You are on page 1of 39

Accepted Manuscript

Title: Locating Exchangers in an EIP-wide Heat Integration


Network

Authors: Sajitha K. Nair, Melvin Soon, I.A. Karimi

PII: S0098-1354(17)30296-X
DOI: http://dx.doi.org/10.1016/j.compchemeng.2017.08.004
Reference: CACE 5871

To appear in: Computers and Chemical Engineering

Received date: 5-5-2017


Revised date: 18-7-2017
Accepted date: 14-8-2017

Please cite this article as: Nair, Sajitha K., Soon, Melvin., & Karimi, I.A., Locating
Exchangers in an EIP-wide Heat Integration Network.Computers and Chemical
Engineering http://dx.doi.org/10.1016/j.compchemeng.2017.08.004

This is a PDF file of an unedited manuscript that has been accepted for publication.
As a service to our customers we are providing this early version of the manuscript.
The manuscript will undergo copyediting, typesetting, and review of the resulting proof
before it is published in its final form. Please note that during the production process
errors may be discovered which could affect the content, and all legal disclaimers that
apply to the journal pertain.
Locating Exchangers in an EIP-wide Heat Integration Network

Sajitha K Nair, Melvin Soon, and I A Karimi*


Department of Chemical & Biomolecular Engineering
National University of Singapore
4 Engineering Drive 4, Singapore 117585


Corresponding author: Email cheiak@nus.edu.sg, Phone +65 6516-6359, Fax +65 6779-1936.

Highlights
 Multi-plant multi-enterprise collaborative heat integration in an eco-industrial park
 Simultaneous intra-plant and inter-plant direct heat integration
 Heat exchangers distributed at plants and centralized sites
 Accounting of ambient heat losses/gains and piping, exchanger, and pumping costs

Abstract

Inter-plant heat integration offers an energy-saving opportunity in EIPs beyond the traditional

intra-plant integration. The economic feasibility of this integration depends critically on the

locations of heat exchangers. In this work, we generalize our previous study on centralized

HENS (Heat Exchanger Network Synthesis) for EIPs to allow exchangers to be located at either

plant or central sites facilitating both intra-plant and inter-plant heat integration in a seamless

manner. We propose a mixed integer non-linear programming (MINLP) model that synthesizes

a maximum-NPV (Net Present Value) EIP-wide HEN, while accounting for all the capital (e.g.

heat exchangers, pumps, and pipelines) and operating (pumping costs, utility savings) cash

flows along with the ambient heat gains/losses during transports. The model is tested on five

examples from the literature and gives better HENs compared to the previous results. This work

highlights and quantifies the impact of heat exchanger locations in an EIP-wide HEN.

Keywords: Eco-Industrial Park, Heat Integration, Mixed-integer Non-linear Programming,

Heat Exchanger Network, Heat Exchanger.


1. Introduction

The world’s population is growing and projected to increase from 6.51 billion in 2005 to 7.79-

10.76 billion in 2050 (Nehring, 2009). Fossil fuels, the main source of energy in the world, are

depleting and causing global warming. These concerns are encouraging global efforts for

energy efficiency and conservation. Heat integration is a well-established technique for energy

conservation in the chemical industry. The heating/cooling demands of streams in a plant can

be synergized to gain economic and environmental benefits. This strategy can also be applied

to a community or complex of closely located plants, called an Eco-Industrial Park (EIP) (Tudor

et al., 2007), to gain environmental, economic, and societal benefits through collaboration

(Manahan, 2004). The collaboration reduces the net consumption of energy, raw materials,

water, and other resources. Some successful EIPs are Kashima Industrial park in Japan, Jurong

Island in Singapore, London Remade eco-industrial sites in the United Kingdom, Ecosite du

Pays de Thau in France (Gibbs et al., 2005; Matsuda et al., 2009).

Interplant heat integration poses unique challenges such as locating exchangers,

transporting streams, terms and conditions of collaboration, and interdependence of plants

(Chen and Lin, 2012; Nair et al., 2016). Hui and Ahmad (1994) developed a nine-step procedure

based on pinch analysis for inter-plant HEN. Instead of using explicit plant-to-plant distances,

they assigned a fixed investment for all plant-to-plant connections. Bagajewicz and Rodera

(2000) used LP (Linear Programming) and MILP (Mixed Integer Linear Programming) models

to integrate four plants, but ignored the critical reality of plant-to-plant distances, transport

costs, and investment for pumps and pipelines. Laukkanen et al. (2012) developed a model for

both direct and indirect heat integration among processes. They assumed that the fixed costs of

the exchangers included the piping costs, which is rather simplistic. Hiete et al. (2012) used

pinch analysis to heat integrate four plants, but did not report any HEN. Cheng et al. (2014)

proposed a game-theoretic strategy where a plant can “trade” a stream with another plant for
heat integration. However, like Bagajewicz and Rodera (2000), Cheng et al. (2014) also ignored

locations, transports, and investments. Wang et al. (2015) proposed a combined direct / indirect

heat integration for two plants. They considered piping costs, but neglected pumps and pumping

costs. Nemet et al. (2016) considered the key issues of pipeline design, heat losses, and pressure

drops, while designing a direct / indirect heat integration network. However, their models were

either too complex or not designed to handle more than two plants.

None of the above interplant heat integration efforts addressed the issue of exchanger

locations. Heat exchanger locations in an EIP-based network are critical, as they affect the

piping and pumping costs and ambient effects. Nair et al. (2016) proposed the idea of locating

all exchangers at a central location and developed an MINLP model for an EIP-wide HEN.

They accounted for the relevant costs of pipes, pumps, exchangers, and pumping, and presented

a shared and rational approach for enterprise collaboration. However, as we show later, forcing

all exchanging streams to gather at a centralized location has economic drawbacks. Chang et

al. (2017) added intervals to the isothermal stagewise superstructure (Yee and Grossmann,

1990) and accounted for exchanger locations, pumping costs, and piping investments. However,

they seem to have unnecessarily complicated the superstructure, as stage numbers can easily be

used as intervals. Furthermore, the assumption of isothermal mixing leads to inferior HEN.

In this work, we propose the idea of a distributed HEN, in which heat exchangers can be

placed at some selected sites or participating plants. By allowing a heat exchanger to be located

at a plant, we can eliminate the movement of streams that already belong to that plant. In

addition, we account for pressure and ambient heat gains/losses. We present a mixed integer

nonlinear programming (MINLP) model that maximizes the NPV for the EIP-wide HEN.

2. Problem Statement

E independent enterprises (𝑒 = 1, 2, … , 𝐸) are interested in an EIP-wide heat integration

project that can save utility costs by conserving energy. Each enterprise has one or more plants
with streams that are currently heated or cooled internally by some utilities (e.g. steam, cooling

water, ...). The proposed project aims to match the heating/cooling demands of these streams

across the EIP to reduce the total utility usage in a cost-effective manner. It involves

constructing a cost-effective and mutually agreeable network of 2-stream exchangers that

enable hot process streams (those that need cooling) from various EIP plants to heat cold

process streams (those that need heating). Hence, there is a need to (1) identify the most cost-

effective HEN for the EIP, (2) locate its exchangers at optimal places, and (3) estimate the costs

and benefits for the various enterprises that may participate in the project. It is agreed that an

exchanger may be located at either a pre-designated centralized site or any plant in the EIP.

With this, the problem for synthesizing an EIP-wide HEN can be stated as follows.

Given:

1) 𝐼 hot EIP streams (𝑖 = 1, 2, … , 𝐼) to be cooled and 𝐽 cold EIP streams (𝑗 = 1, 2, … , 𝐽) to

be heated.

2) Their mass flows, initial and target temperatures, and allowable temperature limits (lower

and upper) of all streams.

3) Their thermophysical properties such as density, heat capacity, viscosity, film heat

transfer coefficients (HTCs).

4) 𝑃 + 1 plants (𝑝 = 1, 2, … , 𝑃 + 1), where exchangers can be located. Here, the first

𝑃 plants (𝑝 = 1, 2, … , 𝑃) are the actual plants belonging to various enterprises, and the last plant

(𝑝 = 𝑃 + 1) is a dummy plant representing the centralized site. The dummy plant belongs to

all 𝐸 enterprises.

5) Plant-to-plant distance for each pair of plants (𝑝 = 1, 2, … , 𝑃 + 1).

6) Minimum temperature approach and pressure drops for each potential 2-stream heat

exchanger.

7) Unit costs and temperature ranges of hot and cold utilities.


8) Purchase and installation costs of pipes, heat exchangers, pumps, and compressors.

9) Estimated life time (N) for the project and annual interest rate.

Obtain:

1) The EIP-wide HEN with participating streams, their temperatures, and flows.

2) Heat exchangers, their streams, duties, areas, and locations.

3) Total capital expenditure or CAPEX for the project.

4) Total operating expenditure or OPEX for the project.

5) Participating enterprises, their CAPEX and OPEX contributions, and IROR (Internal Rate

of return on investment).

Aiming to:

1) Maximize the net present value (NPV) for the project.

Assuming:

1) Inter-plant flows of streams cannot be partial. A stream may split into one or more

substreams after entering a plant, but the substreams must merge again before leaving that plant.

This is to minimize the cost of transporting streams. Transporting multiple substreams would

increase piping and transport costs, and may pose operational issues.

2) All 2-stream heat exchangers are countercurrent.

3) Existing in-plant heaters/coolers for the streams are retained or retrofitted to provide

operational robustness. Retrofit costs are zero.

4) A stream either gains heat from or loses heat to the ambient during its movement from

one plant to another in the EIP, and its gain/loss behavior remains unchanged through its entire

movement.

5) Existing in-plant movers (pumps or compressors) for the streams are retained or

retrofitted to transport streams internally. New pumps are installed for transporting streams to

other plants. Only one new pump is used for such transport.
6) All matches between hot and cold streams are allowable, if thermodynamically feasible.

7) Stream properties and HTCs are constants.

8) All capital expenses occur at time zero.

9) Inflation rate is zero, so utility and transportation costs remain constant throughout the 𝑁

years.

3. Model Formulation

We assume a superstructure (Huang and Karimi, 2013; Yee and Grossmann, 1990) with 𝐾 + 2

stages (𝑘 = 0,1, … , 𝐾, 𝐾 + 1) as shown in Figure 1, where 𝐾 is predefined. The hot streams

pass through stages 1 through 𝐾 + 1 before exiting the HEN. In contrast, the cold streams pass

through stages 𝐾 through 0 before exiting the HEN. However, a stream may bypass one or more

stages from 1 to 𝐾 entirely, if needed. If a hot (cold) stream enters stage k (1 ≤ 𝑘 ≤ 𝐾), then it

splits into 𝐽 (𝐼) substreams to exchange heat with 𝐽 (𝐼) cold (hot) substreams in 2-stream

exchangers. Thus, stage 𝑘 (1 ≤ 𝑘 ≤ 𝐾) can have at most 𝐼 × 𝐽 exchangers. After exchanging

heat, the substreams merge back to remake the parent stream before exiting the stage. Each hot

(cold) stream enters stage 𝐾 + 1 (0) to be cooled (heated) with utilities to meet its target

temperature.

Let 𝑠 (= 𝑖 𝑜𝑟 𝑗) represent an EIP stream with 1 ≤ 𝑠 ≤ 𝑆 = 𝐼 + 𝐽. The HEN and the

streams may not need all the 𝐾 stages. Therefore, we allow a stream to bypass a stage (1 ≤ 𝑘 ≤

𝐾) entirely with the help of the following 0-1 continuous variable.

