile of all the hot streams (i.e., streams that are to be cooled) is called the
hot composite curve. Hot composite curve for this example can be generated by thermodynamically
adding two hot streams, H1 and H2. Figure 4.14.4a shows the two hot streams on temperature-heat
Applications 879

TABLE 4.14.4
Stream Data for Illustrative Example
Stream Tin (°C) Tout (°C) MCp (MW/°C) Duty (MW)
H1 210 60 0.10 15
Downloaded By: 10.3.97.143 At: 15:12 28 Oct 2023; For: 9781315119717, sec4_14, 10.4324/9781315119717-47

H2 180 30 0.20 30
C3 40 100 0.15 9
C4 60 200 0.30 42

(a)
240
210
180
Temperature (°C)

H1
150
H2
120
90
60
30
15 MW 30 MW
0
0 10 20 30 40 50
Heat duty (MW)
(b)
240

210

180
Temperature (°C)

e
150 urv
te c
osi
120 comp
t
Ho
90

60

30

0
0 10 20 30 40 50
Heat duty (MW)

FIGURE 4.14.4 (a) Temperature-heat duty diagrams of hot streams, and (b) Hot composite curve.

duty diagram. The positioning of the streams in the temperature-heat duty diagram is arbitrary on
the heat duty scale. However, the horizontal distance for each stream is fixed. This is because the
starting value is not important; the overall heat duty of each stream is important. Figure 4.14.4b
shows the two hot streams combined together to give a hot composite curve. The addition of heat
availability from both the streams can be performed between each temperature interval. This is
achieved by combining the duties and heat capacity flow rates of all the streams that have a common
temperature range. This can also be done algebraically. Based on the supply and target temperatures
of these streams, three temperature intervals can be identified: 30°–60°, 60°–180°, and 180°–210°.
Between 30°C and 60°C, heat is available from H2 only. In this temperature interval, the total heat
880 CRC Handbook of Thermal Engineering

available is 0.2 × (60 − 30) = 6 MW. Between 60°C and 180°C, heat is available from both H1 and
H2. In this temperature interval, total heat available is (0.2 + 0.1) × (180 − 60) = 36 MW. Similarly,
3 MW of heat, from the hot stream H1, is available in the temperature interval 180°–210°. For each
temperature interval, heat available can be represented as a line segment on a temperature-heat duty
diagram. Therefore, the coordinates of the hot composite curve in Figure 4.14.4b are (0, 30), (6, 60),
Downloaded By: 10.3.97.143 At: 15:12 28 Oct 2023; For: 9781315119717, sec4_14, 10.4324/9781315119717-47

(42, 180), and (45, 210).


Similarly, heat requirement profile of all the cold streams (i.e., streams that are to be heated) is
called the cold composite curve. Based on the supply and target temperatures of two cold streams, C3
and C4, three temperature intervals can be identified: 40°–60°, 60°–100°, and 100°–200°. Between
40°C and 60°C, heat is required by C3 only. In this temperature interval, total heat required is 0.15
× (60 − 40) = 3 MW. Similarly, 18 and 30 MW of heat are required in the temperature intervals 60°–
100° and 100°–200°. By joining these heat requirement line segments, the cold composite curve can
be generated. Individual cold streams and the cold composite curve are shown in Figure 4.14.5. The
coordinates of the cold composite curve are (0, 40), (3, 60), (21, 100), and (51, 200).
By construction, these composite curves, as a function of temperature, are monotonic in nature.
These composite curves are plotted on a single diagram such a way that they are vertically separated

(a)
220
200
180
160
Temperature (°C)

140 C4
120
100 C3
80
60
40
9 MW 42 MW
20
0
0 10 20 30 40 50 60
Heat duty (MW)
(b)
210

180
ve
c ur
150 s ite
o
mp
Temperature (°C)

d co
120 l
Co

90

60

30

0
0 10 20 30 40 50 60
Heat duty (MW)

FIGURE 4.14.5 (a) Temperature-heat duty diagrams of cold streams, and (b) Cold composite curve.
Applications 881

Hot utility
240
12 MW
210

180
rve
cu
Temperature (°C)
Downloaded By: 10.3.97.143 At: 15:12 28 Oct 2023; For: 9781315119717, sec4_14, 10.4324/9781315119717-47

e
150 sit
e
c urv
po te
t com p osi
120 Ho d com
Pinch point l
Co
90

60

30
Cold utility 6 MW
0
0 10 20 30 40 50 60
Heat duty (MW)

FIGURE 4.14.6 Hot and cold composite curves at ΔTmin of 10°C and energy targeting.

by at least a given minimum approach temperature (ΔTmin). This vertical temperature difference
ensures satisfaction of the second law of thermodynamics and proper heat transfer from the hot
streams to the cold streams (see Figure 4.14.6). For this illustrative example, ΔTmin is assumed to
be 10°C. The overlap of these two composite curves indicates the maximum process heat recovery
between hot and cold streams. Overhang of the hot composite curve over the cold composite curve
denotes the requirement of the minimum cold utility (to remove excess heat from the system) and
similarly, overshoot of the cold composite curve over the hot composite curve indicates the require-
ment of the minimum hot utility (to supply additional heat to the system). From Figure 4.14.6, the
minimum hot and cold requirements are identified to be 12 and 6 MW. The point where ΔTmin is
observed is known as the pinch point (see Figure 4.14.6). This is the point where the temperature
driving force is the least or the composite curves are closest.

Algebraic Analysis
Problem Table Algorithm (PTA), originally proposed by Linnhoff and Flower (1978), is the alge-
braic tool of pinch analysis for targeting the minimum external utility requirements in a heat
recovery network. Over the years, different modifications of the original PTA were proposed. For
example, Ozkan and Dincer (2001) combined shifting of stream temperatures and the PTA in one
tabular calculation and applied to target utility requirements in a refinery. Salama (2005) introduced
numerical calculation procedure based on the reversal procedure of PTA. Based on the geometrical
analysis, an algebraic technique was formulated by Salama (2006). An algebraic procedure to target
the minimum utilities, known as the modified Problem Table Algorithm (MPTA), was proposed by
Bandyopadhyay and Sahu (2010). In this chapter, MPTA is applied for energy targeting. Steps of the
proposed algorithm are as follows.

Step 1: Shifting of stream temperatures.


