You are on page 1of 15

Case Studies in Thermal Engineering 43 (2023) 102820

Contents lists available at ScienceDirect

Case Studies in Thermal Engineering


journal homepage: www.elsevier.com/locate/csite

A novel method of variable flow operation in district cooling


system: A case study
Wei Zhang, Xu Jin, Ling Zhang, Wenpeng Hong *
School of Energy and Power Engineering, Northeast Electric Power University, Jilin, 132012, People’s Republic of China

G R A P H I C A L A B S T R A C T

A R T I C L E I N F O A B S T R A C T

Keywords: The district cooling system (DCS) is generally considered to be a key efficient cooling technology to
Multi-cold source cooling system cope with the increasing cooling demand. However, the energy consumption of pumps in DCS ac­
TRNSYS counts for a large proportion. This work demonstrated a novel variable flow control strategy based on
Variable flow operation strategy measured operating parameters. An existing DCS in a subtropical climate zone was used as the
Cooling load baseline model, denoted as Case A. Furthermore, using TRNSYS software, the variable flow control of
the chilled water pumps and the cooling water pumps in Case A was performed in turn, and two
variable flow models were established, denoted as Case B and Case C, respectively. The operation
strategies capable of dealing with the diversified cooling loads were developed and simulated. The
results indicated that compared with Case A, the newly proposed models Case B and Case C reduced
the energy consumption by 9.50% and 14.15%, cut down CO2 emissions by 9.45% and 14.18%, saved
costs by 10.34% and 15.16%, respectively. Moreover, the efficiency of the two models was also
improved by 10.70% and 16.67%, respectively, compared with the Case A. Effective operation
strategies can further improve the performance of DCS, thereby providing energy-saving and efficient
cooling solutions.

* Corresponding author.
E-mail address: hwp@neepu.edu.cn (W. Hong).

https://doi.org/10.1016/j.csite.2023.102820
Received 9 November 2022; Received in revised form 20 January 2023; Accepted 13 February 2023
Available online 15 February 2023
2214-157X/© 2023 The Authors. Published by Elsevier Ltd. This is an open access article under the CC BY-NC-ND license
(http://creativecommons.org/licenses/by-nc-nd/4.0/).
W. Zhang et al. Case Studies in Thermal Engineering 43 (2023) 102820