1 if stream 𝑠 enters stage 𝑘


𝑦𝑠𝑘 = { 1 ≤ 𝑠 ≤ 𝑆; 1 ≤ 𝑘 ≤ 𝐾
0 otherwise

Clearly, a stage can exist, only if at least one stream enters that stage. We force the redundant

stages to be the end ones by defining the following 0-1 continuous variable.

1 if stage 𝑘 exists
𝑌𝑘 = { 1≤𝑘≤𝐾
0 otherwise

𝑌𝑘 ≥ 𝑌𝑘+1 1 ≤𝑘 ≤𝐾−1 (1a)


𝑦𝑠𝑘 ≤ 𝑌𝑘 ≤ ∑𝑆𝑠′=1 𝑦𝑠′𝑘 1 ≤ 𝑠 ≤ 𝑆; 1 ≤ 𝑘 ≤ 𝐾 (1b)

Let 𝐻𝐸𝑖𝑗𝑘 (1 ≤ 𝑖 ≤ 𝐼; 1 ≤ 𝑗 ≤ 𝐽; 1 ≤ 𝑘 ≤ 𝐾) be the 2-stream heat exchanger in which

a part or whole of hot stream 𝑖 heats a part or whole of cold stream 𝑗 in stage 𝑘. We define a 0-

1 continuous variable 𝑥𝑖𝑗𝑘 (1 ≤ 𝑖 ≤ 𝐼; 1 ≤ 𝑗 ≤ 𝐽; 1 ≤ 𝑘 ≤ 𝐾) to denote the existence of 𝐻𝐸𝑖𝑗𝑘 .

1 if 𝐻𝐸𝑖𝑗𝑘 exists
𝑥𝑖𝑗𝑘 = { 1 ≤ 𝑖 ≤ 𝐼; 1 ≤ 𝑗 ≤ 𝐽; 1 ≤ 𝑘 ≤ 𝐾
0 otherwise

For simplicity, we limit the number of substreams in each stage.

∑𝐽𝑗=1 𝑥𝑖𝑗𝑘 ≤ 𝑆𝑆𝑖 1 ≤ 𝑖 ≤ 𝐼; 1 ≤ 𝑘 ≤ 𝐾 (2a)

∑𝐼𝑖=1 𝑥𝑖𝑗𝑘 ≤ 𝑆𝑆𝑗 1 ≤ 𝑗 ≤ 𝐽; 1 ≤ 𝑘 ≤ 𝐾 (2b)

It is possible that two stages have identical stream matches, i.e. the same combinations of hot

and cold substreams exchange heat. In this case, we can merge the two stages into a single stage

by demanding that a stream can exchange heat with the same stream in two consecutive stages

only if it also exchanges heat with another substream in one of the stages.

𝑥𝑖𝑗𝑘 + 𝑥𝑖𝑗(𝑘+1) − ∑𝐽𝑗′ =1,𝑗 ′ ≠𝑗 𝑥𝑖𝑗 ′ 𝑘 − ∑𝐽𝑗′ =1,𝑗 ′ ≠𝑗 𝑥𝑖𝑗 ′ (𝑘+1) − ∑𝐼𝑖′ =1,𝑖 ′ ≠𝑖 𝑥𝑖 ′ 𝑗𝑘 −

∑𝐼𝑖′ =1,𝑖 ′ ≠𝑖 𝑥𝑖 ′ 𝑗(𝑘+1) ≤ 1 1 ≤ 𝑖 ≤ 𝐼; 1 ≤ 𝑗 ≤ 𝐽; 1 ≤ 𝑘 ≤ 𝐾 − 1 (3)

Let 𝑓𝑖𝑗𝑘 (1 ≤ 𝑖 ≤ 𝐼; 1 ≤ 𝑗 ≤ 𝐽; 1 ≤ 𝑘 ≤ 𝐾) and 𝑔𝑖𝑗𝑘 (1 ≤ 𝑖 ≤ 𝐼; 1 ≤ 𝑗 ≤ 𝐽; 1 ≤ 𝑘 ≤ 𝐾)

respectively be the fractional flows of streams 𝑖 and 𝑗 entering 𝐻𝐸𝑖𝑗𝑘 . For a stream to flow

through an exchanger, the exchanger must exist, and the stream must enter the stage.

Furthermore, if a stream enters the stage, the fractional flows must sum to one. Therefore,

𝑦𝑖𝑘 = ∑𝐽𝑗=1 𝑓𝑖𝑗𝑘 1 ≤ 𝑖 ≤ 𝐼; 1 ≤ 𝑘 ≤ 𝐾 (4a)

𝑓𝑖𝑗𝑘 ≤ 𝑥𝑖𝑗𝑘 ≤ 𝑦𝑖𝑘 1 ≤ 𝑖 ≤ 𝐼; 1 ≤ 𝑗 ≤ 𝐽; 1 ≤ 𝑘 ≤ 𝐾 (4b)

𝑦𝑗𝑘 = ∑𝐼𝑖=1 𝑔𝑖𝑗𝑘 1 ≤ 𝑗 ≤ 𝐽; 1 ≤ 𝑘 ≤ 𝐾 (4c)

𝑔𝑖𝑗𝑘 ≤ 𝑥𝑖𝑗𝑘 ≤ 𝑦𝑗𝑘 1 ≤ 𝑖 ≤ 𝐼; 1 ≤ 𝑗 ≤ 𝐽; 1 ≤ 𝑘 ≤ 𝐾 (4d)

Exchanger Locations

Locating exchangers at right plants in an EIP is crucial, as the key costs of pumping and
piping depend heavily on the distances that the heat-exchanging streams must travel. An

exchanger can only be placed in a plant that accepts to host it. Hence, we define,

𝑃𝑠 = {𝑝 | 𝑝𝑙𝑎𝑛𝑡 𝑝 𝑚𝑎𝑦 𝑝𝑜𝑠𝑠𝑖𝑏𝑙𝑦 ℎ𝑜𝑠𝑡 𝑎𝑛 𝑒𝑥𝑐ℎ𝑎𝑛𝑔𝑒𝑟 𝑤𝑖𝑡ℎ 𝑠𝑡𝑟𝑒𝑎𝑚 𝑠}

𝑝𝑠 = 𝑃𝑙𝑎𝑛𝑡 𝑡ℎ𝑎𝑡 𝑜𝑤𝑛𝑠 𝑠𝑡𝑟𝑒𝑎𝑚 𝑠.

Then, we define a binary variable 𝑎𝑖𝑗𝑘𝑝 for locating 𝐻𝐸𝑖𝑗𝑘 .

1 if 𝐻𝐸𝑖𝑗𝑘 is located at plant 𝑝


𝑎𝑖𝑗𝑘𝑝 = { 1 ≤ 𝑖 ≤ 𝐼; 1 ≤ 𝑗 ≤ 𝐽; 1 ≤ 𝑘 ≤ 𝐾, 𝑝 ∈ 𝑃𝑖 ∩ 𝑃𝑗
0 otherwise

∑𝑝∈𝑃𝑖 ∩𝑃𝑗 𝑎𝑖𝑗𝑘𝑝 = 𝑥𝑖𝑗𝑘 1 ≤ 𝑖 ≤ 𝐼; 1 ≤ 𝑗 ≤ 𝐽; 1 ≤ 𝑘 ≤ 𝐾 (5)

Eq. 5 fixes 𝑥𝑖𝑗𝑘 to be binary. Now, we identify the location of each stream at each stage using

the following 0-1 continuous variables.

1 if stream 𝑠 with all its substreams is located at plant 𝑝 in stage 𝑘


𝑧𝑠𝑘𝑝 = {
0 otherwise

1 ≤ 𝑠 ≤ 𝑆; 0 ≤ 𝑘 ≤ 𝐾 + 1; 𝑝 ∈ 𝑃𝑠

Because a stream must begin from and return to its own plant, we set 𝑧𝑠(𝐾+1)𝑝𝑠 = 𝑧𝑠0𝑝𝑠 = 1,

and 𝑧𝑠(𝐾+1)𝑝 = 𝑧𝑠0𝑝 = 0 for 𝑝 ∈ 𝑃𝑠 and 𝑝 ≠ 𝑝𝑠 . Furthermore, we demand that stream 𝑠 must

be at one plant in each stage 𝑘, even if it skips (𝑦𝑠𝑘 = 0) stage 𝑘 or does not even have it.

∑𝑝∈𝑃𝑠 𝑧𝑠𝑘𝑝 = 1 1 ≤ 𝑠 ≤ 𝑆; 1 ≤ 𝑘 ≤ 𝐾 (6)

Furthermore, a stream must also be at the same plant where its exchanger is.

𝑧𝑖𝑘𝑝 ≥ 𝑎𝑖𝑗𝑘𝑝 1 ≤ 𝑖 ≤ 𝐼; 1 ≤ 𝑗 ≤ 𝐽; 1 ≤ 𝑘 ≤ 𝐾; 𝑝 ∈ 𝑃𝑖 ∩ 𝑃𝑗 (7a)

𝑧𝑗𝑘𝑝 ≥ 𝑎𝑖𝑗𝑘𝑝 1 ≤ 𝑖 ≤ 𝐼; 1 ≤ 𝑗 ≤ 𝐽; 1 ≤ 𝑘 ≤ 𝐾; 𝑝 ∈ 𝑃𝑖 ∩ 𝑃𝑗 (7b)

Having located the streams at each stage, we now track their movements through the HEN by:

1 if stream 𝑖 is at 𝑝′ at stage 𝑘 − 1 and 𝑝 at stage 𝑘


𝑍𝑖𝑘𝑝′𝑝 = { = 𝑧𝑖(𝑘−1)𝑝′ 𝑧𝑖𝑘𝑝
0 otherwise

1 ≤ 𝑖 ≤ 𝐼; 1 ≤ 𝑘 ≤ 𝐾 + 1; 𝑝 ∈ 𝑃𝑖 , 𝑝′ ∈ 𝑃𝑖

1 if stream 𝑗 is at 𝑝′ at stage 𝑘 + 1 and 𝑝 at stage 𝑘


𝑍𝑗𝑘𝑝′𝑝 = { = 𝑧𝑗(𝑘+1)𝑝′ 𝑧𝑗𝑘𝑝
0 otherwise

1 ≤ 𝑗 ≤ 𝐽; 0 ≤ 𝑘 ≤ 𝐾; 𝑝 ∈ 𝑃𝑗 , 𝑝′ ∈ 𝑃𝑗
We linearize 𝑍𝑠𝑘𝑝′𝑝 as follows,

∑𝑝∈𝑃𝑖 𝑍𝑖𝑘𝑝′𝑝 = 𝑧𝑖(𝑘−1)𝑝′ 1 ≤ 𝑖 ≤ 𝐼; 1 ≤ 𝑘 ≤ 𝐾 + 1; 𝑝′ ∈ 𝑃𝑖 (8a)

∑𝑝′ ∈𝑃𝑖 𝑍𝑖𝑘𝑝′𝑝 = 𝑧𝑖𝑘𝑝 1 ≤ 𝑖 ≤ 𝐼; 1 ≤ 𝑘 ≤ 𝐾 + 1; 𝑝 ∈ 𝑃𝑖 (8b)

∑𝑝∈𝑃𝑗 𝑍𝑗𝑘𝑝′𝑝 = 𝑧𝑗(𝑘+1)𝑝′ 1 ≤ 𝑗 ≤ 𝐽; 0 ≤ 𝑘 ≤ 𝐾; 𝑝′ ∈ 𝑃𝑗 (8c)