Hot stream temperatures are shifted down by ½ ΔTmin and the cold stream tempera-
tures are shifted up by ½ ΔTmin. However, it may be noted that any shifting may be
permitted as long as the net shifting between the hot and cold stream temperatures are
exactly ΔTmin. This ensures that there is an adequate driving force (ΔTmin) between the
hot and cold streams.
Step 2: Determination of sources and demands from shifted stream data.
Inlet of each stream, irrespective of whether it is hot or cold, is considered as a
source. Each source produces the heat capacity flow rate (MCP) at its inlet temperature.
882 CRC Handbook of Thermal Engineering

Similarly, the outlet of each stream is considered as a demand that accepts a heat capac-
ity flow rate (MCP) at respective outlet temperature. Source and demand data for the
illustrative example are given in Table 4.14.5.
Step 3: Determination of temperature intervals (column 1)
Temperatures of all sources and demands, as determined in the previous step, are
Downloaded By: 10.3.97.143 At: 15:12 28 Oct 2023; For: 9781315119717, sec4_14, 10.4324/9781315119717-47

tabulated in decreasing order in the first column. If the value of a particular temperature
occurs more than once, the same need not be repeated. The temperature for kth row is
denoted as Tk . In the first column of Table 4.14.6, temperatures are arranged in descend-
ing order for the example.
Step 4: Calculation of heat capacity flows (column 2)
Heat capacity flows (i.e., algebraic sum of heat capacity flows corresponding to any
particular temperature) are tabulated in the second column (see Table 4.14.6). Consider
flows corresponding to sources as positive and demands as negative. For kth row, total
heat capacity flow is denoted as Fk.
Step 5: Calculation of net MCP (column 3)
Cumulative heat capacity flows are tabulated in the third column. Summation of heat
capacity flows of all previous rows denotes the cumulative heat capacity flows for any
particular row. Due to conservation of mass flow rate and constant specific heat for each

TABLE 4.14.5
Source and Demand Data for Illustrative Example with Shifted Temperatures
Heat Capacity Flow (MW/°C) Shifted Temperature (°C)
Sources
SH1 0.10 205
SH2 0.20 175
SC3 0.15 45
SC4 0.30 65
Demands
DH1 0.10 55
DH2 0.20 25
DC3 0.15 105
DC4 0.30 205

TABLE 4.14.6
Modified Problem Table Algorithm Applied to Illustrative Example
Tint (°C) Fk (kW/°C) MCp,int (kW/°C) Qint (kW) Qcas (kW) Rcas (kW)
205 −0.2 −0.2 0 0 12
175 0.2 0 −6 −6 6
105 −0.15 −0.15 0 −6 6
65 0.3 0.15 −6 −12 0
55 −0.1 0.05 1.5 −10.5 1.5
45 0.15 0.2 0.5 −10 2
25 −0.2 0 4 −6 6
Applications 883

stream, last entry should be zero. Mathematically cumulative heat capacity flow at Tk is
given as:
k

CumFk = ∑F l (4.14.4)
Downloaded By: 10.3.97.143 At: 15:12 28 Oct 2023; For: 9781315119717, sec4_14, 10.4324/9781315119717-47

l =1

Step 6: Calculation of net enthalpy (column 4)


The fourth column represents the enthalpy (Qk) for each temperature interval. The
first entry in the fourth column is kept 0. For all subsequent columns, the difference
between the last two temperatures is multiplied by the cumulative heat capacity flows to
calculate the enthalpy and tabulated in the fourth column. Mathematically, net enthalpy
(Qk) for each temperature interval can be calculated using the following formula:

0, for k = 1


Qk =   k −1  (4.14.5)
(Tk−1 − Tk ) 
 
 l=1 

Fl  ,

for k > 1

Net enthalpy, calculated in this column, represents net surplus (Qi > 0) or net deficit
(Qi < 0) of enthalpy in each temperature interval.
Step 7: Calculation of cascaded heat duty (column 5)
Cascaded heat duty for each temperature interval can be calculated by summing
enthalpies for all previous rows and tabulated in the fifth column. Using Equation 4.14.5,
cascaded heat duty for kth row may be expressed as:

0, for k = 1
k
  
Qcas,k = ∑ Ql  =  k −1 (4.14.6)
 l =1   ∑
 l=1
Fl (Tl − Tk ), for k > 1

Step 8: Revised cascade (column 6)


As some of the entries in column 5 may be negative, it suggests an infeasible cascade.
The minimum value of Qcas in column 5 is subtracted from each value in the fifth column
to obtain the revised feasible cascaded heat duty and tabulated in the sixth column.

Rcas,k = Qcas,k − min (Qcas ) (4.14.7)

In Table 4.14.6, the minimum value of Qcas in column 5 is 12 MW. Using Equation 4.14.7, revised
cascade data are generated and tabulated in column 6.
The first and last values of the sixth column suggest the minimum hot (Qhu,min) and the mini-
mum cold (Qcu,min) utility requirements, respectively. Based on the data reported in column 6 of
Table 4.14.6, the minimum hot and cold requirements are identified to be 12 and 6 MW, identical to
the values obtained graphically (see Figure 4.14.6).
The temperature in column 1 that corresponds to zero revised cascaded heat in column 6 is the
pinch temperature. For the illustrative example, pinch corresponds to the shifted temperature of
65°C. Therefore, the hot and cold side pinch temperatures are 70° and 60°C.
884 CRC Handbook of Thermal Engineering

Golden Rules of Pinch Analysis


The pinch temperature boundary divides the process into two sub-processes, above and below the
pinch. The process hot stream segments above pinch temperature should not transfer their energy
to cold stream segments below pinch temperature, if a design has to meet minimum utility targets.
This is nicely illustrated by the source–sink diagram shown in Figure 4.14.7a. It may be noted that
Downloaded By: 10.3.97.143 At: 15:12 28 Oct 2023; For: 9781315119717, sec4_14, 10.4324/9781315119717-47

the sub-process above the pinch is a net sink since it requires external hot utility to meet its energy
requirements, while the sub-process below the pinch is a net source since it requires external cold
utility to transfer the excess energy available. If an amount of energy α is transferred across pinch,
then both the hot utility and cold utility requirements increase by the same amount α to satisfy the
energy balances above and below pinch. Similarly, if cold utility is used above pinch or hot utility
is used below pinch, then both hot and cold utility requirements increase by an identical amount.
Figure 4.14.7b–d illustrates this graphically. Therefore, in order to achieve the minimum energy
targets for a process, following rules must be obeyed. If these rules are not satisfied, requirement
of hot and cold utilities goes up simultaneously, leading to a double energy penalty. These rules are
known as golden rules of pinch analysis.

• No heat transfer across the pinch,


• No cold utility above the pinch, and
• No hot utility below the pinch.