1. Introduction
Energy is the lifeblood and foundation of human development. With the continuous progress of society, the global energy demand
is increasing daily [1]. Buildings are the primary contributor to the total energy consumption [2], accounting for more than 20% in
China [3], and are also a substantial source of CO2 emissions [4]. With global warming and rising temperatures, cooling has become
increasingly indispensable and vital [5]. The energy consumption of air conditioning systems is almost half of the total consumption of
the building [6]. It is crucial for the building sector to conserve energy and reduce emissions while meeting the demand for cooling by
the residents to advance the “dual-carbon” goals announced by the Chinese government [7]. Therefore, it is vital to provide
low-energy, low-carbon, cost-effective, green and clean cooling options. Among these, the district cooling system (DCS) [8,9] over­
comes the shortcomings of traditional cooling systems, also has the advantages of high efficiency and cleanliness. DCS technology is
that one or more cooling stations provide cooling energy for buildings in a district through the water supply and return pipe network
[10]. The typical DCS with various cold sources include not only the electric cooling technology, but also combined with the waste heat
technology [11] and the cold storage technology that utilizes the difference in peak and valley electricity tariffs [12]. The specific cold
source should be determined by the locally available resources. Overall, DCS transforms traditional cooling systems into sustainable
energy systems [13], resulting in emerging cooling systems being a subject of considerable interest. To explore the performance of DCS,
researchers have done extensive research.
Thakar et al. [14] used a campus cooling system in a hot and dry climate zone. Aiming at diversifying the cooling load, they
proposed a detailed design of a DCS with thermal energy storage and conducted a comprehensive lifecycle cost analysis. The results
demonstrated that the new design can save 0.34 million USD and reduce CO2 emissions by 50%. In Hong Kong, a distributed energy
system (DES) integrating DCS was applied in a typical university campus by Kang et al. [15]. Subsequently, they explored the energy
saving and economic performance of the new integrated cooling system. The results illustrated that the integrated system could save
energy and costs by 9.6% and 44%, respectively. In response to a series of challenges, such as high power consumption caused by the
standalone compression chiller-based cooling method, Jannatabadi et al. [16] proposed to implement a DCS in northern Iran and
performed a performance analysis. The application of the new cooling solution could reduce the current high power consumption, and
the payback period could be shortened to 6 years. Alajmi et al. [17] used simulation methods to study a low-rise building DCS in
Kuwait and analyzed its lifecycle cost. The results indicated that the peak electricity demand and CO2 emissions were cut by 50% and
53%, respectively. A DCS with different cooling sources in a subtropical climate zone was selected as the case study by Zhang et al.
[18]. Furthermore, they explored the impact of ice thermal storage (ITS) on the entire cooling system. The results revealed that the
addition of ITS can provide 20% more cooling capacity. Although the integrated system increased energy consumption and carbon
emissions by a small percentage, but the system was more economical with an annual cost saving of 6.78%.
The above-mentioned previous studies have shown that DCS can bring considerable energy-saving performance and economic
benefits. However, the energy consumption of the pumps accounts for about 20%–30% of the entire cooling system [19], implying that
there is a large space for energy saving. Therefore, how to cut the energy consumption of the pumps has become a critical link to further
energy-saving and high-performance operation of this emerging cooling system [20]. This work aims to explore the energy-saving
performance of a DCS with variable flow pumps (VFPs). Through literature research, it was found that the current research on the
application of VFPs in DCS was insufficient; however, it had received extensive attention in the field of district heating (DH) [21,22]. In
North China, it was estimated that the heating and domestic hot water demand of more than 10 million m2 of building areas was
provided by DH systems with VFPs [23,24]. Sheng et al. [25] indicated that DH with VFPs can reduce electricity consumption by
49.41% per year compared with the previous constant flow system. Gu et al. [26] explored the energy saving potential of VFPs applied
in DH with a case in Shenyang, China. They found that the system using VFPs can save 28.52% of electricity per year compared to the
traditional system. To explore the hydraulic performance of the variable-flow DH, Yan et al. [27] implemented such a system in Kuerle,
China. Their results indicated that the new system saved more than 30% of the energy compared with the constant-flow system.
Simultaneously, the capacities of the pumps were reduced accordingly. Wang et al. [28] developed a novel method for application to
an actual DH with distributed VFPs and examined two scenarios to explore the performance. The results demonstrated that the
maximum energy savings of the two scenarios were 66.8% and 90.3%, respectively, compared with the constant flow system.
From the above studies, it can be concluded that the application of VFPs in DH systems can greatly promote energy conservation
and emission reductions. Integrating the VFPs with the DCS sounds to be also a potential energy efficient cooling solution. However,
after literature research we found that there were few reports about such a new cooling option. To address this gap, this work develops
appropriate control methods to explore the rationality of variable flow cooling based on the measured data. An existing DCS operating
at constant flow, located in Zhuhai, China, was selected as the case study. Then, based on this constant flow model, the TRNSYS
software [29] was used to perform variable flow control program and established the models, and set reasonable operation strategies to
deal with the diversified cooling loads. Additionally, the comprehensive evaluation of the three models was carried out. Finally, the
feasibility of variable flow DCS was explored.
The remainder of this study proceeds as follows. Section 2 explains the key variable flow methods used in this study, and forecasts
the user cooling load. In Section 3, an existing typical DCS in a subtropical climate zone is used as the case, and the detailed system
information is categorized. Moreover, the variable flow control logics of different pumps and the elaborate evaluation methods are
shown. Additionally, the constant flow and variable flow models are built. Section 4 presents the multi-objective discussion based on
the simulation results. The study is concluded in Section 5.

2
W. Zhang et al. Case Studies in Thermal Engineering 43 (2023) 102820

2. Methodology
The primary study object of this work is a multi-cold source DCS, located in Zhuhai, China, which has a subtropical climate. The
peak cooling load is 75,000 kW. To minimizing the energy consumption of the cooling system while meeting user cooling loads, this
study proposes the variable flow operation strategy.