∑𝑝′ ∈𝑃𝑗 𝑍𝑗𝑘𝑝′𝑝 = 𝑧𝑗𝑘𝑝 1 ≤ 𝑗 ≤ 𝐽; 0 ≤ 𝑘 ≤ 𝐾; 𝑝 ∈ 𝑃𝑗 (8d)

If a stream does not enter stage 𝑘, then we keep it at its current location to minimize unnecessary

transport. Hence,

∑𝑝∈𝑃𝑠 𝑍𝑠𝑘𝑝𝑝 ≥ 1 − 𝑦𝑠𝑘 1 ≤ 𝑠 ≤ 𝑆; 1 ≤ 𝑘 ≤ 𝐾 (9)

Stream Temperatures and Exchanger Duties

Let

𝐹𝑖 = Heat content (flow times heat capacity) of hot stream 𝑖 (1 ≤ 𝑖 ≤ 𝐼)

𝐺𝑗 = Heat content of cold stream 𝑗 (1 ≤ 𝑗 ≤ 𝐽)

𝑇𝐼𝑁𝑠 = Initial temperature of stream 𝑠 (1 ≤ 𝑠 ≤ 𝑆)

𝑇𝑂𝑈𝑇𝑠 = Target temperature of stream 𝑠 (1 ≤ 𝑠 ≤ 𝑆)

𝑄𝑖𝑗𝑘 = Duty of 𝐻𝐸𝑖𝑗𝑘 (1 ≤ 𝑖 ≤ 𝐼; 1 ≤ 𝑗 ≤ 𝐽; 1 ≤ 𝑘 ≤ 𝐾)

𝑄𝑖𝑘 = Total heat duty of hot stream 𝑖 in stage 𝑘 (1 ≤ 𝑖 ≤ 𝐼, 1 ≤ 𝑘 ≤ 𝐾 + 1)

𝑄𝑗𝑘 = Total heat duty of cold stream 𝑗 in stage 𝑘 (1 ≤ 𝑗 ≤ 𝐽, 0 ≤ 𝑘 ≤ 𝐾)

𝐷𝑖𝑗𝑘 = Temperature drop of hot substream 𝑖 in 𝐻𝐸𝑖𝑗𝑘 (1 ≤ 𝑖 ≤ 𝐼; 1 ≤ 𝑗 ≤ 𝐽; 1 ≤

𝑘 ≤ 𝐾)

𝑅𝑖𝑗𝑘 = Temperature rise of cold substream 𝑗 in 𝐻𝐸𝑖𝑗𝑘 (1 ≤ 𝑖 ≤ 𝐼; 1 ≤ 𝑗 ≤ 𝐽; 1 ≤

𝑘 ≤ 𝐾)

𝐷𝑖𝐿 = Minimum desired temperature change in substream 𝑖 at any stage (1 ≤ 𝑖 ≤ 𝐼)

𝑅𝑗𝐿 = Minimum desired temperature change in substream 𝑗 at any stage (1 ≤ 𝑗 ≤ 𝐽)

Let stream 𝑠 enter stage 𝑘 at temperature 𝑇𝑠𝑘 and leave stage 𝑘 at 𝑇𝐸𝑠𝑘 . Thus, its
temperature may change, as it moves from one stage to another due to ambient heat losses/gains.

Assuming that the temperature change is linearly proportional to the distance travelled by 𝑠, we

obtain,

𝑇𝑖𝑘 − 𝑇𝐸𝑖(𝑘−1) = ∆𝑇𝑖 ∑𝑝∈𝑃𝑖 ∑𝑝′ ∈𝑃𝑖 (𝑍𝑖𝑘𝑝′ 𝑝 𝑑𝑑𝑝′ 𝑝 ) 1 ≤ 𝑖 ≤ 𝐼; 2 ≤ 𝑘 ≤ 𝐾 + 1 (10a)

𝑇𝑖1 − 𝑇𝐼𝑁𝑖 = ∆𝑇𝑖 ∑𝑝∈𝑃𝑖 ∑𝑝′ ∈𝑃𝑖(𝑍𝑖1𝑝′ 𝑝 𝑑𝑑𝑝′ 𝑝 ) 1≤𝑖≤𝐼 (10b)

𝑇𝑗𝑘 − 𝑇𝐸𝑗(𝑘+1) = ∆𝑇𝑗 ∑𝑝∈𝑃𝑗 ∑𝑝′ ∈𝑃𝑗 (𝑍𝑗𝑘𝑝′ 𝑝 𝑑𝑑𝑝′ 𝑝 ) 1 ≤ 𝑗 ≤ 𝐽; 0 ≤ 𝑘 ≤ 𝐾 − 1 (11a)

𝑇𝑗𝐾 − 𝑇𝐼𝑁𝑗 = ∆𝑇𝑗 ∑𝑝∈𝑃𝑗 ∑𝑝′ ∈𝑃𝑗 (𝑍𝑗𝐾𝑝′ 𝑝 𝑑𝑑𝑝′ 𝑝 ) 1≤𝑗≤𝐽 (11b)

where, ∆𝑇𝑠 is the average rise/drop in the temperature of stream 𝑠 per unit distance and

𝑑𝑑𝑝′ 𝑝 (1 ≤ 𝑝, 𝑝′ ≤ 𝑃 + 1) is the distance between plants 𝑝 and 𝑝′ . ∆𝑇𝑠 is positive, if 𝑠 gains

heat from the ambient, and vice versa.

Nair et al. (2016) allowed two utility exchangers for each stream at stages 0 and (𝐾 + 1). In

contrast, we allow only one utility exchanger to be placed at the HEN exit. Hence, 𝑄𝑖(𝐾+1)

represents the utility cooling of stream 𝑖, and 𝑄𝑗0 represents the utility heating of stream 𝑗. Also,

we have modified the expressions for temperature changes.

𝐷𝑖𝐿 𝑥𝑖𝑗𝑘 ≤ 𝐷𝑖𝑗𝑘 ≤ 𝐷𝑖𝑗𝑈 𝑥𝑖𝑗𝑘 1 ≤ 𝑖 ≤ 𝐼; 1 ≤ 𝑗 ≤ 𝐽; 1 ≤ 𝑘 ≤ 𝐾 (12a)

𝑅𝑗𝐿 𝑥𝑖𝑗𝑘 ≤ 𝑅𝑖𝑗𝑘 ≤ 𝑅𝑖𝑗


𝑈
𝑥𝑖𝑗𝑘 1 ≤ 𝑖 ≤ 𝐼; 1 ≤ 𝑗 ≤ 𝐽; 1 ≤ 𝑘 ≤ 𝐾 (12b)

Then, the heat exchanged by 𝑠 at stage 𝑘 is given by,

𝑄𝑖𝑘 = ∑𝐽𝑗=1 𝑄𝑖𝑗𝑘 1 ≤ 𝑖 ≤ 𝐼; 1 ≤ 𝑘 ≤ 𝐾 (13a)

𝑄𝑗𝑘 = ∑𝐼𝑖=1 𝑄𝑖𝑗𝑘 1 ≤ 𝑗 ≤ 𝐽; 1 ≤ 𝑘 ≤ 𝐾 (13b)

𝑄𝑖𝑈 = 𝐹𝑖 (𝑇𝐼𝑁𝑖 − 𝑇𝑂𝑈𝑇𝑖 ) = ∑𝐾+1 𝐾+1


𝑘=1 [𝑄𝑖𝑘 ] + ∑𝑘=2 [𝐹𝑖 (𝑇𝐸𝑖(𝑘−1) − 𝑇𝑖𝑘 )] + 𝐹𝑖 (𝑇𝐼𝑁𝑖 − 𝑇𝑖1 )]

1≤𝑖≤𝐼 (14a)

𝑄𝑗𝑈 = 𝐺𝑗 (𝑇𝑂𝑈𝑇𝑗 − 𝑇𝐼𝑁𝑗 ) = ∑𝐾 𝐾−1


𝑘=0[𝑄𝑗𝑘 ] − ∑𝑘=0 [𝐺𝑗 (𝑇𝐸𝑗(𝑘+1) − 𝑇𝑗𝑘 )] − 𝐺𝑗 (𝑇𝐼𝑁𝑗 − 𝑇𝑗𝐾 )

1≤𝑗≤𝐽 (14b)

𝑄𝑖𝑗𝑘 ≤ 𝑓𝑖𝑗𝑘 𝐹𝑖 𝐷𝑖𝑗𝑘 1 ≤ 𝑖 ≤ 𝐼; 1 ≤ 𝑗 ≤ 𝐽; 1 ≤ 𝑘 ≤ 𝐾 (15a)


𝑄𝑖𝑗𝑘 ≤ 𝑔𝑖𝑗𝑘 𝐺𝑗 𝑅𝑖𝑗𝑘 1 ≤ 𝑖 ≤ 𝐼; 1 ≤ 𝑗 ≤ 𝐽; 1 ≤ 𝑘 ≤ 𝐾 (15b)

𝑄𝑖𝑗𝑘 ≤ 𝑚𝑖𝑛{𝑄𝑖𝑈 , 𝑄𝑗𝑈 , 𝐹𝑖 𝐷𝑖𝑗𝑈 , 𝐺𝑗 𝑅𝑖𝑗


𝑈
}𝑥𝑖𝑗𝑘 1 ≤ 𝑖 ≤ 𝐼; 1 ≤ 𝑗 ≤ 𝐽; 1 ≤ 𝑘 ≤ 𝐾 (16)

Then, the heat duty and area of exchanger 𝐻𝐸𝑖𝑗𝑘 are given by,

𝑄𝑖𝑗𝑘 ≤ 𝑈𝑖𝑗 𝐴𝑖𝑗𝑘 𝐿𝑀𝑇𝐷𝑖𝑗𝑘 1 ≤ 𝑖 ≤ 𝐼; 1 ≤ 𝑗 ≤ 𝐽; 1 ≤ 𝑘 ≤ 𝐾 (17a)

𝐴𝐿𝑖𝑗 ≤ 𝐴𝑖𝑗𝑘 ≤ 𝐴𝐿𝑖𝑗 + (𝐴𝑈𝑖𝑗 − 𝐴𝐿𝑖𝑗 )𝑥𝑖𝑗𝑘 1 ≤ 𝑖 ≤ 𝐼; 1 ≤ 𝑗 ≤ 𝐽; 1 ≤ 𝑘 ≤ 𝐾 (17b)

where, 𝑈𝑖𝑗 (1 ≤ 𝑖 ≤ 𝐼; 1 ≤ 𝑗 ≤ 𝐽) is the overall heat transfer coefficient, 𝐿𝑀𝑇𝐷𝑖𝑗𝑘 (1 ≤ 𝑖 ≤

𝐼; 1 ≤ 𝑗 ≤ 𝐽; 1 ≤ 𝑘 ≤ 𝐾) is the Logarithmic Mean Temperature Difference (LMTD),

𝐴𝑖𝑗𝑘 (1 ≤ 𝑖 ≤ 𝐼; 1 ≤ 𝑗 ≤ 𝐽; 1 ≤ 𝑘 ≤ 𝐾) is the area, and 𝐴𝐿𝑖𝑗 (1 ≤ 𝑖 ≤ 𝐼; 1 ≤ 𝑗 ≤ 𝐽) and

𝐴𝑈𝑖𝑗 (1 ≤ 𝑖 ≤ 𝐼; 1 ≤ 𝑗 ≤ 𝐽) are the lower and upper bounds of 𝐴𝑖𝑗𝑘 . Appendix A gives the

remaining equations for the temperature changes and approaches from Nair et al. (2016).