Cost Targeting
Once the energy targets are established, the operating cost of the HEN can easily be established.
In addition to the operating cost, it is useful to estimate the capital cost of the network. The capital
cost of a HEN depends on the number of heat exchangers, number of shells, type of heat exchangers,
heat exchange area, material of construction, pressure rating, etc. Capital cost targeting for HEN is
briefly discussed.

(a) (b)
HUmin HUmin + α

Net sink Net sink

α
Pinch boundary Pinch boundary

Net source Net source

CUmin CUmin + α

(c) (d)
HUmin + α HUmin

Net sink Net sink


α CU
Pinch boundary Pinch boundary
HU
Net source α Net source

CUmin CUmin + α

FIGURE 4.14.7 Derivation of golden rules of pinch analysis. (a) No heat transfer across pinch. (b) Heat
transfer of α across pinch. (c) Cold utility of α used above pinch. (d) Hot utility of α used below pinch.
Applications 885

Targeting for Number of Heat Exchangers


A minimum number of heat exchangers can be estimated using the mathematical results from graph
theory (Linnhoff et al., 1979). The number of heat exchanger in a HEN is related to the number of
streams by the following equation:
Downloaded By: 10.3.97.143 At: 15:12 28 Oct 2023; For: 9781315119717, sec4_14, 10.4324/9781315119717-47

 Number of heat   Number of streams  Number of independent  Number of independent 


  =  +  − 
 exchangers   including utilities   loops in the network   components in the network 
 (4.14.8)

Before designing a HEN, a number of independent loops and independent components are not
known. For simplicity, it is assumed that there is no loop in the network and there is a single com-
ponent. These assumptions can simplify Equation 4.14.8 significantly.

 Number of heat   Number of streams 


  =  −1 (4.14.9)
 exchangers   including utilities 

Equation 4.14.9 can be used as a targeting tool to estimate the number of heat exchangers, prior to
the design of the HEN. Typically, Equation 4.14.9 is applied for the above pinch and below pinch
portion separately to estimate the total number of heat exchangers. For the illustrative example, all
four process streams exist above pinch and only hot utility can be used. There are total five streams
and hence, four exchangers are expected. Similarly, there are three process streams (H1, H2, and
C3), which exist below pinch with cold utility. This leads to a target of three heat exchangers below
pinch. The total number of heat exchangers for this example is targeted to be 7.

Targeting for Heat Transfer Area


Heat transfer area may be estimated from the composite curves, after incorporating utility streams.
Composite curves may be divided into vertical enthalpy intervals as shown in Figure 4.14.8.
Assume a constant overall heat transfer coefficient U for the entire network and countercurrent
heat transfer, the heat transfer area for each enthalpy interval may be calculated using the follow-
ing formula:

 Heat transfer area in  heat duty of the enthalpy interval


 = (4.14.10)
 an enthalpy interval  U × LMTD

where LMTD stands for the log mean temperature difference of the interval. Different heat transfer
coefficient for each stream can easily be incorporated and the following expression may be used for
targeting the total heat transfer area for the HEN (Linnhoff and Ahmad, 1990):

Intervals
1  hot streams q cold streams
q j,k 
A= ∑ 
LMTDk  i ∑ i,k
hi
+ ∑ h

j 
(4.14.11)
k  j 

It should be noted that Equation 4.14.11 does not predict the true minimum network area for dif-
ferent heat transfer coefficient for each stream. Another difficulty is to estimate heat transfer coef-
ficient for each stream without a detailed design of the heat exchanger.
886 CRC Handbook of Thermal Engineering

240

210

180

Temperature (°C)
150
Downloaded By: 10.3.97.143 At: 15:12 28 Oct 2023; For: 9781315119717, sec4_14, 10.4324/9781315119717-47

120

90

60

30

0
0 10 20 30 40 50 60
Heat duty (MW)

FIGURE 4.14.8 Enthalpy intervals and area targeting.

Cost Optimization
Typically, the cost of a single heat exchanger with surface area A can be estimated using some sim-
plifying relation of the following form:

(Cost of a Heat Exchanger ) = a + bAc (4.14.12)

where a, b, and c are various cost constants that vary according to the materials of construction,
pressure rating, and type of exchanger. Using this, a simplified expression may be derived to esti-
mate the capital cost of the entire network (Ahmad et al., 1990).

(Capital Cost of HEN ) = N (a + b( A / N )c ) (4.14.13)

where N is the number of heat exchangers.


Once the capital cost is targeted, total annualized cost of the entire network may also be
estimated by adding annual operating cost with the annualized capital cost. By varying the total
annualized cost of HEN with respect to the ΔTmin, the optimum value of ΔTmin may be deter-
mined (Linnhoff and Ahmad, 1989). A typical capital-energy trade-off, as a function of ΔTmin,
is illustrated in Figure 4.14.9. As ΔTmin increases, the consumption of both hot and cold utilities
increases together in order to maintain the energy balances. This leads to increase in operating
cost. Typically, energy cost or the operating cost is a piece-wise linear function of ΔTmin. On
the other hand, the temperature differences for heat transfer between the hot and cold streams
throughout the process become larger as ΔTmin increases. This means that the heat transfer area
decreases, hence decreasing the capital costs. The jumps in the capital cost curve are created due
to variation in the expected number of heat exchangers. Combining the energy cost and capital
cost allows an optimal value of ΔTmin to be identified. In practice, the shape of the optimization
curve tends to be quite flat and a very precise optimization is often not necessary. As long as the
value of ΔTmin is somewhere in the region of the optimum point, there is little to be lost or gained
in terms of the overall cost by small changes in ΔTmin. For chemical processes, a reasonable value
of ΔTmin is usually around 10°C–15°C. A larger value is usually required for refinery processes
(typically 15°–30°C). For low-temperature processes, a reasonable value of ΔTmin is often lower
(typically 5°C or less), because of the expense of providing cooling through refrigeration in low-
temperature processes.
Applications 887

Total
annualized
cost

Annualized cost
Downloaded By: 10.3.97.143 At: 15:12 28 Oct 2023; For: 9781315119717, sec4_14, 10.4324/9781315119717-47

Capital
cost

Operating
cost
Optimum

ΔTmin

FIGURE 4.14.9 Capital-energy trade-offs and determination of optimum ΔTmin.

240
Hot utility 12 mw
210

180
Shifted temperature (°C)

150

120

90

60 Pinch point

30
Cold utility 6 mw
0
0 5 10 15
Heat duty (MW)

FIGURE 4.14.10 Grand composite curve (GCC).