2.1. Variable flow control method


This study adjusts the working speed and flow rate of the pumps according to the real-time load change to realize the set tem­
perature of the chilled water and save the energy to the utmost extent under the premise of meeting the cooling supply. According to
the similarity law of the pump [30], the flow rate, head, and efficiency of the pump conform to the following relationship:


⎪ m(ω) ω

⎪ =

⎪ m(ω0 ) ω0



⎨ H(ω) ( )2
ω
= (1)
⎪ H(ω0 )


ω0



⎪ ηp (ω)

⎪ =1

ηp (ω0 )

where m, H, and ηp represent the flow rate (in m3/h), head (in m), and efficiency of the pump, respectively. ω and ω0 are the actual
speed (in r/min) and rated speed (in r/min) of the pump, respectively, and the ratio of the two is the speed ratio of the pump, rep­
resented by I, calculated as:
ω
I= (2)
ω0
The relationship between the flow rate and head at the rated speed is as follows:

H(ω0 ) = a0 + a1 m(ω0 ) + a2 m(ω0 )2 (3)

where a0 , a1 , and a2 are polynomial fitting coefficients.


Then, combined with Eq. (1) and Eq. (2), Eq. (3) can be derived in terms of actual speed, resulting in Eq. (4):

H(ω) = a0 I 2 + a1 m(ω)I + a2 m(ω)2 (4)


Based on the above deduction process, the flow rate, head, and speed ratio of the pump can be calculated as follows:

H = a0 I 2 + a1 mI + a2 m2 (5)
The relationship between the power supply frequency of the pump and speed ratio can be expressed as follows:
f = 50I (6)

where f represents the power supply frequency of the inverter, in Hz.


Generally, in China, the commonly used frequency range of the VFP is 25 Hz–50 Hz. Therefore, when the simulated speed ratio is
less than 0.5, the VFP will run at 25 Hz, and when the speed ratio is greater than 1, it will run at 50 Hz. When it is between 0.5 and 1, it
will run at the frequency corresponding to the current real speed ratio. Additionally, we define the flow rate at which the design

Fig. 1. Hourly outdoor meteorological data in Zhuhai throughout the year.

3
W. Zhang et al. Case Studies in Thermal Engineering 43 (2023) 102820

temperature difference is reached as the expected flow rate, which is expressed as follows:
Qload
m0 = (7)
cΔt

where m0 is the expected flow rate of the pump, in m3/h, Qload is the current load of the users, in kW, c is the specific heat capacity of the
fluid, in kJ/(kg⋅◦ C), and Δt is the temperature difference between the supply and return water, in ◦ C.

2.2. Cooling load forecast


Accurate cooling load prediction is a key prerequisite for system modeling and simulation. Weather data is an indispensable
condition for forecasting. The annual outdoor meteorological data in Zhuhai is shown in Fig. 1, which reveals that the hourly tem­
perature fluctuates greatly throughout the year, and the minimum and maximum values of dry bulb temperature appear around the
287th and 4552nd hours, and the values are 5.83 ◦ C and 34.39 ◦ C, respectively.
This work selects TRNSYS for cooling load prediction. Based on the characteristics of user buildings (public buildings and resi­
dential buildings) obtained from the investigation, combined with the design standards [31,32], the design parameters of the buildings
are summarized in Table 1. Additionally, the TRNBuild system is used to set up and simulate hourly cooling loads, as shown in Fig. 2.
The region where the DCS is located has a hot summer and warm winter climate, and the cooling load is dominant, and the heating load
can almost be neglected [33].
Fig. 2 plots the load pattern in different periods, in which Fig. 2(a) reveals the hourly distribution for 8760 hours, indicating that the
loads in the initial and end periods are much smaller than those in the intermediate cooling period. To ensure the economic operation
of the system, we choose a period when the cooling load of the system is relatively large to eliminate the factors that cause an un­
economic operation of the system, namely 1752 h–7656 h, which is equivalent to March 15 to November 15 in the month, and plotted
Fig. 2(b).
In order to compare the simulation results with the actual values in a more detailed and comprehensive method to verify the
reliability of the simulation cooling load, we further processed the hourly cooling load data and selected the cooling load data of 24
hours in a typical day with peak cooling load for comparison, and drew the comparison curve as shown in Fig. 3.
It can be seen from Fig. 3 that the user’s peak cooling load occurs at 4090 h, and the simulation value (Qs) and actual value (Qa) are
74,800 kW and 75,000 kW respectively, with an error of 0.27%. Furthermore, in addition to the support of the above peak comparison,
it can be seen from the figure that the trend of the two curves is basically consistent, and the fluctuation of the simulated value and the
actual value at each time is very small. This trend maintains the whole typical day. The above comparison illustrates that the simulated
cooling load data has a certain reliability and provides support and foundation for subsequent simulations.