NPV for HEN

We now compute the capital and operating expenses (CAPEX and OPEX) of HEN to obtain its

NPV. 𝐶𝐴𝑃𝐸𝑋 is the total capital cost invested at time zero. It includes the cost of purchasing

and installing the pumps, heat exchangers, and pipelines. As the enterprises retain their existing

heaters/coolers and pumps for each stream, we do not include their costs. For the new pumps

required for transporting streams out of their parent plants, we use the following 0-1 continuous

variable.

1 if stream 𝑠 needs a new pump


𝑁𝑃𝑠 = { 1≤𝑠≤𝑆
0 otherwise

1 − 𝑧𝑠𝑘𝑝𝑠 ≤ 𝑁𝑃𝑠 ≤ 𝐾 − ∑𝐾
𝑘 ′ =1 𝑧𝑠𝑘 ′ 𝑝𝑠 1 ≤ 𝑠 ≤ 𝑆; 1 ≤ 𝑘 ≤ 𝐾 (18)

Then, we have,
𝐽
𝐶𝐴𝑃𝐸𝑋 = ∑𝑆𝑠=1[𝐶𝑃𝐿𝑠 𝐿𝑃𝐿𝑠 + 𝑀𝑃𝑠 𝐶𝑃𝑠 ] + ∑𝐾 𝐼
𝑘=1 ∑𝑖=1 ∑𝑗=1 𝑀𝐻𝑖𝑗 𝐻𝐸𝐶𝑖𝑗𝑘 (19a)

𝛽𝑖𝑗 𝛽𝑖𝑗
𝐻𝐸𝐶𝑖𝑗𝑘 = 𝐹𝐶𝐻𝑖𝑗 𝑥𝑖𝑗𝑘 + 𝑉𝐶𝐻𝑖𝑗 (𝐴𝑖𝑗𝑘 ) − 𝑉𝐶𝐻𝑖𝑗 (𝐴𝐿𝑖𝑗 ) (1 − 𝑥𝑖𝑗𝑘 )

1 ≤ 𝑖 ≤ 𝐼; 1 ≤ 𝑗 ≤ 𝐽; 1 ≤ 𝑘 ≤ 𝐾 (19b)
𝐿𝑃𝐿𝑖 = ∑𝐾+1
𝑘=1 ∑𝑝∈𝑃𝑖 ∑𝑝′ ∈𝑃𝑖 (𝑍𝑖𝑘𝑝′ 𝑝 𝑑𝑑𝑝′ 𝑝 ) 1≤𝑖≤𝐼 (19c)

𝐿𝑃𝐿𝑗 = ∑𝐾
𝑘=0 ∑𝑝∈𝑃𝑗 ∑𝑝′ ∈𝑃𝑗(𝑍𝑗𝑘𝑝′ 𝑝 𝑑𝑑𝑝′ 𝑝 ) 1≤𝑗≤𝐽 (19d)

where,

𝐶𝑃𝐿𝑠 = installed cost of pipeline per unit length

𝐿𝑃𝐿𝑠 = length of pipeline for stream 𝑠

𝑀𝑃𝑠 = installation/purchase cost multiplier for a pump for stream 𝑠

𝐶𝑃𝑠 = Purchase cost of a pump for stream 𝑠 with 𝐹𝐶𝑃𝑠 as the fixed component, 𝑉𝐶𝑃𝑠 as the

coefficient in the variable component, and 𝛼𝑠 as the exponent

𝑀𝐻𝑖𝑗 = installation/purchase cost multiplier for 𝐻𝐸𝑖𝑗𝑘

𝐻𝐸𝐶𝑖𝑗𝑘 = Purchase cost of heat exchanger 𝐻𝐸𝑖𝑗𝑘 with 𝐹𝐶𝐻𝑖𝑗 as the fixed component, 𝑉𝐶𝐻𝑖𝑗 as

the coefficient in the variable component, and 𝛽𝑖𝑗 as the exponent

For costing new pumps, we compute the total pump power for each stream as follows.

𝜂𝑠 𝐸𝑠 = 𝑉𝐹𝑠 [∆𝑝𝑠𝐿 𝐿𝑃𝐿𝑠 + ∑𝐾 𝐻𝐸


𝑘=1 𝑦𝑠𝑘 ∆𝑝𝑠 ] 1≤𝑠≤𝑆 (20)

where, ∆𝑝𝑠𝐻𝐸 is the pressure drop for stream 𝑠 in one heat exchanger, ∆𝑝𝑠𝐿 is the pressure drop

for stream 𝑠 per unit length of its pipeline, 𝑉𝐹𝑠 is the volumetric flow rate (𝑚3 /𝑎) of stream 𝑠,

and 𝜂𝑠 is the pump efficiency. Then, the purchase cost of a new pump for stream 𝑠 is,

𝐶𝑃𝑠 ≥ 𝐹𝐶𝑃𝑠 𝑁𝑃𝑠 + 𝑉𝐶𝑃𝑠 (𝐸𝑠 )𝛼𝑠 − 𝑉𝐶𝑃𝑠 (𝐾 ∙ 𝑉𝐹𝑠 ∙ ∆𝑝𝑠𝐻𝐸 /𝜂𝑠 )𝛼𝑠 (1 − 𝑁𝑃𝑠 )

1≤𝑠≤𝑆 (21)

When stream 𝑠 needs a new pump (𝑁𝑃𝑠 = 1), then eq. 21 will give its cost. If it does not, then

eq. 21 will give a zero cost, as stream 𝑠 will use its own in-plant pump.

The OPEX for the HEN is the annual cost of moving streams, which is 𝐶𝑂𝐸 ∑𝑆𝑠=1 𝐸𝑠 ,

where 𝐶𝑂𝐸 is the cost of electricity ($/kW). The use of the HEN may save utility costs annually

for the EIP. We compute the NPV for these annual OPEX and utility savings by using a factor

𝛾 as done by Nair et al. (2016). Then, the objective for the EIP-wide HEN project is to maximize
NPV.

(1+𝑟)𝑁 −1
𝑀𝑎𝑥 𝑁𝑃𝑉 = {∑𝐽𝑗=1[𝑈𝐶𝑗 (𝑄𝑗𝑈 − 𝑄𝑗0 )] + ∑𝐼𝑖=1[𝑈𝐶𝑖 (𝑄𝑖𝑈 − 𝑄𝑖(𝐾+1) )] −
𝑟(1+𝑟)𝑁

𝐶𝑂𝐸 ∑𝑆𝑠=1 𝐸𝑠 } − 𝐶𝐴𝑃𝐸𝑋 (22)

where, 𝑈𝐶𝑠 (1 ≤ 𝑠 ≤ 𝑆) is the unit utility cost ($/kW-a) for stream 𝑠, 𝑟 is the rate of interest,

and 𝑁 years is the project life of HEN.

This completes our MINLP model for the synthesis of EIP-wide HEN. The model

comprises eqs. 1-22 except eq. 19a-b that are eliminated by substitution into eq. 22. We now

illustrate its application via several examples from the literature.

5. Examples

We take five examples with only one plant for each enterprise, hence plant and enterprise

are interchangeable labels. We assume that for a plant to host an exchanger, the plant or its

enterprise must own at least one of the two exchanging streams. Table 1 lists all the data; Table

2 gives the cost correlations, pressure drops, stream densities, and pipe sizes; and Table 3 lists

the various model parameters such as variable bounds. For examples with no plant-to-plant

distance data, we assume the distances as in Table 1. For all examples, we used SCIP 3.1

(Achterberg, 2009) in GAMS 24.4.6 (2015) as the MINLP solver. GAMS/SCIP uses CPLEX

as the LP solver, COIN-OR Interior Point Optimizer IPOPT (Wächter and Biegler, 2006) as the

nonlinear solver, and CppAD to compute the derivatives of nonlinear functions.

5.1. Example 1

This example is from Nair et al. (2016) who adopted it from Björk and Westerlund (2002). It

has three enterprises or plants: E1, E2, and E3. E1 owns two hot streams (H1 and H2), E2 owns

three cold streams (C1, C2, and C3) and E3 owns one cold stream (C4). We consider three

variations for this example with 𝐾 = 3 and 𝐾 = 4. Case 1 is the original example from Nair et

al. (2016) with no ambient heat effects. Case 2 considers the ambient heat effects. In Case 3,

we move E2 away from E1 and E3 to see the impact on HEN. For all cases except one, we
achieve an optimality gap of 0.1 % for both 𝐾 = 3 and 𝐾 = 4, and compare the results of our

distributed HENs with the centralized HEN reported by Nair et al. (2016) in Table 4. The final

HEN and results are the same for both 𝐾 = 3 and 𝐾 = 4. However, the computation effort is

much higher for the latter than the former as expected.

Case 1

This case ignores the ambient heat gains/losses. Figure 2 shows the best distributed HEN.

Circled numbers denote the plants where streams and exchangers are located in various stages.

Only H1 and H2 flow from E1 to other plants for heat exchange. C1-C4 remain in their parent

plants. H1 bypasses stage 1, moves from E1 to E3 in stage 2, moves to E2 in stage 3, and returns

to E1 to bypass stage 4. H2 moves from E1 to E2 in stage 1, where it splits into two substreams.

It remains at E2 in stage 2, bypasses stage 3, and returns to E1 in stage 4. H1 and C1 do not

need their in-plant heaters/coolers, while the others do.

The distributed HEN has NPV = $2.39 million and IROR = 161.5 %. In contrast, Nair et

al. (2016) reported NPV = $1.84 million and IROR = 78.8 % for their centralized HEN. Thus,

by optimally locating the exchangers, the distributed HEN increases NPV by 30 % and IROR

by nearly 100 %. Only two streams flow out of their plants in the distributed HEN versus six in

the centralized HEN. The lower transport needs reduce CAPEX by 45.9 %, and increase NPV

and IROR. However, the distributed HEN saves less energy (i.e. utility heating and cooling)

than the centralized HEN. As a result, the savings in the utility costs are slightly lower ($1.35

million/a vs $1.39 million/a).

The centralized HEN involved five exchanges: H1-C2, H1-C3, H2-C1, H2-C2, and H2-

C4. In contrast, the distributed HEN involves five exchanges: H1-C1, H1-C4, H2-C1, H2-C2,

and H2-C3. Clearly, the exchanges have changed due to the economics of transporting streams.

The restricted availability of streams at each plant has limited exchanges and reduced energy

savings. In the centralized HEN, H2 from E1 could heat in parallel both C1 (E2) and C4 (E3)
that belong to different enterprises. In the distributed HEN, this is not economical, as both C1

and C4 must travel from E2 and E3 to E1, increasing piping and pumping costs. The distributed

HEN saved 58 % in piping costs, and found alternate exchanges. While this reduced energy

savings slightly, it increased NPV and IROR drastically. This shows the importance of trading

off CAPEX versus OPEX, and true costs rather that simple energy savings to give a better HEN.

Case 2

Now, we examine the impact of ambient heat gains/losses. We assumed 10 W/m2-K

(Engineering Toolbox) as the overall heat transfer coefficient of all pipelines. In this case, all

streams lose heat to the ambient. This represents a loss for both hot and cold streams, as hot

streams lose the heating energy, and cold streams need more cooling energy. These ambient

heat losses reduce the energy savings as compared to Case 1. The utility cost savings reduce

from $1.35 million/a to $1.14 million/a. The best HEN (Figure 3) is also very different from

Case 1. NPV is lower ($2.02 million versus $2.3 million), but IROR remains nearly the same.