Targeting for Multiple Utilities


In the previous section, energy targeting is performed with a single hot utility and a single cold util-
ity. However, in many practical situations, multiple utilities are present. To target multiple utilities a
piece-wise linear curve, known as the Grand Composite Curve (GCC), may be utilized. GCC may
be generated by drawing the shifted temperature (in column 1 of MPTA) against the feasible cas-
cade (in column 6 of MPTA) in a temperature-heat duty diagram. GCC for the example is shown in
Figure 4.14.10. The GCC represents the heat demand (segments with positive slope) and supply (seg-
ments with negative slope) within each temperature interval. GCC touches the temperature axis at
the shifted pinch temperature, signifying no heat transfer across the pinch. The portion of the GCC
above the pinch point represents heat demand and requires only hot utilities. Similarly, the portion
below the pinch has surplus heat and this surplus heat is rejected through cold utilities.
Figure 4.14.11 illustrates some examples in the use of the grand composite curve. Figure 4.14.11a
shows high-pressure steam and cooling water used to satisfy the heating and cooling requirements
of a GCC. In Figure 4.14.11b, high-pressure and low-pressure steams are used to satisfy the hot util-
ity requirement of a GCC. Flue gas from a fired heater or from gas turbine exhaust may also be used
888 CRC Handbook of Thermal Engineering

(a) (b) (c)


High pressure steam

as
High pressure steam

rg
st r o
Shifted temperature

au te
Shifted temperature

xh hea
Shifted temperature

ee d
in r e
Low pressure

rb f i
tu o m
steam

r
Downloaded By: 10.3.97.143 At: 15:12 28 Oct 2023; For: 9781315119717, sec4_14, 10.4324/9781315119717-47

sf
ga
ue
Fl
Cooling water Cooling water
Heat duty Heat duty
Cooling water

Heat duty

FIGURE 4.14.11 Examples in the use of the GCC: (a) use of high-pressure steam and cooling water, (b) use
of multiple steam levels, and (c) use of flue gas as hot utility.

to provide hot utility requirement in a process (see Figure 4.14.11c). Essentially utility profiles are
matched against the GCC to target multiple utilities. During profile matching, utility profiles may
touch the GCC. This does not imply that ΔTmin is zero. It should be noted that the temperature axis
in GCC is in shifted scale and the same temperature shift (by ΔTmin/2) should also be incorporated
in generating the utility profile. Therefore whenever the utility profile touches the GCC, the tem-
perature driving force is exactly equal to ΔTmin.
Shenoy et al. (1998) proposed a targeting methodology to determine the optimum distributions of
loads for multiple utilities considering the cost tradeoffs. GCC plays an important role in the ther-
mal integration of fired heaters (Varghese and Bandyopadhyay, 2007), integration between various
processes (Bandyopadhyay et al., 2010), cogeneration plant (Bade and Bandyopadhyay, 2015), etc.

Network Design and Evolution


In the previous sections, methodologies for setting various targets are discussed. It should be noted
that these targets are set prior to the detailed design of the HEN. In this section, a simple methodol-
ogy, called the pinch design method, to design HEN is discussed (Linnhoff and Hindmarsh, 1983).
The starting point is the energy targets and the pinch temperatures. The design procedure starts by
setting up a design grid, as shown in Figure 4.14.12. The hot streams are shown at the top of the dia-
gram running in the direction left to right and the cold streams at the bottom of the diagram running
right to left. A vertical line at the pinch temperature of 70°C for the hot streams and 60°C for the
cold streams divides the grid into two parts (see Figure 4.14.12). To the left of the pinch line is the
above pinch portion. To the right of the pinch line is the below pinch portion. It should be noted that

Pinch line
Above pinch portion 70°
210° 60°
H1
180° 30°
H2

100° 40°
C3
200°
C4 Below pinch portion

60°

FIGURE 4.14.12 Parallel grid diagram for heat exchanger network design.
Applications 889

all four streams exist above pinch, but only three streams exist in the below pinch portion. Pinch
is the most constrained part of the composite curve. Temperature driving force is the minimum at
the pinch (see Figure 4.14.6). The design procedure starts at the pinch region and proceeds further.
For the successful design of a HEN, two feasibility criteria have to be met. First feasibility cri-
terion is related to the heat capacity flow rates (MCp) of the hot and the cold streams. When a hot
Downloaded By: 10.3.97.143 At: 15:12 28 Oct 2023; For: 9781315119717, sec4_14, 10.4324/9781315119717-47

stream and a cold stream, just above the pinch point, are selected to transfer heat through a heat
exchanger, the temperature driving force at the pinch side is exactly ΔTmin. The temperature driv-
ing force at the other side of the exchanger must be above ΔTmin. This is only possible if the MCp of
hot stream is lower than that of the cold stream. The exactly opposite relation has to be obeyed just
below the pinch. The MCp feasibility criterion may be summarized as:

(MCP )hot stream ≤ (MCP )cold stream just above pinch (4.14.14)

(MCP )hot stream ≥ (MCP )cold stream just below pinch (4.14.15)

Second feasibility criterion is related to the number of streams present at the pinch. To achieve
the energy target, no cold utility to be placed above the pinch. This implies that every hot stream,
present just above the pinch, must be brought to the pinch temperature by exchanging heat with
a cold stream. Therefore, the number of cold streams, just above the pinch, must be greater than
that of the hot streams. The similar argument just below the pinch suggests that the number of
hot streams must be greater than that of the cold streams. The number feasibility criterion may
be summarized as:

Number of hot streams ≤ Number of cold streams just above pinch (4.14.16)

Number of hot streams ≥ Number of cold streams just below pinch (4.14.17)

Whenever the number criterion is not met, a stream may be split into multiple fractions. It should
be remembered that these feasibility criteria are applicable only at the pinch and not away from the
pinch. As the temperature driving force is higher, these criteria are not directly applicable.
The next step in the design is to determine the duty of each exchanger. By choosing a large duty,
the number of heat exchangers can be minimized. The maximum duty a heat exchanger can have
is the lower of the heat duties of the exchanging streams. This is known as tick-off heuristic. These
steps can be used to design a HEN. Applications of these steps are demonstrated with the illustra-
tive example.
We start with the below pinch portion (see Figure 4.14.13a). There are three streams below pinch.
Both the hot streams start at 70°C, the hot side pinch temperature. The cold stream terminates at
60°C, the cold side pinch temperature. MCp criterion below pinch (Equation 4.14.15) suggests that
C3 with a MCp of 0.15 can be matched with a hot stream with higher MCp. Only H2 with a MCp
of 0.2 matches that. This also satisfies the number criterion (Equation 4.14.17). Therefore, a match
between H2 and C4 is chosen (see Figure 4.14.13b). The heat exchanger in the grid diagram is rep-
resented by a vertical line joining two open circles on the streams that are being matched. H2 has
8 MW of heat available and C3 can accept 3 MW of heat. Therefore, the duty of this exchanger is
3 MW. Now, no cold stream is available. The remaining heat of the hot streams must be removed
by cold utilities. Two coolers, one on H1 and other on H2, are placed as shown in Figure 4.14.13c.
This completes the design of below pinch HEN. It matches with the cold utility target of 6 MW and
expected number of heat exchangers of 3.
Now, we start with the above pinch portion (see Figure 4.14.14a). All four streams are present
above pinch. Both the hot streams terminate at 70°C, the hot side pinch temperature. Both the cold
streams start at 60°C, the cold side pinch temperature. MCp criterion above pinch (Equation 4.14.14)
890 CRC Handbook of Thermal Engineering