3. Case study on multi-cold source DCS


This section presents detailed equipment information for the studied case and the evaluation method. To explore a more efficient
and energy-saving cooling solution, the control method of variable flow operation is expounded, and the establishment process of
several comparative models is shown.

3.1. Equipment information collection


After the field investigation, it was concluded that there are three main cold sources in the DCS: electric chiller/chiller(e) driven by
electricity, absorption chiller/chiller (ab) driven by waste heat steam, and ITS utilizing the advantage of time-of-use (TOU) electricity
tariffs. The main equipment of the cooling system also includes dual-mode chiller/chiller(d), cooling tower/c(tower), chilled water
pump/CHWP, cooling water pump/CLWP, plate heat exchanger for cooling (PHEcooling) and for melting (PHEmelting). The connection of
the chillers and auxiliary equipment and the layout of the water supply and return pipelines are illustrated in Fig. 4 based on the actual

Table 1
The design parameters of the buildings.

Items Values

Public building people convection heat dissipation 58 W


people humidity 0.02 kg/h⋅m2
lighting heat dissipation 9 W/m2
electrical heat dissipation 15 W/m2
density of people 0.10 people/m2

Residential building people convection heat dissipation 65 W


people humidity 1.12 kg/h⋅m2
lighting heat dissipation 5 W/m2
electrical heat dissipation 3.8 W/m2
density of people 0.04 people/m2

Lighting convection ratio 60%


radiation ratio 40%
Electrical convection ratio 80%
radiation ratio 20%

4
W. Zhang et al. Case Studies in Thermal Engineering 43 (2023) 102820

Fig. 2. Cooling load pattern of users.

Fig. 3. Comparison curve of hourly cooling load in typical day.

data collected onsite.


Specifically, chiller(d) − 1 is different from the other chillers(d) because it has dual evaporators. Therefore, as shown in the figure, it
can be connected to the return water and ice storage tank for cooling or ice storage cycles simultaneously. While the chillers(d) ∼ 2− 4
need to switch between these two working conditions. Furthermore, chillers(ab) − 1, 2 and chillers(e) − 1, 2 are connected in series, and
the return water at 12 ◦ C is cooled by the chillers(ab) to 7.5 ◦ C, then further cooled by the chillers(e) to 3 ◦ C, and finally returned to the
users. The other 7.5 ◦ C chilled water obtained by chillers(ab) ∼ 3 − 5 exchanges heat with the 1.5 ◦ C glycol from chillers(d) ∼ 2− 4 in
the PHEcooling , after which the temperature of the glycol is increased and then returned to the chillers(d) ∼ 2 − 4. During ice storage, the
− 5.6 ◦ C glycol flowing out of chillers(d) enters the ice coil to participate in cold storage, and then rises to − 1.71 ◦ C and returns to
chillers(d). The 12 ◦ C return water can exchange heat with the 1.5 ◦ C chilled water flowing out of the ice storage tank in the PHEmelting
and return the 3 ◦ C cold water to the collector. The return water is warmed to 6.5 ◦ C and enters the tank for the next ice melting cycle.
Moreover, the cooling water cycle occurs between the chillers, CLWPs, and c(tower)s, and the inlet and outlet temperatures are 32 ◦ C
and 37 ◦ C, respectively. The detailed equipment information of the cooling system is listed in Table 2.