In contrast to Case 1, H2 and C3 do not need their utility exchangers, but all other streams need

them. In contrast to Case 1, the heat losses in Case 2 reduced the incentive for heat exchange

between H1 and C1. As a result, the best HEN for this case does not send H1 to E2. This reduced

the pipeline for H1, and overall CAPEX to $0.65 million. Also, the exchange duties for H1-C4

and H2-C2 are also reduced because of the ambient heat losses compared to Case 1. Thus, it is

important to consider the ambient heat gains/losses.

Case 3

In this case, we move E2 away from both E1 and E3, and ignore the ambient heat gains/losses.

Thus, the distances of E2 are now 3 km versus 0.61 km from E1, and 3.06 km versus 0.70 km

from E3. Since E2 owns three of the four cold streams, we expect this change to impact the

distributed HEN. Figure 4 shows the best distributed HEN for this case. Even though E2 is

farther away from E1 and E3, the best HEN still sends H2 to E2, and it does not send H1 to E2
as in Case 1. As a result, the duty of the utility cooler for H1 increases. Since H2 flows to E2,

the heat from H2 is fully utilized for the cold streams of E2, and H2 does not need its own utility

cooler. However, as the energy from H2 is still not sufficient for C1, C2, and C4; they need

their in-plant utility heaters. These differences lower the utility cost savings from $1.35

million/a to $1.27 million/a. Moving H2 to E2 increases piping costs and CAPEX considerably.

CAPEX is 98 % higher than Case 1. This reduces IROR and NPV to 59 % and $1.36 million

respectively.

5.2. Example 2

Consider Example 2b from Nair et al. (2016). It has 3 enterprises with 4 hot and 5 cold streams.

We solve this example for 𝐾 = 3 only, as computational effort is excessive for 𝐾 = 4. Figure

5 shows the best distributed HEN for this example with an optimality gap of 4.8 %. It involves

6 exchanges: H1-C1, H2-C3, H2-C4, H3-C1, H3-C2 and H4-C5. In contrast, the centralized

HEN involved only 3 exchanges; H2-C1, H3-C1, and H3-C2. A major difference between the

two sets of exchanges is the fact that the exchanges in the distributed HEN involve both intra-

plant and inter-plant heat integration. The dotted lines indicate the intra-plant heat integration.

In this example, the availability of both hot and cold streams at each plant allows the possibility

of intra-plant exchanges, which are well exploited by the distributed HEN model to reduce the

piping and pump costs. This illustrates the ability of our model to naturally address both intra-

plant and inter-plant integrations. The centralized HEN of Nair et al. (2016) for this example

did not use H1, H4, C3, C4, and C5 for integration, because transporting them to the central

site was not economical compared to the potential utility savings. In contrast, the distributed

HEN uses all of them for integration. This leads to a greater energy savings for the distributed

HEN and utility cost savings ($0.190 million/a vs $0.158 million/a), The distributed HEN has

49.7 % lower CAPEX, as no streams are transported. IROR increases from 14.2 % to 54.5 %

and NPV increases from $0.09 million to $0.587 million. These results show the significant
advantage of using the distributed versus centralized HEN model. Table 5 compares the full

details of the two HENs.

This example was also addressed using a game-theory based optimization strategy for

inter-plant heat integration by Cheng et al. (2014). However, no meaningful comparison can be

made, as they ignored the piping and pumping costs.

5.3. Example 3

This has four plants, six hot streams, and six cold streams. Again, we could solve this example

for at most 𝐾 = 3. We compare our distributed HEN with the solution reported by Hiete et al.

(2012) using thermal pinch analysis. While we use their plant-to-plant distances and cost

correlations, a fair and direct comparison is not possible due to the following differences:

a) Some streams could use two different utility costs in Hiete et al. (2012), while we fixed

utility cost for each stream. Besides, Hiete et al. (2012) did not justify their utility costs.

b) Their heat exchanger costs differ with exchanger type, while we fixed our heat exchanger

cost at their highest investment price.

c) Their piping costs are functions of volumetric flowrate; but they provided mass flowrates

without giving stream densities. Therefore, we assumed a density of 900 kg/m3 for all streams.

d) In their model, plants hosted the heat exchangers, but it was not clear how they selected

plant locations. Their solutions did not report the order and locations of exchangers.

e) They gave no data on heat transfer coefficients and minimum temperature approach.

Table 6 summarizes the results for our best distributed HEN with an optimality gap of 5.2

%. Figure 6 shows the plants with both hot and cold streams exploited intra-plant heat

exchanges. Only four streams (C3, C4, C5 and C6) were sent away from their owner plants. C3

was transported from E1 to E2 in stage 2, and returned to E1 in stage 0. C4 was sent to E1 in

stage 3 and returned to E2 in stage 2. C5 was sent from E3 to E1 in stage 3, and returned to E3

in stage 0. Finally, C6 was sent from E4 to E1 in stage 3, and returned to E4 in stage 1. The
distributed HEN involves 10 exchanges: H1-C2, H2-C2, H2-C5, H2-C6, H3-C1, H3-C2, H3-

C4, H5-C3, H5-C4, and H6-C6. Of these, H1-C2, H2-C2, H3-C1, H3-C2, H3-C2, H5-C4, and

H6-C6 are intra-plant exchanges. In other words, our proposed model optimizes both intra-plant

and inter-plant exchanges simultaneously. Zero fixed costs of exchangers in this example leads

to many exchangers in HEN. Despite using lower hot (25.7 MW) and cold (27.1 MW) utilities,

the distributed HEN has a higher utility cost ($3.676 million/a vs $3.406 million/a) than Hiete

et al. (2012). This is due to a fixed utility cost for each stream, as Hiete et al. (2012) allowed

two utility costs for some streams. The amortized capital cost ($0.395 million/a) from the

distributed HEN model is significantly lower than Hiete et al. (2012) ($0.773 million/a). The

missing stream densities could cause this difference; resulting in different volumetric flowrates

and stream piping costs. The distributed HEN has NPV = $8.9 million and IROR = 113 %.

5.4. Example 4

Wang et al. (2015) used this example. It has two plants with E1 having four hot streams and E2

having seven cold streams. While they illustrated their model for direct, indirect, and combined

heat integration, we consider only direct heat integration for the sake of a fair comparison.

However, comparing the two solutions (ours and theirs) was not easy because of the following

differences. First, their model applies to two plants only. Second, they ignored the costs of

pumps and pumping. Third, they reported the pipe costs for their HEN, but did not report the

heat exchanger locations. Fourth, they did not report crucial stream properties such as specific

heat capacities and densities for their direct integration example. Last, their model assumed

isothermal mixing. Therefore, we set pump and pumping costs to zero for this example, and

take the stream properties from their example on indirect heat integration.

The computational effort again limited us to 𝐾 = 3 for this exercise. Table 7 compares

our HEN (optimality gap = 3.6 %) with theirs. Our best HEN in Figure 7 shows four streams

being transported. Wang et al. (2015) also had transported four streams. However, we cannot
know which four streams, as heat exchanger locations are unknown. Our HEN has 10

exchanges: H1-C1, H1-C2, H1-C3, H1-C4, H1-C5, H2-C5, H2-C6, H3-C6, H3-C7, and H4-

C7. Of these, H1-C5, H2-C6, and H3-C7 do not exist in Wang et al. (2015). In both HENs, the

cold streams fully meet hot stream duties. Both HENs achieve the same maximum energy

savings and utility costs ($0.228 million/a). However, our HEN is better, as it needs less heat

exchange area and cost (1667 m2 vs. 2069 m2 and $0.280 million/a vs. $0.334 million/a). The

difference is mainly due to the isothermal mixing assumption. Although both HENs have four

transported streams, our pipe costs are lower ($0.371 vs $0.632 million/a) probably due to the

missing stream properties like heat capacities and densities. Overall, our HEN has a lower TAC

($0.878 vs $1.195 million/a).

Example 5

Chang et al. (2017) proposed to use Generalized Disjunctive Programming (GDP) for inter-

plant HENS. In addition to stages, they introduced “intervals” to locate streams at different

plants for heat exchange. They minimized TAC rather than maximize NPV, so we do the same

for a fair comparison. While we used their parameters, costs, and pressure drop correlations,

we encountered certain differences. First, eqs. 56 and 57 for pressure drops in Chang et al.

(2017) are incorrect, so we used the equations from the original source of Soltani and Shafiei

(2011). Second, the exchanger fixed cost is stated as $1,100 instead of $11,000. Third, our

calculations for their network give higher piping costs than those reported by them. Fourth, they

included the costs of in-plant utility exchangers as well, so we also did the same for a valid

comparison. Last, they assumed isothermal mixing, but we do not.

We solved this example for both 𝐾 = 3 and 𝐾 = 4. Figure 8a and Table 8 show our

optimal HEN with a gap of 1 % for 𝐾 = 3. It involves 7 exchanges: H1-C1, H1-C3, H2-C2,

H2-C3, H2-C5, H4-C1, and H4-C4. Among these, H1-C3, H2-C5 and H4-C1 are inter-plant

heat exchanges. In Chang et al. (2017), exchanges H1-C3 and H4-C1 are not present. Also, H1
and C5 are transported in Chang et al. (2017), whereas H1, H2, and H4 in this work. Thus, our

HEN has a higher piping cost ($0.083 million/a Vs $0.054 million/a). However, allowing non-

isothermal mixing gives 41.9 % lower utility costs ($0.96 million/a vs $1.66 million/a) and 33.4

% lower TAC. Thus, our model gives a much better HEN.

For 𝐾 = 4 (see Table 8), we obtain the network in Figure 8b. This has 0.5 % lower TAC

than 𝐾 = 3. However, it takes 200,000 s of CPU time (versus 61,186 s for 𝐾 = 3) to reach a

relative optimality gap of 21.9 % (versus 1 % for 𝐾 = 3). The primary reason for the greater

optimality gap seems to be the significantly poorer relaxation resulting in a much worse lower

bound of $ 1,210,887 versus $ 1,469,050 (for 𝐾 = 3).

6. Conclusion

The costs of piping and pumping streams can be a deal-breaker for an EIP-wide HEN, as they

constitute major capital and operating expenses. We presented an MINLP model for

synthesizing an HEN where the heat exchangers can be located at any plant or central sites.

Thus, we generalized our previous work (Nair et al, 2016) on centralized EIP-wide HENS. In

addition, we included the effects of ambient heat gains/losses, pressure drops during stream

transports between locations, and a seamless treatment of both inter-plant and intra-plant heat

exchanges. Our numerical evaluation on five literature examples shows that our distributed

HEN model correctly trades off between energy and transport costs. The idea of distributed

HEN reduces stream movements and piping/pumping costs (both CAPEX and OPEX)

significantly. For the first two examples that allowed a central site, no exchanger was placed at

that site in the best HENs. Our model gives better results for all the five examples. However,

reaching a low optimality gap was a challenge for most examples. Thus, a method for efficient

model solution remains an area of further interest.

Acknowledgement

Sajitha K Nair acknowledges financial support under the President’s Graduate Fellowship from
the National University of Singapore. The work was also funded in part by the National

University of Singapore through a seed grant (R-261-508-001-646/733) for CENGas (Center

of Excellence for Natural Gas).


Appendix

Nair et al. (2016) derived/used the following for temperature changes and approaches. For

detailed explanation, please refer Nair et al. (2016).