(a) (b)
Duty (MW) MCp Duty (MW) MCp
70° 60° 70° 60°
1 H1 0.1 1 H1 0.1
70° 30° 70° 55° 30°
8 H2 0.2 8 H2 0.2
Downloaded By: 10.3.97.143 At: 15:12 28 Oct 2023; For: 9781315119717, sec4_14, 10.4324/9781315119717-47

60° 40° 60° 40°


3 C3 0.15 3 C3 0.15
5 MW
(c)
Duty (MW) MCp
70° 60°
1 H1 C 0.1
70° 55° 1 MW 30°
8 H2 C 0.2
5 MW
60° 40°
3 C3 0.15
3 MW

FIGURE 4.14.13 Design of below pinch HEN: (a) below pinch parallel grid diagram, (b) the first match, and
(c) placement of coolers.

(a) (b)
Duty (MW) Duty (MW) MCp
MCp
210° 130° 70°
210° 70° 14 H1 0.1
14 H1 0.1 70°
180° 70° 180°
22 H2 0.2
22 H2 0.2

100° 60°
100° 60° 6 C3 0.15
6 C3 0.15 6 MW
200° 133.3° 60°
200° 60° 42 C4 0.3
42 C4 0.3
22 MW
(c) (d)
Duty (MW) MCp Duty (MW) MCp
210° 70° 70°
14 H1 0.1 210°
70° 14 H1 0.1
180° 180° 70°
22 H2 0.2 22 H2 0.2

100° 60° 100° 60°


6 C3 0.15 6 C3 H 0.15
200° 60° 200° 6MW 60°
42 C4a 0.3 170° 0.3
42 C4a H
22 MW
C4b 6MW 22 MW
C4b
14MW 14MW

FIGURE 4.14.14 Design of above pinch HEN: (a) above pinch parallel grid diagram, (b) the hidden pinch
phenomenon, (c) stream splitting, and (d) placement of heaters.

suggests that C3 with a MCp of 0.15 can be matched with H1 with a MCp of 0.1 and C4 with a MCp
of 0.3 can be matched with H2 with a MCp of 0.2. Number criterion is automatically satisfied as two
hot streams are matched with two cold streams. Duties of the H1-C3 and H2-C4 matches are 6 MW
(duty of C3) and 22 MW (duty of H2). These matches are shown in Figure 4.14.14b. Now H1 and
C4, with a residual duty of 20 MW, are left unsatisfied. However, have a look at the temperatures of
the remaining streams. 8 MW of heat is available (remaining H1 stream) from 210° to 130°C and
20 MW of heat is required (remaining C4 stream) from 130.3° to 200°C. This is now impossible to
match these two streams without any penalty. This is known as the hidden pinch phenomenon. This
is primarily due to the failure of the tick-off heuristic.
One simple way to avoid hidden pinch phenomenon for this example is to avoid using C3 stream.
In this case, both H1 and H2 have to match with C4. This leads to two hot streams and a single cold
stream, which leads to a violation of the number criterion above pinch (Equation 4.14.16). Therefore,
Applications 891

C4 is split into two (C4a and C4b) and each portion is matched with different hot streams, as shown
in Figure 4.14.14c. Duties of the H1-C4b and H2-C4a matches are decided to be 14 MW (duty of H1)
and 22 MW (duty of H2). As the MCp of C4 is exactly sum of the MCps of the hot streams, MCp of C4a
and C4b must be 0.2 and 0.1, respectively. This also satisfies the MCp criterion. Now, no hot stream is
available. Remaining heat requirement of the cold streams must be satisfied by hot utilities. Two heat-
Downloaded By: 10.3.97.143 At: 15:12 28 Oct 2023; For: 9781315119717, sec4_14, 10.4324/9781315119717-47

ers, one on C3 and other on C4a are placed as shown in Figure 4.14.14d. Two split fractions of C4 are
joined and this completes the design of above pinch HEN. It matches with the hot utility target of 12
MW and expected number of heat exchangers of 4. Complete HEN is shown in Figure 4.14.15.
The pinch design method creates a network structure that satisfies the energy and unit targets.
The minimum temperature driving force for the network is ΔTmin. Based on this initial structure, the
network may be evolved and optimized further. Two important concepts are used while evolving a
given network: loop and path. A loop is defined by a series of connected matches such that it starts
and ends at the same node. One loop is shown in Figure 4.14.16. This loop has the following con-
nections: H(C3), H2-C3, H2-C4, and H(C4). It should be noted that the heaters are connected due
to the presence of a single hot utility. It is interesting to note that every loop in a HEN must contain
an even number of matches. A path is defined by a series of connected matches such that it starts
from a heater and ends at a cooler. One path is shown in Figure 4.14.17. This path has the following
connections: H(C3), H2-C3, and C(H2). It is interesting to note that path in a HEN must contain an
odd number of matches.
Loops and paths can be used to evolve any network. One classical technique is to identify a loop
and distribute the duties of various heat exchanges in the loop. The distribution of the duties may

Pinch line
Above pinch portion 70° Below pinch portion
210° 60°
H1 C
180° 55° 1 MW 30°
H2 C
5 MW
100° 40°
C3 H
200° 6 MW 170° 3 MW
C4 H
6 MW 22 MW
14 MW
60°

FIGURE 4.14.15 Complete heat exchanger network.

Pinch line
Above pinch portion 70° Below pinch portion
210° 60°
H1 C
180° 55° 1 MW 30°
H2 C
5 MW
100° 40°
C3 H
200° 6 MW 170° 3 MW
C4 H
6 MW 22 MW

14 MW
60°

FIGURE 4.14.16 Loop in a heat exchanger network.