3.2. Performance evaluation


Based on the detailed information of the DCS mentioned above, combined with the actual situation of the energy-consuming
equipment in the system, this section presents methods to evaluate the performance from different perspectives.

3.2.1. Energy consumption


In this study, this index is composed of the energy consumption of the chillers, pumps, and c (tower)s. Among them, the chillers

5
W. Zhang et al. Case Studies in Thermal Engineering 43 (2023) 102820

Fig. 4. Distribution diagram of equipment in the cold source side of DCS.

Table 2
The main basic information of the multi-cold source DCS.

Items Equipment Values

CAP (kW) chillers(e) chiller(e) − 1 3165


chiller(e) − 2 7034
chillers(ab) chiller(ab) − 1 3165
chiller(ab) − 2 7034
chiller(ab) − 3 5627
chiller(ab) − 4, 5 8441
chillers(d) chiller(d) − 1, 2, 4 8441
chiller(d) − 3 5627
PHEs PHEmelting − 1, 2, 3 7034
PHEcooling − 1 5627
PHEcooling − 2, 3 8441
COP chillers(e) – 5.60
chillers(ab) – 0.75
chillers(d) – 5.60

Head (m) CHWPs – 20


CLWPs – 25

Quantity (units) CHWPs – 9


CLWPs – 11
c(tower)s – 36

Temperature (◦ C) return water – 12


chilled water – 3
ice tank outlet – 1.5
cooling water inlet/outlet – 32/37

efficiency pumps – 0.70

6
W. Zhang et al. Case Studies in Thermal Engineering 43 (2023) 102820

include chillers(e), chillers(ab), and chillers(d), and the pumps include CHWPs and CLWPs.
The calculation process is presented in Eqs. 8–11:
Powerchillers = Powerchillers(e) + Powerchillers(d) + Powerchillers(ab) (8)

Powerpumps = PowerCHWPs + PowerCLWPs (9)

Powertotal = Powerchillers + Powerpumps + Powerc(tower)s (10)


∫ t2
ECtotal = Powertotal dt (11)
t1

To discuss the energy consumption of the cooling system, the power of each energy-consuming equipment in the system should be
counted first. In Eqs. 8–10, the left side of the equation is the total power of the chillers, pumps and system, and the right side is the
single item, where Powerchillers(e) , Powerchillers(d) , Powerchillers(ab) , PowerCHWPs , PowerCLWPs and Powerc(tower)s are the powers (in kW) of
chillers(e), chillers(d), chillers(ab), CHWPs, CLWPs, and c(tower)s respectively. After the simulation by TRNSYS, the power of the above
equipment is calculated, and then the total power of the equipment is integrated during the cooling period by Eq. (11) to obtain the
system consumption, namely ECtotal (in kWh), Where t1 and t2 are the two end-point time values of the system cooling period, in h.

3.2.2. CO2 emissions


At present, the world is committed to carbon reduction, so CO2 emissions are an important indicator to evaluate the environmental
friendliness and performance of the system. In this study, this indicator is derived from the system’s cumulative energy consumption
and carbon emissions factor (factorCO2 ). According to the geographical location of the system, the factorCO2 is selected based on carbon
emission standard [34], and the value is 0.5271 kgCO2/kWh. The calculation process is as shown in Eq. (12).
ECO2 = ECtotal · factorCO2 (12)

where ECO2 represent the CO2 emissions, in kg.

3.2.3. Economy and efficiency


Economy is a key factor that determines whether the cooling scheme is adopted and is often the greatest concern for operators. In
this regard, it is necessary to evaluate the economy of the DCS. The system studied in this work contains ITS, which can take advantage
of the difference in peak-to-valley electricity price to store cold to maximize the economy of the system and ease the grid supply
pressure. The operating costs (OCs) are calculated using Eq. (13).
OCs = ECtotal · PTOU (13)

where PTOU is the electricity price corresponding to different periods. The product of the two is the OCs, in million RMB.
To more prominently compare the economics of the different systems, we defined the index of OC reduction ratio (OCRR), that is,
the difference between OC of the constant flow system (OCcon) and OC of the variable flow system (OCvar) divided by the OC of the
constant flow system. As demonstrated in Eq. (14), a positive value of the ratio indicates that the variable flow system is more
economical, and the larger the ratio, the better the economy of the system. Correspondingly, the OCRR of models Case B and Case C are
shown in Eqs. (15) and (16).
OCcon − OCvar
OCRR = × 100% (14)
OCcon