𝑇𝑖𝐿 = Lowest acceptable temperature for substream 𝑖 (1 ≤ 𝑖 ≤ 𝐼)

𝑇𝑗𝑈 = Highest acceptable temperature for substream 𝑗 (1 ≤ 𝑗 ≤ 𝐽)

𝑀𝑇𝐴𝑖𝑗 = Minimum temperature approach for 𝐻𝐸𝑖𝑗𝑘 (1 ≤ 𝑖 ≤ 𝐼; 1 ≤ 𝑗 ≤ 𝐽)

𝐻𝑇𝐴𝑖𝑗𝑘 = Hot end temperature approach (1 ≤ 𝑖 ≤ 𝐼; 1 ≤ 𝑗 ≤ 𝐽; 1 ≤ 𝑘 ≤ 𝐾)

𝐶𝑇𝐴𝑖𝑗𝑘 = Cold end temperature approach (1 ≤ 𝑖 ≤ 𝐼; 1 ≤ 𝑗 ≤ 𝐽; 1 ≤ 𝑘 ≤ 𝐾)

𝑄𝑖𝑘
𝑇𝐸𝑖𝑘 = 𝑇𝑖𝑘 − 1 ≤ 𝑖 ≤ 𝐼; 1 ≤ 𝑘 ≤ 𝐾 (A.1)
𝐹𝑖

𝑄𝑗𝑘
𝑇𝐸𝑗𝑘 = 𝑇𝑗𝑘 + 1 ≤ 𝑗 ≤ 𝐽; 1 ≤ 𝑘 ≤ 𝐾 (A.2)
𝐺𝑗

𝑄𝑖(𝐾+1)
𝑇𝑂𝑈𝑇𝑖 = 𝑇𝑖(𝐾+1) − 1≤𝑖≤𝐼 (A.3)
𝐹𝑖

𝑄𝑗0
𝑇𝑂𝑈𝑇𝑗 = 𝑇𝑗0 + 1≤𝑗≤𝐽 (A.4)
𝐺𝑗

𝐷𝑖𝑗𝑈 = 𝑚𝑖𝑛[𝑇𝐼𝑁𝑖 − 𝑇𝑖𝐿 , 𝑚𝑎𝑥{𝐷𝑖𝐿 , 𝑇𝐼𝑁𝑖 − 𝑇𝐼𝑁𝑗 − 𝑀𝑇𝐴𝑖𝑗 }]

1 ≤ 𝑖 ≤ 𝐼; 1 ≤ 𝑗 ≤ 𝐽 (A.5)

𝑈
𝑅𝑖𝑗 = 𝑚𝑖𝑛[𝑇𝑗𝑈 − 𝑇𝐼𝑁𝑗 , 𝑚𝑎𝑥{𝑅𝑗𝐿 , 𝑇𝐼𝑁𝑖 − 𝑇𝐼𝑁𝑗 − 𝑀𝑇𝐴𝑖𝑗 }]

1 ≤ 𝑖 ≤ 𝐼; 1 ≤ 𝑗 ≤ 𝐽 (A.6)

𝑓𝑖𝑗𝑘 𝐹𝑖 𝐷𝑖𝐿 ≤ 𝑄𝑖𝑗𝑘 ≤ 𝑓𝑖𝑗𝑘 𝐹𝑖 𝐷𝑖𝑗𝑈 1 ≤ 𝑖 ≤ 𝐼; 1 ≤ 𝑗 ≤ 𝐽; 1 ≤ 𝑘 ≤ 𝐾 (A.7)

𝑔𝑖𝑗𝑘 𝐺𝑗 𝑅𝑗𝐿 ≤ 𝑄𝑖𝑗𝑘 ≤ 𝑔𝑖𝑗𝑘 𝐺𝑗 𝑅𝑖𝑗


𝑈
1 ≤ 𝑖 ≤ 𝐼; 1 ≤ 𝑗 ≤ 𝐽; 1 ≤ 𝑘 ≤ 𝐾 (A.8)

𝑇𝑖𝑘 − 𝐷𝑖𝑗𝑘 ≥ 𝑇𝑖𝐿 1 ≤ 𝑖 ≤ 𝐼; 1 ≤ 𝑗 ≤ 𝐽; 1 ≤ 𝑘 ≤ 𝐾 (A.9)

𝑇𝑗𝑘 + 𝑅𝑖𝑗𝑘 ≤ 𝑇𝑗𝑈 1 ≤ 𝑖 ≤ 𝐼; 1 ≤ 𝑗 ≤ 𝐽; 1 ≤ 𝑘 ≤ 𝐾 (A.10)

𝑀𝑇𝐴𝑖𝑗 ≤ 𝐻𝑇𝐴𝑖𝑗𝑘 ≤ 𝑇𝑖𝑘 − 𝑇𝑗𝑘 − 𝑅𝑖𝑗𝑘 + (𝑀𝑇𝐴𝑖𝑗 + |𝑇𝑂𝑈𝑇𝑖 − 𝑇𝑂𝑈𝑇𝑗 |)(1 − 𝑥𝑖𝑗𝑘 )

1 ≤ 𝑖 ≤ 𝐼; 1 ≤ 𝑗 ≤ 𝐽; 1 ≤ 𝑘 ≤ 𝐾 (A.11)

𝑀𝑇𝐴𝑖𝑗 ≤ 𝐶𝑇𝐴𝑖𝑗𝑘 ≤ 𝑇𝑖𝑘 − 𝑇𝑗𝑘 − 𝐷𝑖𝑗𝑘 + (𝑀𝑇𝐴𝑖𝑗 + |𝑇𝑂𝑈𝑇𝑖 − 𝑇𝑂𝑈𝑇𝑗 |)(1 − 𝑥𝑖𝑗𝑘 )
1 ≤ 𝑖 ≤ 𝐼; 1 ≤ 𝑗 ≤ 𝐽; 1 ≤ 𝑘 ≤ 𝐾 (A.12)

𝐻𝑇𝐴𝑖𝑗𝑘 ≥ 𝐶𝑇𝐴𝑖𝑗𝑘 + 𝐿𝑀𝑇𝐷𝑖𝑗𝑘 [𝑙𝑛(𝐻𝑇𝐴𝑖𝑗𝑘 ) − 𝑙𝑛(𝐶𝑇𝐴𝑖𝑗𝑘 )]

1 ≤ 𝑖 ≤ 𝐼; 1 ≤ 𝑗 ≤ 𝐽; 1 ≤ 𝑘 ≤ 𝐾 (A.13)

2𝐿𝑀𝑇𝐷𝑖𝑗𝑘 ≤ 𝐻𝑇𝐴𝑖𝑗𝑘 + 𝐶𝑇𝐴𝑖𝑗𝑘 1 ≤ 𝑖 ≤ 𝐼; 1 ≤ 𝑗 ≤ 𝐽; 1 ≤ 𝑘 ≤ 𝐾 (A.14)


Nomenclature

Subscripts

𝑒 Enterprise in an EIP

𝑖 Hot process stream

𝑗 Cold process stream

𝑝 Plant

𝑘 Stage in the HEN superstructure

𝑠 Process stream

Superscripts

𝐿 Lower limit

𝑈 Upper Limit

Parameters

𝐶𝑂𝐸 Cost of electricity ($/W)

𝐶𝑃𝐿 Installed Pipe cost per unit length ($/m)

𝑑𝑑 Distance between plants (m)

𝐸 Number of enterprises

𝐹 Heat-content flow (mass flow x heat capacity or kW/K) of hot stream

𝐹𝐶𝐻 Fixed component of the purchase cost of a heat exchanger ($)

𝐹𝐶𝑃 Fixed component of the purchase cost of a pump ($)

𝐺 Heat-content flow (mass flow x heat capacity or kW/K) of cold stream

𝐼 Number of hot process streams

𝐽 Number of cold process streams

𝐾 Number of stages in the superstructure excluding the utility exchanger stages

𝑀𝐻 Installed cost-Purchased cost multiplier for an exchanger

𝑀𝑃 Installed cost-Purchased cost multiplier for a pump


𝑀𝑇𝐴 Minimum allowable temperature approach in an exchanger (K)

𝑁 Number of years (y)

𝑃 Total number of plants

𝑟 Interest rate for NPV computation

𝑆 Number of process streams

𝑆𝑆 Number of substreams in a stage

𝑇𝐼𝑁 Initial temperature of stream (K)

𝑇𝑂𝑈𝑇 Final (target) temperature of stream (K)

𝑈 Overall Heat Transfer Coefficient (kW/m2-K)

𝑈𝐶 Unit utility cost ($/kW-a)

𝑉𝐶𝐻 Coefficient in the variable component of the purchase cost of a heat exchanger

𝑉𝐶𝑃 Coefficient in the variable component of the purchase cost of a pump

𝑉𝐹 Volumetric Flow rate (m3/a)

𝛽 Exponent in the cost correlation for an exchanger

∆𝑃𝐻𝐸 Pressure drop per exchanger (Pa)

∆𝑃𝐿 Pressure drop per unit length (Pa/m)

∆𝑇 Average Temperature rise per unit distance (K/m)

η Pump efficiency

𝜌 Density (kg/m3)

Continuous Variables

𝐴 Area of an exchanger (m2)

𝐶𝐴𝑃𝐸𝑋 Total capital expense ($)

𝐶𝑃 Purchase cost of pump ($)

𝐶𝑇𝐴 Temperature approach at the cold end of an exchanger (K)


𝐷 Temperature change for hot substream in an exchanger (K)

𝐸 Pump Power (W)

𝑓 Fractional heat-content flow of hot substream in an exchanger

𝑔 Fractional heat-content flow of cold substream in an exchanger

𝐻𝐸𝐶 Heat exchanger cost ($)

𝐻𝑇𝐴 Temperature approach at the hot end of an exchanger (K)

𝐿𝑃𝐿 Length of pipeline (m)

𝐿𝑀𝑇𝐷 Logarithmic Mean Temperature Difference (K)

𝑁𝑃𝑉 Net Present Value ($)

𝑂𝑃𝐸𝑋 Operating expense ($/a)

𝑄 Heat duty (kW)

𝑅 Temperature change for cold substream in an exchanger (K)

𝑇 Entry Temperature of a stream at a stage (K)

𝑇𝐸 Exit Temperature of a stream at a stage (K)