892 CRC Handbook of Thermal Engineering

Pinch line
Above pinch portion 70° Below pinch portion
210° 60°
H1 C
180° 55° 1 MW 30°
H2 C
Downloaded By: 10.3.97.143 At: 15:12 28 Oct 2023; For: 9781315119717, sec4_14, 10.4324/9781315119717-47

5 MW
100° 40°
C3 H
200° 6 MW 170° 3 MW
C4 H
6 MW 22 MW
14 MW
60°

FIGURE 4.14.17 Path in a heat exchanger network.

be performed in such a way that the duty of one exchanger may be reduced to zero while the duties
of the other exchangers are positive. This is known as loop breaking. Typically, the heat exchanger
with the minimum duty is removed by loop breaking (Linnhoff and Hindmarsh, 1983). By break-
ing a loop one heat exchanger may be reduced. However, in most cases, this leads to a violation of
the minimum approach temperature for some exchangers (certain cases it may become negative).
A path is then identified through a heat exchanger where the minimum approach temperature is
violated. Heater and cooler duties are increased and duties of other exchangers are adjusted to
restore the minimum approach temperature of the network. Distribution of duties of different heat
exchangers in a path is known as path relaxation. A combination of loop breaking and path relax-
ation is performed to evolve a HEN. In general, the breaking of a loop and recovery of temperature
violations using a path leads to energy transfer across the pinch and consequent increase in energy
requirements. The extra energy penalty incurred can be compared with the decrease in capital cost
due to elimination of an exchanger and if the trade-off is favorable, then the evolved network can be
accepted. The energy penalty incurred depends on the choice of loop and path used for eliminating
an exchanger and there are no guidelines for an appropriate choice of loop and paths for evolution-
ary synthesis. Alternatively, a rigorous optimization-based formulation proposed by Mehta et al.
(2001) using all loops and paths of the process can be formulated which can be used to determine
the least energy penalty for eliminating an exchanger. This formulation can also be used to deter-
mine whether it is feasible to eliminate an exchanger from a given HEN.

Integrating HEN with Other Energy Devices


In a process, electrical energy is used to drive pumps and compressors. The process may also have
a captive power plant for generating electrical power from high-pressure steam. Pinch technology
provides clear rules for heat and power integration which can increase the overall reduction in
energy utilization.
Figure 4.14.18a shows a process represented as a source–sink diagram with HUmin and CUmin being
the minimum utility requirements. Figure 4.14.18a also shows a heat engine, which is part of the pro-
cess (such as a turbine) which converts part of the thermal energy it receives (Qin) into useful work W
and rejects the remaining energy Qout = Qin − W to the environment. The heat engine operates between
the temperatures Tsource and Tsink. The process and heat engine are not heat integrated, and the com-
bined hot utility required by the process and heat engine is HUmin + Qin, while the combined cold util-
ity required is HUmin + Qout. The net useful work obtained is W (in the form of electrical energy) which
may be used in the process itself or may be exported. Figure 4.14.18b–d conceptually depicts different
ways of thermally coupling the heat engine with the process depending on the temperatures, Tsource,
Tsink and Tpinch. It is observed from these diagrams that if the heat engine is heat integrated across pinch,
then there is no net change in the total utility consumption or useful work produced when compared
Applications 893

(a)
HUmin Tsource

Net sink Qin


Total HU = HUmin+Qin
Heat
Total CU = CUmin+Qout engine
Tpinch Pinch boundary W
Downloaded By: 10.3.97.143 At: 15:12 28 Oct 2023; For: 9781315119717, sec4_14, 10.4324/9781315119717-47

Total work = W Qout


Net source

Tsink
CUmin

(b) Tsource
Qin

Heat
HUmin–Qout engine
W

Net sink Qout


Total HU = HUmin+Qin–Qout
Tsink
Total CU = CUmin
Tpinch Pinch boundary
Total work = W
Net source

CUmin

(c) HUmin

Net sink
Total HU = HUmin
Total CU = CUmin–Qin +Qout
Tpinch Pinch boundary
Total work = W
Net source Tsource
CUmin–Qin Qin

Heat
engine
W
Qout

Tsink

FIGURE 4.14.18 Integration of heat engine with process. (a) Process and heat engine not thermally inte-
grated. (b) Heat engine integrated above pinch. (c) Heat engine integrated below pinch.

to the non-integrated scheme of Figure 4.14.18a. However, if the heat engine is integrated by rejecting
heat to the process above pinch or by receiving energy from the process below pinch, then there is net
reduction in overall energy consumption of either Qout or Qin, respectively. It may be noted that the
diagrams represent the maximum energy reduction that can be obtained by integrating the heat engine
with the process. The actual reduction depends on the temperature levels at which the integration is
performed, and the available or required process energy at that temperature level.
It may be noted that a distillation column is similar to a heat engine since it takes in heat from
a reboiler and rejects heat to the condenser, while performing useful work of separating a multi-
component mixture into products of the desired composition. As an example, consider the process
described in Table 4.14.4. A binary distillation column operating at atmospheric pressure is also part
of the process. The bottom product has a bubble point temperature of 120°C and the top product
has a dew point temperature of 90°C. The reboiler duty is 2.2 MW and condenser duty is 2 MW.
One possibility to integrate the column reboiler and condenser with the process HEN is to include
the top product stream as a hot stream and bottom product as a cold stream in the stream table.
The HEN designed for this extended process will also include the reboiler and condenser of the
894 CRC Handbook of Thermal Engineering

distillation column. Alternatively, if the product streams of the column are not included in the pro-
cess stream table, then we can consider the integration of this column with the background process
HEN by treating it like a heat engine. We will assume that a minimum approach temperature of
10°C is required in the reboiler and condenser. Since the condenser temperature is greater than the
pinch temperature, the only possibility is to reject heat to the process above pinch by integrating
Downloaded By: 10.3.97.143 At: 15:12 28 Oct 2023; For: 9781315119717, sec4_14, 10.4324/9781315119717-47

the condenser with the process. In order to satisfy the minimum approach temperature, the cold
process stream receiving the condenser duty should be <80°C. From the grand composite curve
(Figure 4.14.6), we observe that the net energy deficit from pinch temperature up to 80°C is 1.5 MW.
Thus, the maximum amount of energy that can be rejected from the condenser to the background
process is 1.5 MW and the remaining energy of 0.5 MW must be met by an external cold utility. The
net energy savings by heat integrating the column with the process is 1.5 MW. It is also possible to
obtain additional savings by considering the option of increasing the column pressure which will
increase both the reboiler and condenser temperatures. The trade-off, in this case, is the increase
in capital cost of the column (to withstand the increased pressure) and increased pumping costs as
against the reduction in the cost of external utilities.
A similar analysis can be carried out for integrating a heat pump with the HEN. A heat pump
takes in energy from a low-temperature source and does useful work on it and rejects it to a high-
temperature sink. If the source temperature is below the pinch temperature, then the heat pump can
receive energy from the process below pinch. Similarly, if the sink temperature is above pinch, then
the heat pump can reject heat to the process above pinch. Thus, it is best if the heat pump is heat
integrated across the pinch.
This leads to the following pinch technology rules for heat and power integration:

• Integrate a heat engine above or below pinch, but not across the pinch temperature.
• Integrate a heat pump across pinch (receive heat from below pinch and reject heat above pinch).