OCCaseA − OCCaseB
OCRRCaseB = × 100% (15)
OCCaseA

OCCaseA − OCCaseC
OCRRCaseC = × 100% (16)
OCCaseA
The well-known advantage of a DCS is its high efficiency. Therefore, system efficiency (η) is also a key item to evaluate the DCS
performance, as shown in Eq. (17), which is determined by the total cooling capacity (Qcooling,system , in kWh) and energy consumption of
the system. Therefore, system efficiency and energy consumption are inseparable.
/
η = Qcooling,system ECtotal (17)

3.3. System controlling


The constant flow operation system Case A shown in Fig. 4 is used as the reference model, and based on this, the variable flow
control method is set for different pumps to realize the variable flow operation.

3.3.1. Variable flow control of the CHWPs


Based on the constant flow model Case A, the CHWPs are firstly controlled by variable flow, and the new model is represented by
Case B. According to the actual arrangement, function and CAP of each chiller in the cooling system, the start-up control signals for the

7
W. Zhang et al. Case Studies in Thermal Engineering 43 (2023) 102820

chillers in the model Case B are listed in Table 3.


Table 3 lists the information of the chillers corresponding to the constant flow signal/ Signal(con) and the variable flow signal/
Signal(var) and other equipment controlled by the corresponding signals. The control process for these signals is set according to the
actual connection layout and the cooling capacity. The start-up program is written based on the logic of cooling capacity accumulation.
When the total cooling capacity of melting ice and running chillers cannot meet the current demand, a new chiller needs to be turned
on. The starting programs of Signals(con) SN (N = 1) and SN (2 ≤ N ≤ 11) need to satisfy the following Eqs. (18) and (19), respectively.
Qload > 3 · QPHEmelting · Si 3 · 0.90 (N = 1) (18)


N
Qload > 0.80 · QSn− 1 + 3 · QPHEmelting · Si 3 · 0.90 (2 ≤ N ≤ 11) (19)
n=2

In the above equations, Qload (in kW) is the cooling capacity that the system needs to provide. QSn (in kW) refers to the cooling
capacity of the chiller currently associated with the signal Sn . QPHEmelting (in kW) means the rated heat exchange of the PHEmelting. Si 3 is
the signal to start the third PHEmelting, because when it is opening, it can be enough to prove that all the PHEmelting are running. N
represents the number of the control signal, and the figure ‘3’ in Eqs. (18) and (19) represents three PHEmelting. Additionally, referring
to the actual work of the chillers and the PHEs, combined with the compilation logic of cooling capacity superposition for the TRNSYS
model [35], the operating upper limit ratios of these two types of equipment are set to 0.80 and 0.90, respectively, as shown in Eq. (19).
The above settings can ensure that when the load rate is close to the working upper limit of the equipment, new equipment will be
started in sequence, so that the system can always run safely.
In order to maximize the economic benefits of the cold storage system, the three PHEmelting are first activated to release the cold
energy as much as possible. If the cooling capacity is not sufficient to supply current demand, the two chillers chiller(e) − 1 and
chiller(ab) − 1 arranged in series with smaller CAP are started first to cope with the smaller cooling load demand. As the cooling load
demand continues to increase, the other two series-connected chillers with slightly larger CAP controlled by signals S3 and S4 are
turned on in sequence. As the cooling continues, the chillers controlled by the signals S5 –S10 are operated according to the sequence in
Table 3, and the reasons are as follows. As presented in Fig. 4, the chilled water at 7.5 ◦ C produced by chiller(ab) − 3/4/5 will exchange
heat with glycol (1.5 ◦ C) from chiller(d) − 2/3/4. Therefore, before starting the chillers(ab) controlled by the signals S8 –S10 , the
chiller(d) − 2/3/4 should be started first in the descending order of CAP, so that the cooling capacity can be provided by as few chillers
as possible. Furthermore, considering that chiller(d) − 1 is a dual-evaporator-type chiller, it is started at the end to facilitate rapid
unloading during ice storage, so that it can be directly used for cooling or ice storage.
The detailed control principle for the Signals(var) is explained as follows. In the constant flow model illustrated in Fig. 4, the chillers
chiller(e) − 1 and chiller(ab) − 1 controlled by S1 and S2 are connected in series, and the pump CHWP(a) − 1 conveys the return water
to the two chillers. Therefore, in the variable flow system, signal S1− controls the pump CHWP(a) − 1 and the chillers chiller(e)− 1