Binary variables

𝑎 Heat exchanger location

𝑁𝑃 Requirement of new pump

𝑥 Existence of a heat exchanger

𝑦 Entry of stream in a stage

𝑌 Existence of stage

𝑧 Stream location in a stage

𝑍 Stream movement

Acronyms

EIP Eco-Industrial Park

HE Heat Exchanger
HEN Heat Exchanger Network

HTC Heat Transfer Coefficient

IROR Internal rate of return

MINLP Mixed-Integer Non-Linear Programming


References

Achterberg, T. (2009). Scip: Solving Constraint Integer Programs. Mathematical Programming Computation, 1, 1-
41.
Bagajewicz, M., & Rodera, H. (2000). Energy Savings in the Total Site Heat Integration across Many Plants.
Computers & chemical engineering, 24, 1237-1242.
Björk, K.-M., & Westerlund, T. (2002). Global Optimization of Heat Exchanger Network Synthesis Problems with
and without the Isothermal Mixing Assumption. Computers & chemical engineering, 26, 1581-1593.
Chang, C., Chen, X., Wang, Y., & Feng, X. (2017). Simultaneous Synthesis of Multi-Plant Heat Exchanger Networks
Using Process Streams across Plants. Computers and Chemical Engineering, 101, 95-109.
Chen, C.-L., & Lin, C.-Y. (2012). Design of Entire Energy System for Chemical Plants. Industrial & Engineering
Chemistry Research, 51, 9980-9996.
Cheng, S.-L., Chang, C.-T., & Jiang, D. (2014). A Game-Theory Based Optimization Strategy to Configure Inter-
Plant Heat Integration Schemes. Chemical Engineering Science, 118, 60-73.
Gams Development Corporation, General Algebraic Modeling System (Gams) Release 24.4.6. (2015).
Washington, DC, USA, .
Gibbs, D., Deutz, P., & Proctor, A. (2005). Industrial Ecology and Eco‐Industrial Development: A Potential
Paradigm for Local and Regional Development? Regional Studies, 39, 171-183.
Hiete, M., Ludwig, J., & Schultmann, F. (2012). Intercompany Energy Integration. Journal of industrial Ecology,
16, 689-698.
Huang, K. F., Al-mutairi, E. M., & Karimi, I. (2012). Heat Exchanger Network Synthesis Using a Stagewise
Superstructure with Non-Isothermal Mixing. Chemical Engineering Science, 73, 30-43.
Huang, K. F., & Karimi, I. A. (2013). Simultaneous Synthesis Approaches for Cost-Effective Heat Exchanger
Networks. Chemical Engineering Science, 98, 231-245.
Hui, C., & Ahmad, S. (1994). Minimum Cost Heat Recovery between Separate Plant Regions. Computers &
chemical engineering, 18, 711-728.
Laukkanen, T., Tveit, T.-M., & Fogelholm, C.-J. (2012). Simultaneous Heat Exchanger Network Synthesis for Direct
and Indirect Heat Transfer inside and between Processes. Chemical Engineering Research and Design,
90, 1129-1140.
Manahan, S. E. (2004). Environmental Chemistry, Eighth Edition. Crc Press, Boca Raton.
Matsuda, K., Hirochi, Y., Tatsumi, H., & Shire, T. (2009). Applying Heat Integration Total Site Based Pinch
Technology to a Large Industrial Area in Japan to Further Improve Performance of Highly Efficient
Process Plants. Energy, 34, 1687-1692.
Nair, S. K., Guo, Y., Mukherjee, U., Karimi, I., & Elkamel, A. (2016). Shared and Practical Approach to Conserve
Utilities in Eco-Industrial Parks. Computers & chemical engineering, 93, 221-233.
Nehring, R. (2009). Traversing the Mountaintop: World Fossil Fuel Production to 2050. Philosophical
Transactions: Biological Sciences, 364, 3067-3079.
Nemet, A., Čuček, L., & Kravanja, Z. (2016). Procedure for the Simultaneous Synthesis of Heat Exchanger
Networks at Process and Total Site Level Chemical Engineering Transactions, 52, 1057-1062.
Overall Heat Transfer Coefficient. Engineering ToolBox. Accessed on 09/04/ 2017,
http://www.engineeringtoolbox.com/overall-heat-transfer-coefficient-d_434.html.
Soltani, H., & Shafiei, S. (2011). Heat Exchanger Networks Retrofit with Considering Pressure Drop by Coupling
Genetic Algorithm with Lp (Linear Programming) and Ilp (Integer Linear Programming) Methods. Energy,
36, 2381-2391.
Tudor, T., Adam, E., & Bates, M. (2007). Drivers and Limitations for the Successful Development and Functioning
of Eips (Eco-Industrial Parks): A Literature Review. Ecological Economics, 61, 199-207.
Wächter, A., & Biegler, L. T. (2006). On the Implementation of an Interior-Point Filter Line-Search Algorithm for
Large-Scale Nonlinear Programming. Mathematical Programming, 106, 25-57.
Wang, Y., Chang, C., & Feng, X. (2015). A Systematic Framework for Multi-Plants Heat Integration Combining
Direct and Indirect Heat Integration Methods. Energy, 90, 56-67.
Yee, T. F., & Grossmann, I. E. (1990). Simultaneous Optimization Models for Heat Integration—Ii. Heat Exchanger
Network Synthesis. Computers & chemical engineering, 14, 1165-1184.
List of Tables

Table 1 Stream temperatures, heat contents, utility costs, and plant-to-plant distances for

Examples 1-5.

Table 2 Stream densities, pressure drops, pipe/pump sizing data, and cost data for Examples 1-

5.

Table 3 Parameters for Examples 1-5

Table 4 Comparison of centralized versus distributed HENs for Example 1

Table 5 Comparison of centralized versus distributed HENs for Example 2

Table 6 Comparison of our HEN with Hiete et al. (2012) for Example 3

Table 7 Comparison of our HEN with Wang et al. (2015) for Example 4

Table 8 Comparison of our HEN with Chang et al. (2017) for Example 5

List of Figures

Figure 1 Schematic of HEN superstructure

Figure 2 Best HEN solution for Example 1: Case 1

Figure 3 Best HEN solution for Example 1: Case 2

Figure 4 Best HEN solution for Example 1: Case 3

Figure 5 Best HEN solution for Example 2

Figure 6 Best HEN solution for Example 3

Figure 7 Best HEN solution for Example 4

Figure 8a Best HEN solution for Example 5 (𝐾 = 3)

Figure 8b Best HEN solution for Example 5 (𝐾 = 4)


Table 1. Stream temperatures, heat contents, utility costs, and distances for Example 1-5.

𝑻𝑰𝑵𝑻𝑶𝑼𝑻 Duty 𝒌𝑾 𝒌𝑱 𝒌𝒈 Utility Cost Plant/


Stream 𝒉 ( ) 𝑪𝒑 ( ) 𝝆 ( 𝟑)
(K) (K) (kW) 𝟐
𝒎 𝑲 𝒌𝒈𝑲 𝒎 ($/kW-a) Enterprise
Example 1
∆𝑇𝑚𝑖𝑛 = 5 K; 𝑁 = 3 y; 𝑟 = 12 %.
𝑑𝑑 (𝐸1, 𝐸2) = 609 𝑚, 𝑑𝑑 (𝐸1, 𝐸3) = 782 𝑚, 𝑑𝑑(𝐸2, 𝐸3) = 700 𝑚.
H1 453.15 348.15 3150 2.0 2.0 900 20 E1
H2 513.15 333.15 7200 2.0 2.0 900 20 E1
C1 313.15 503.15 3800 1.5 2.0 900 120 E2
C2 393.15 533.15 2100 1.5 2.0 900 120 E2
C3 313.15 403.15 2250 2.0 2.0 900 120 E2
C4 353.15 463.15 2200 2.0 2.0 900 120 E3
HU 598.15 598.15 - 1.0 - - - E1-E3
CU 298.15 313.15 - 2.0 - - - E1-E3
Example 2
∆𝑇𝑚𝑖𝑛 = 5 K; 𝑈 = 1 kW/m2-K; 𝑁 =7 y; 𝑟 = 10 %.
𝑑𝑑 (𝐸1, 𝐸2) = 609 𝑚, 𝑑𝑑 (𝐸1, 𝐸3) = 700 𝑚, 𝑑𝑑(𝐸2, 𝐸3) = 782 𝑚.
H1 423.15 313.15 770 - 2.0 900 10.0 E1
H2 473.15 343.15 715 - 2.0 900 22.5 E2
H3 643.15 423.15 660 - 2.0 900 30.0 E3
H4 473.15 313.15 880 - 2.0 900 30.0 E3
C1 333.15 413.15 720 - 2.0 900 90.0 (HU1) E1
C2 383.15 463.15 640 - 2.0 900 90.0 (HU1) E1
C3 303.15 383.15 280 - 2.0 900 30.0 (HU1) E2
C4 413.15 463.15 375 - 2.0 900 30.0 (HU1) E2
C5 383.15 633.15 1125 - 2.0 900 40.0 (HU2) E3
HU1 473.15 473.15 - - - - - E1-E3
HU2 773.15 773.15 - - - - - E3
CU 298.15 303.15 - - - - - E1-E3
Example 3
∆𝑇𝑚𝑖𝑛 = 5 K; 𝑈 = 1 kW/m2-K; 𝑁 = 10 y; 𝑟 = 10 %.
OPEX = 20% of piping cost for pumping + 5% of exchanger costs for maintenance.
𝑑𝑑 (𝐸1, 𝐸2) = 200 𝑚, 𝑑𝑑 (𝐸1, 𝐸3) = 200 𝑚, 𝑑𝑑 (𝐸1, 𝐸4) = 200 𝑚.
𝑑𝑑 (𝐸2, 𝐸3) = 300 𝑚, 𝑑𝑑 (𝐸2, 𝐸4) = 500 𝑚, 𝑑𝑑(𝐸3, 𝐸4) = 200 𝑚.
H1 354.15 348.15 1590 - 3.60 900 18.57 (CU2) E1
H2 354.15 289.15 16644 - 4.19 900 111.69 (CU1) E1
H3 343.15 289.15 15006 - 4.17 900 111.69 (CU1) E1
H4 323.15 293.15 1397 - 4.19 900 18.57 (CU2) E2
H5 433.15 303.15 7258 - 1.00 900 18.57 (CU2) E2
H6 413.15 353.15 838 - 1.01 900 18.57 (CU2) E4
C1 314.15 338.15 1253 - 4.27 900 37.23 (HU1) E1
C2 333.15 393.15 26537 - 4.19 900 37.23 (HU1) E1
C3 351.15 393.15 6979 - 4.18 900 37.23 (HU1) E1
C4 287.15 393.15 1184 - 1.01 900 74.46 (HU2) E2
C5 287.15 423.15 1215 - 1.01 900 74.46 (HU2) E3
C6 287.15 423.15 4116 - 4.19 900 74.46 (HU2) E4
HU1 453.15 453.15 - - - - - E1
HU2 473.15 473.15 - - - - - E1-E4
CU1 277.65 277.65 - - - - - E1-E4
CU2 287.15 303.15 - - - - - E1-E4
Example 4
∆𝑇𝑚𝑖𝑛 = 8 K; 𝑁 = 3 years; 𝑟 = 20 %; 𝑑𝑑(𝐸1, 𝐸2) = 650 𝑚.
H1 421.15 395.15 6000 0.969 4.2 950 20 E1
H2 403.15 373.15 2300 0.599 4.2 950 20 E1
H3 378.15 348.15 1600 0.621 4.2 950 20 E1
H4 368.15 333.15 1200 0.657 4.2 950 20 E1
C1 385.15 410.15 1500 0.667 4.2 950 190 E2
C2 384.15 409.15 1500 0.665 4.2 950 190 E2
C3 383.15 409.15 1500 0.668 4.2 950 190 E2
C4 382.15 407.15 1500 0.667 4.2 950 190 E2
C5 363.15 393.15 2000 0.772 4.2 950 190 E2
C6 328.15 371.15 2100 0.651 4.2 950 190 E2
C7 323.15 348.15 2200 0.598 4.2 950 190 E2
HU 423.15 423.15 - - - - - E2
CU 303.15 313.15 - - - - - E1
Example 5
∆𝑇𝑚𝑖𝑛 = 10 K; Annualization factor = 0.264.
𝑑𝑑 (𝐸1, 𝐸2) = 250 𝑚, 𝑑𝑑 (𝐸1, 𝐸3) = 250 𝑚, 𝑑𝑑(𝐸2, 𝐸3) = 250 𝑚.
H1 523.15 393.15 39000 1 10 621 10 (CU1) E1
H2 773.15 393.15 95000 1.2 10 802 10 (CU1) E2
H3 398.15 392.15 15000 1.1 50 830 10 (CU1) E3
H4 473.15 303.15 34000 1.3 10 725 10 (CU1) E3
C1 458.15 493.15 17500 1.4 8 640 160 (HU3) E1
C2 412.15 773.15 54150 1.2 3 680 200 (HU4) E2
C3 293.15 523.15 23000 1.1 4 760 160 (HU3) E2
C4 383.15 433.15 12500 1.1 5 780 100 (HU2) E3
C5 468.15 478.15 25000 1.3 50 810 100 (HU2) E3
CU1 288.15 293.15 - 1 - - - E1-E3
HU1 423.15 422.15 - 1.2 - - - E1-E3
HU2 493.15 492.15 - 1.5 - - - E1-E3
HU3 553.15 552.15 - 1.8 - - - E1-E3
HU4 1073.15 1023.15 - 2.5 - - - E1-E3
Table 2. Stream densities, pressure drops, pipe, pump sizing data, and costs data for
Examples 1-5.