Conclusion
Pinch analysis was originally developed as a thermodynamic-based approach for conserving ther-
mal energy through a network of heat exchangers (Linnhoff et al., 1982). It was developed as a
holistic tool to analyze overall heat recovery systems for energy conservation in process industries.
Problems of pinch analysis can easily be generalized where streams are characterized by both quan-
tity and quality measures. In the case of heat recovery, energy streams are quantified in terms of
heat duty (quantity) and temperature (quality). The fundamental problem involves minimization of
the externally sourced, high-value resource (e.g., hot utilities in heat recovery pinch analysis) subject
to quality constraints (e.g., temperature constraints due to the second law of thermodynamics). It
is important to note that a consequence of such an optimization is the simultaneous reduction of
unusable waste streams (e.g., rejected heat in heat recovery pinch analysis) (Bandyopadhyay, 2006).
Based on the similarities between heat and mass transfer phenomena, El-Halwagi and Manousiouthakis
(1989) proposed the synthesis of mass exchange networks for the efficient use of industrial mass separat-
ing agents such as solvents, adsorbents, etc. The primary objective is to recover impurity load from a
set of rich streams (process streams containing impurity) to a set of lean streams (process mass sepa-
rating agent(s) that can remove impurity) and thus reducing the use of external mass separating agent.
Advanced problems in a mass exchange network were addressed by Hallale and Fraser (1998).
Further work on mass integration in the 1990s led to the development of pinch analysis techniques
for material resource conservation networks including that for industrial water network (Wang
and Smith, 1994), refinery hydrogen system (Alves, 1999), and property integration (Kazantzi and
El-Halwagi, 2005). In these problems, valuable resources (e.g., water, hydrogen, and other materi-
als) from a set of process sources (stream containing resources) to a set of process sinks (units
requiring resources) are recovered and thus, the use of external resources (such as fresh water, fresh
hydrogen, etc.) is reduced (El-Halwagi, 2011; Foo, 2012).
Applications 895

Tan and Foo (2007) proposed pinch analysis–based methodology for carbon-constrained energy
planning. In carbon-constrained energy planning, energy sources (fossil fuels) are allocated to vari-
ous energy demands (e.g., regions/areas that require energy), in order to reduce the overall CO2
Downloaded By: 10.3.97.143 At: 15:12 28 Oct 2023; For: 9781315119717, sec4_14, 10.4324/9781315119717-47

emission. Carbon capture and storage planning is one of the important issues for carbon-constrained
energy sector planning and applied for country-specific energy sector planning like Ireland (Crilly
and Zhelev, 2008), New Zealand (Atkins et al., 2010), China (Chen et al., 2011), India (Krishna Priya
and Bandyopadhyay, 2013), etc. Techniques of pinch analysis have been extended to study problems
such as to determine the extent of carbon capture to existing power plant that minimizes power loss
(Tan et al., 2009; Sahu et al., 2014), to determine the maximum amount of CO2 to be stored in various
storage systems with given capacity and injectivity constraints (Diamante et al., 2014), etc.
There are numerous other extensions of pinch analysis. Other than the examples listed above,
pinch analysis has also been applied to aggregate production planning problems (Singhvi and Shenoy,
2002), isolated energy systems (Arun et al., 2007; Bandyopadhyay, 2011), human resource planning,
and work scheduling (Foo et al., 2010). Over the years, pinch analysis has established itself as a
structural tool for analyzing and conserving resources in numerous diversified applications, such as
energy sector planning, financial analysis, supply chain management, isolated energy system design,
batch process scheduling, carbon dioxide sequestration, etc. (Linnhoff, 1993; Smith, 2016).

REFERENCES
Ahmad, S., B. Linnhoff, and R. Smith (1990). Cost optimum heat exchanger networks—2. Targets and design
for detailed capital cost models. Computers & Chemical Engineering, 14(7): 751–767.
Alves, J. (1999). Analysis and design of refinery hydrogen systems. Ph.D. Thesis, University of Manchester
Institute of Science and Technology, Manchester, UK.
Arun, P., R. Banerjee, and S. Bandyopadhyay (2007). Sizing curve for design of isolated power systems.
Energy for Sustainable Development, 11: 21–28.
Atkins, M.J., A.S., Morrison, and M.R. Walmsley (2010). Carbon emissions pinch analysis (CEPA) for emis-
sions reduction in the New Zealand electricity sector. Applied Energy, 87: 982–987.
Bade, M.H. and S. Bandyopadhyay (2015). Analysis of gas turbine integrated cogeneration plant: Process
integration approach. Applied Thermal Engineering, 78: 118–128.
Bandyopadhyay, S. (2006). Source composite curve for waste reduction. Chemical Engineering Journal, 125:
99–110.
Bandyopadhyay, S. (2011). Design and optimization of isolated energy systems through pinch analysis. Asia-
Pacific Journal of Chemical Engineering, 6: 518–526.
Bandyopadhyay, S. and G.C. Sahu (2010). Modified problem table. Algorithm for energy targeting. Industrial
& Engineering Chemistry Research, 49: 11557–11563.
Bandyopadhyay, S., J. Varghese, and V. Bansal (2010). Targeting for cogeneration potential through total site
integration. Applied Thermal Engineering, 30(1): 6–14.
Chen, Q., C. Kang, Q. Xia, and D. Guan (2011). Preliminary exploration on low-carbon technology roadmap
of China’s power sector. Energy, 36: 1500–1512.
Crilly, D. and T. Zhelev (2008). Emissions targeting and planning: An application of CO2 emissions pinch
analysis (CEPA) to the Irish electricity generation sector. Energy, 33: 1498–1507.
Diamante, J.A.R., R.R. Tan, D.C.Y. Foo, D.K.S. Ng, K.B. Aviso, and S. Bandyopadhyay (2014). Unified pinch
approach for targeting of carbon capture and storage (CCS) systems with multiple time periods and
regions. Journal of Cleaner Production, 71: 67–74.
El-Halwagi, M.M. (2011). Sustainable Design through Process Integration. Butterworth-Heinemann, Boston, MA.
El-Halwagi, M.M. and V. Manousiouthakis (1989). Synthesis of mass exchange networks. AIChE Journal, 35:
1233–1244.
Foo, D.C.Y. (2012). Process Integration for Resource Conservation. CRC Press, Boca Raton, FL.
Foo, D.C.Y., N. Hallale, and R.R. Tan (2010). Optimize shift scheduling using pinch analysis. Chemical
Engineering, 117: 48–52.
Furman, K.C. and N.V. Sahinidis (2002). A critical review and annotated bibliography for heat exchanger net-
work synthesis in the 20th century. Industrial & Engineering Chemistry Research, 41(10): 2335–2370.
Grip, C.E., J. Isaksson, S. Harvey, and L. Nilsson (2013). Application of pinch analysis in an integrated steel
plant in northern Sweden. ISIJ International, 53(7): 1202–1210.
896 CRC Handbook of Thermal Engineering