2
and chiller(ab) − 1 simultaneously. The same rule applies to signal S3− 4 . Notably, the control of signals S5 –S7 are different. Taking S5
′ ′ ′ ′

as an example, its corresponding Signal(con) is S5 , and the controlled chiller is chiller(d) − 2. This chiller is connected in series with the
chiller(ab) − 5, so S5 controls the CHWPs in front of the two chillers at the same time, namely CHWP(d) − 2 and CHWP(a) − 5.

Similarly, signals S6 and S7 have the same control idea. Furthermore, the S8 , S9 , and S10 control the chillers chiller(ab) − 5, 4, 3, so the
′ ′

new Signals(var) also control the CHWPs connected to these chillers(ab). Next, we take the signals S5 and S8 as examples to illustrate
′ ′

the control process of the two signals on the same pump as follows.
Fig. 5 presents the variable flow calculation process of CHWPs. As mentioned above, CHWP(a) − 5 is controlled by signals S5 and

S8 at the same time, and the process can be integrated according to the process in Fig. 5. Where QSn ′ (in kW) in the first step represent

the load that the chiller controlled by signal Sn should bear. Si is the signal that controls the start of melting ice. In the second step, the

expected flow rate of CHWP mExp CHWP (in m3/h) is the theoretical flow value required to achieve the design temperature difference.
The cCHWP (in kJ/(kg⋅◦ C)) and Δt (in ◦ C) represent the specific heat capacity and design temperature difference of the fluid in the pump,
respectively. The ratio of mExp CHWP to the real flow rate (mCHWP ) when the pump is running is the speed ratio ICHWP . This value is used

Table 3
Control signals and equipment information of the Case B model.

Signal(con) Signal(var) Chiller CAP (kW) Equipment controlled by signals

S1 S1− 2

chiller(e) − 1 3165 chiller(e) − 1, chiller(ab) − 1, CHWP(a) − 1
S2 chiller(ab) − 1 3165
S3 S3− 4

chiller(e) − 2 7034 chiller(e) − 2, chiller(ab) − 2, CHWP(a) − 2
S4 chiller(ab) − 2 7034
S5 S5

chiller(d) − 2 8441 chiller(d) − 2, CHWP(d) − 2, CHWP(a) − 5
S6 S6

chiller(d) − 4 8441 chiller(d) − 4, CHWP(d) − 4, CHWP(a) − 4
S7 S7

chiller(d) − 3 5627 chiller(d) − 3, CHWP(d) − 3, CHWP(a) − 3
S8 S8

chiller(ab) − 5 8441 chiller(ab) − 5, CHWP(a) − 5
S9 S9

chiller(ab) − 4 8441 chiller(ab) − 4, CHWP(a) − 4
S10 S10

chiller(ab) − 3 5627 chiller(ab) − 3, CHWP(a) − 3
S11 S11

chiller(d) − 1 8441 chiller(d) − 1, CHWP(d) − 1

8
W. Zhang et al. Case Studies in Thermal Engineering 43 (2023) 102820

to calculate the speed ratio of pump under the control of signal Sn , as shown in step 4, where ICon CHWP is the control speed ratio of the

pump. gt and le represent “greater than” and “less than or equal to”, respectively. The logic for controlling the speed ratio is that under
the premise that the signal Sn is greater than zero, when the operating speed ratio of pump is greater than or equal to 1, the value is 1;

when the speed ratio is less than or equal to 0.5, the value is 0.5; when the speed ratio is between 0.5 and 1, the input value is the
current true speed ratio. In step 5, ITru CHWP indicates the true speed ratio of pump in variable flow system. When ice melting occurs, the
pump runs at full flow, which is controlled by the Si . When it is controlled by the signal Sn , the pump speed ratio obeys the calculation

process of ICon CHWP . According to the above calculation process, the true speed ratio of CHWP(d) − 2 and CHWP(a) − 5 controlled by
signals S5 and S8 can be obtained as follows:
′ ′