Cost of Electricity (𝐶𝑂𝐸) [Example 1, 2] 525 $/kW-a


Cost of Electricity (𝐶𝑂𝐸) [Example 5] 800 $/kW-a
Pump Efficiency (η) 0.60
Pipe Diameter (𝐷 𝑖𝑛 𝑚)a [Examples 1 & 2] (𝑀 0.664𝑀0.51 𝜌−0.36 for 0.025 𝑚 ≤ 𝐷 ≤ 0.2 𝑚
in kg/s) 0.534𝑀0.43 𝜌 −0.30 for 0.25 𝑚 ≤ 𝐷 ≤ 0.6 𝑚
880𝐷0.74 for 0.025 𝑚 ≤ 𝐷 ≤ 0.2 𝑚
Piping Cost (𝐶𝑃𝐿𝑠 in $/m) [Examples 1, 2a]
1900𝐷1.73 for 0.25 𝑚 ≤ 𝐷 ≤ 0.6 𝑚
Piping Cost (𝐶𝑃𝐿𝑠 in $/m) [Example 3b] 0.448 6 𝑉 + 292.82 (𝑉 = Flow in m3/h)
4𝐹
𝐷𝑖𝑛 = √𝜋𝜌𝐶
𝑝𝑣
Piping Cost (𝐶𝑃𝐿𝑠 in $/m) [Example 4c] 𝐷𝑜𝑢𝑡 = 1.052𝐷𝑖𝑛 + 0.005251
𝑘𝑊 𝑚 𝑘𝐽
(𝐹 𝑖𝑛 , 𝑣 = 1.2 , 𝐶𝑝 𝑖𝑛 ) 𝑤𝑡 = 644.3𝐷𝑖𝑛2 + 72.5𝐷𝑖𝑛 + 0.4611
𝐾 𝑠 𝑘𝑔−𝐾
𝐶𝑃𝐿 = 0.82𝑤𝑡 + 185𝐷𝑜𝑢𝑡 0.48 + 6.8
+ 265𝐷𝑜𝑢𝑡
𝐷𝑖𝑛 = 0.363 𝐹0.45 𝐶𝑝−0.45 𝜌−0.32
𝐷𝑜𝑢𝑡 = 1.101𝐷𝑖𝑛 + 0.006349
f
Piping Cost (𝐶𝑃𝐿𝑠 in $/m) [Example 5 ] 𝑤𝑡 = 1330𝐷𝑖𝑛2 + 75.18𝐷𝑖𝑛 + 0.9268
𝐶𝑃𝐿 = 0.82𝑤𝑡 + 185𝐷𝑜𝑢𝑡 0.48 + 6.8
+ 265𝐷𝑜𝑢𝑡
0.9
8000 + 240𝑉 (𝑉 = Flow in 𝐿/𝑠)
Pump Purchase Cost (𝐶𝑃𝑠 $)a [Example 1, 2]
Pipeline Pressure Drop (∆𝑝𝑠𝐿 ) [Example 1, 2] 0.5 kPa/m
4𝐹 𝜌𝑣𝐷𝑖𝑛
𝑣 = 𝜋𝜌𝐶 ; 𝑅𝑒 = ;𝑓 =
𝐷𝑖𝑛 2 µ
Pipeline Pressure Drop (∆𝑝𝑠𝐿 ) [Example 5] 0.046
𝑝
2𝑓𝜌𝑣 2
; ∆𝑝𝐿 =
𝑅𝑒 𝐷𝑖𝑛
Exchanger Pressure Drop (∆𝑝𝑠𝐻𝐸 ) [Example 1,
35 kPa
2]
Exchanger Pressure Drop (∆𝑃𝑠𝐻𝐸 ) [Example 5] Soltani and Shafiei (2011)
Heat Exchanger Purchase Costs (HEC in $) with A = heat transfer area in m2
Example 1d 8000 + 50𝐴0.75
Example 2e 670𝐴0.83
Example 3b 600𝐴
Example 4c ($/a) 3000 + 150𝐴
Example 5f ($) 11000 + 150𝐴
Installed / Purchase Cost Ratio for pumps (𝑀𝑃𝑠 ) 4.0
Installed Cost / Purchase Cost Ratio (𝑀𝐻𝑖𝑗 ) for exchanger
Examples 1a, 2a 3.5
Examples 3, 4, 5 1
a
Towler & Sinnott (2013); Hiete et al. (2012); Wang et al. (2015); d Björk and Westerlund
b c

(2002); e Cheng et al. (2014); f Chang et al. (2017)


Table 3. Parameters for Examples 1-5
Parameter Example 1 Example 2 Example 3 Example 4 Example 5
𝐴𝐿𝑖𝑗 (𝑚2 ) 1 4 5 50 50
𝐷𝑖𝐿 (𝐾 ) 5 0.2(𝑇𝐼𝑁𝑖 1 (H1); 5 (H2- 5 5 (H1, H3-
− 𝑇𝑂𝑈𝑇𝑖 ) H4, H6); 10 H4); 50 (H2)
(H5)
𝑅𝑗𝐿 (𝐾) 5 0.2(𝑇𝑂𝑈𝑇𝑗 2 (C1); 5 (C2); 12 5 (C1, C4-C5)
− 𝑇𝐼𝑁𝑗 ) 4 (C3); 10 50 (C2-C3)
(C4-C6)
𝑇𝑖𝐿 (𝐾) 𝑇𝑂𝑈𝑇𝑖 − 5 𝑇𝑂𝑈𝑇𝑖 𝑇𝑂𝑈𝑇𝑖 − 5 𝑇𝑂𝑈𝑇𝑖 − 5 𝑇𝑂𝑈𝑇𝑖 − 5
𝑇𝑗𝑈 (𝐾) 𝑇𝑂𝑈𝑇𝑗 + 5 𝑇𝑂𝑈𝑇𝑗 𝑇𝑂𝑈𝑇𝑗 + 5 𝑇𝑂𝑈𝑇𝑗 + 5 𝑇𝑂𝑈𝑇𝑗 + 5
𝑆𝑆(𝑖) 2 2 2 (H1-H3, 4 (H1); 2 2 (H1-H2,
H5); 1 (H4, (H2); 1 H4); 1 (H3)
H6) (H3, H4)
𝑆𝑆(𝑗) 2 2 (C1-C2, C5); 2 (C1- C3, C6); 1 1 (C1-C5); 2 (C1-C3,
1 (C3-C4) (C4-C5) 2 (C6-C7) C5); 1 (C4)

Table 4. Comparison of centralized versus distributed HENs for Example 1


Performance Metric Distributed Distributed Distributed Centralized
HEN Case 1 HEN Case 2 HEN Case 3 HEN
IROR 161.5% 161.0% 59.4% 78.8%
NPV ($) 2,390,702 2,022,553 1,359,313 1,837,937
Total Utility Savings ($/a) 1,347,938 1,141,612 1,274,000 1,388,177
Hot utility demand (kW) 722 2,459 1,250 434
Cold utility demand (kW) 722 615 1,250 434
CAPEX ($) 769,535 653,766 1,522,342 1,422,229
OPEX ($/a) 31,172 26,478 73,310 29,774
Relative Gap (%) 0.1 0.1 0.1 -
CPU time (s) (𝑲 = 𝟑) 302 303 19,399 -
CPU time (s) (𝑲 = 𝟒) 376 682 100,000 -

Table 5. Comparison of centralized versus distributed HENs for Example 2


Performance Metric Distributed HEN Centralized HEN
IROR 54.5 14.2%
NPV ($) 587,173 93,270
Total Utility Savings ($/a) 190,282 157,950
Hot utility demand (kW) 1,005 1,780
Cold utility demand (kW) 890 1,665
CAPEX ($) 328,594 652,758
OPEX ($/a) 4,270 4,711
Relative Gap (%) 4.8 -

Table 6. Comparison of our HEN with Hiete et al. (2012) for Example 3
Performance Metric Distributed HEN Hiete et al. (2012)
IROR 112.9 % -
NPV ($) 8,897,517 -
Total Utility Savings ($/a) 1,844,015 -
CAPEX ($) 1,497,138 -
OPEX ($/a) 152,342 -
Utility cost ($/a) 3,676,342 3,406,000
Hot utility demand (kW) 25,678 31,319
Cold utility demand (kW) 27,125 32,766
Amortized capital and operating cost ($/a) 395,993 773,000
Total annual cost ($/a) 4,072,334 4,180,000
Relative Gap (%) 5.2 -

Table 7. Comparison of our HEN with Wang et al. (2015) for Example 4
Performance Metric Distributed HEN Wang et al. (2015)
Hot utility demand (kW) 1,200 1,200
Cold utility demand (kW) 0 0
Energy savings (kW) 22,200 22,200
Additional heat exchangers 10 8
Heat exchanger area (m2) 1,667 2,069
Heat exchanger cost ($/a) 280,005 334, 465
Pipeline cost ($/a) 370,649 632,220
Cost of utilities ($/a) 228,000 228,000
Total annual Cost ($/a) 878,653 1,194,686
Relative Gap (%) 3.6 -

Table 8. Comparison of our HEN with Chang et al. (2017) for Example 5
Performance Metric Distributed Distributed Chang et al.
HEN (𝑲 = 𝟑) HEN (𝑲 = 𝟒) (2017)
Hot utility demand (kW) 2,766 2,760 6,250
Cold utility demand (kW) 53,616 53,610 57,100
Total number of heat exchangers 15 17 14
Total number of inter-plant
3 4 4
matches
Total annualized heat exchanger
377,304 327,162 460,404
cost ($/a)
Pipeline cost ($/a) 83,032 76,660 54,238
Pumping cost ($/a) 60,676 109,803 51,789
Utilities cost ($/a) 962,729 962,137 1,661,000
Total annual Cost ($/a) 1,483,741 1,475,762 2,227,431
Relative Gap (%) 1 21.9 -
CPU time (s) 61,186 200,001
Figure 1. Schematic of HEN superstructure

Figure 2. Best HEN solution for Example 1: Case 1

Figure 3. Best HEN solution for Example 1: Case 2


Figure 4. Best HEN solution for Example 1: Case 3

Figure 5. Best HEN solution for Example 2


Figure 6. Best HEN solution for Example 3

Figure 7. Best HEN solution for Example 4


Figure 8a. Best HEN solution for Example 5 (𝐾 = 3)

Figure 8b. Best HEN solution for Example 5 (𝐾 = 4)

You might also like