Hallale, N. and D.M. Fraser (1998). Capital cost targets for mass exchange networks a special case: Water
minimization. Chemical Engineering Science, 53: 293–313.
Hohmann, E.C. (1971). Optimum networks for heat exchange. Ph.D. Thesis, University of Southern California,
Los Angeles, CA.
Kazantzi, V. and M.M. El-Halwagi (2005). Targeting material use via property integration. Chemical
Engineering Progress, 101: 28–37.
Downloaded By: 10.3.97.143 At: 15:12 28 Oct 2023; For: 9781315119717, sec4_14, 10.4324/9781315119717-47

Kemp, I.C. (2007). Pinch Analysis and Process Integration: A User Guide for the Efficient Use of Energy.
Elsevier, Oxford, UK.
Klemeš, J.J. (Ed.) (2013). Handbook of Process Integration (PI): Minimization of Energy and Water Use,
Waste and Emissions. Woodhead Pub. Ltd., Cambridge, UK.
Krishna Priya, G.S. and S. Bandyopadhyay (2013). Emission constrained power system planning: A pinch anal-
ysis based study of Indian electricity sector. Clean Technologies and Environmental Policy, 15: 771–782.
Linnhoff, B. (1993). Pinch analysis: A state-of-art overview. Transactions of the American Institute of
Chemical Engineers (Part A), 71: 503–522.
Linnhoff, B. and S. Ahmad (1989). Supertargeting: Optimum synthesis of energy management systems.
Journal of Energy Resources Technology, 111(3): 121–130.
Linnhoff, B. and S. Ahmad (1990). Cost optimum heat exchanger networks—1. Minimum energy and capital
using simple models for capital cost. Computers & Chemical Engineering, 14(7): 729–750.
Linnhoff, B. and J.R. Flower (1978). Synthesis of heat exchanger networks: 1. AIChE Journal, 24: 633–642.
Linnhoff, B. and E. Hindmarsh (1983). The pinch design method for heat exchanger networks. Chemical
Engineering Science, 38(5): 745–763.
Linnhoff, B., D.R. Mason, and I. Wardle (1979). Understanding heat exchanger networks. Computers &
Chemical Engineering, 3(1–4): 295–302.
Linnhoff, B., D.W. Townsend, D. Boland, G.F. Hewitt, B.E.A. Thomas, A.R. Guy, and R.H. Marshall, (1982).
A User Guide on Process Integration for the Efficient Use of Energy. Institute of Chemical Engineers,
Rugby, UK.
Mehta, R.K.C., S.K. Devalkar, and S. Narasimhan (2001). An optimization approach for evolutionary synthe-
sis of heat exchanger networks. Transactions of the IChemE, 79: 143:150.
Ozkan, S. and S. Dincer (2001). Application for pinch design of heat exchanger networks by use of a computer code
employing an improved problem algorithm table. Energy Conversion and Management, 42: 2043–2051.
Sahu, G.C., S. Bandyopadhyay, D.C.Y. Foo, D.K.S. Ng, and R.R. Tan (2014). Targeting for optimal grid-
wide deployment of carbon capture and storage (CCS) technology. Process Safety and Environmental
Protection, 92: 835–848.
Salama, A.I.A. (2005). Numerical techniques for determining heat energy targets in pinch analysis. Computers
and Chemical Engineering, 29: 1861–1866.
Salama, A.I.A. (2006). Determination of the optimal heat energy targets in heat pinch analysis using a geom-
etry-based approach. Computers and Chemical Engineering, 30: 758–764.
Shenoy, U.V. (1995). Heat Exchanger Network Synthesis: Process Optimization by Energy and Resource
Analysis. Gulf Publishing, Houston, TX.
Shenoy, U.V., A. Sinha, and S. Bandyopadhyay (1998). Multiple utilities targeting for heat exchanger networks.
Chemical Engineering Research and Design, 76: 259–272.
Singhvi, A. and U.V. Shenoy (2002). Aggregate planning in supply chains by pinch analysis. Chemical
Engineering Research and Design, 80: 597–605.
Smith, R. (2005). Chemical Process Design and Integration, 2nd Edition. John Wiley & Sons Inc, Chichester, UK.
Smith, R. (2016). Chemical Process: Design and Integration. John Wiley & Sons, Chichester, UK.
Svensson, E. and S. Harvey (2011). Pinch analysis of a partly integrated pulp and paper mill. World Renewable
Energy Congress 2011, May 8–13, Sweden.
Tan, R.R. and D.C.Y. Foo (2007). Pinch analysis approach to carbon-constrained energy sector planning.
Energy, 32: 1422–1429.
Tan, R.R., D.K.S. Ng, and D.C.Y. Foo (2009). Pinch analysis approach to carbon-constrained planning for
sustainable power generation. Journal of Cleaner Production, 17: 940–944.
Umeda, T., J. Itoh, and K. Shiroko (1978). Heat exchange system synthesis. Chemical Engineering Progress,
74(7): 70−76.
Upadhyaya, S. (2010). Compilation of energy statistics for economic analysis, Working paper 01/2010,
UNIDO, Vienna.
Varghese, J. and S. Bandyopadhyay (2007). Targeting for energy integration of multiple fired heaters.
Industrial & Engineering Chemistry Research, 46(17): 5631–5644.
Wang, Y.P. and R. Smith (1994). Wastewater minimization. Chemical Engineering Science, 49: 981–1006.

You might also like