(20)
′ ′
ITru CHWP(d)− 2 = Si · (1 − S5 ) + ICon CHWP(d)− 2 · S5

(21)
′ ′ ′
ITru CHWP(a)− 5 = ICon S5

CHWP(a)− 5 · S5 · (1 − S8 ) + ICon S8

CHWP(a)− 5 · S8

The integrated speed ratio is the true speed ratio (ITru CHWP(a)− 5 ) when CHWP(a) − 5 is running. The control procedures for signal
groups S6 and S9 , S7 and S10 are based the same process.
′ ′ ′ ′

3.3.2. Variable flow control of the CLWPs


Based on the above methods, variable flow control of CLWPs is carried out to further explore the energy saving space of the system,
which is represented by Case C. The control program is presented in Table 4.
Table 4 lists the variable flow control program of all CLWPs in the system, including the coupling procedure in which one pump is
controlled by two signals. Among them, the four pumps CLWP(d) ∼ 1 − 4 corresponding to the chillers(d) are all under the two
conditions of cooling and ice storage, so the control process should include these two conditions, and the method is consistent with that
of the CHWPs described earlier. It is worth noting that the chiller(d) − 1, corresponding to CLWP(d) − 1, can be directly used for cooling
or ice storage. During ice storage, this chiller runs at full speed to ensure that the low electricity price is used as much as possible to
complete the ice storage; when cooling, CLWP(d) − 1 runs at a speed ratio (ICon CHWP(d)− 1 ). chiller(d) − 2/3/4 must be unloaded from
the existing conditions. The working speed ratios of the corresponding pumps are coupled as shown in Table 4.
The speed ratios control program of the five pumps CLWP(a) ∼ 1 − 5 connected to the chillers(ab) is also listed in Table 4. Because
CHWP(a) ∼ 3 − 5 in front of the chillers are simultaneously controlled by the signals S7 / S10 , S6 / S9 , and S5 / S8 , respectively, the
′ ′ ′ ′ ′ ′

speed ratios of the CLWP(a) ∼ 3 − 5 must be switched and integrated according to the calculation process shown in Fig. 5. Moreover,

Fig. 5. Calculation process of signal coupling control with S5’ and S8’ as examples.

9
W. Zhang et al. Case Studies in Thermal Engineering 43 (2023) 102820

Table 4
The CLWP variable flow control program of Case C model.

Pump Control program

CLWP(d) − Si · 1 + (1 − Si ) · ICon CHWP(d)− 1


1
CLWP(d) − Si ·(1 − S5 ) + S5 · ICon
′ ′
CHWP(d)− 2
2
CLWP(d) − Si ·(1 − S7 ) + S7 · ICon
′ ′
CHWP(d)− 3
3
CLWP(d) − Si ·(1 − S6 ) + S6 · ICon
′ ′
CHWP(d)− 4
4
CLWP(a) − gt(S2 , 0) ·(gt(ICLWP(a)− 1 , 1) · 1 + le(ICLWP(a)− 1 , 1) · gt(ICLWP(a)− 1 , 0.5) · ICLWP(a)− 1 + le(ICLWP(a)− 1 , 0.5) · 0.5)
1
CLWP(a) − gt(S4 , 0) ·(gt(ICLWP(a)− 2 , 1) · 1 + le(ICLWP(a)− 2 , 1) · gt(ICLWP(a)− 2 , 0.5) · ICLWP(a)− 2 + le(ICLWP(a)− 2 , 0.5) · 0.5)
2
CLWP(a) − S7 ·(1 − S10 ) · ICon
′ ′
+ S10 · ICon

S7 CHWP(a)− 3 S10 CHWP(a)− 3


′ ′

3
CLWP(a) − S6 ·(1 − S9 ) · ICon
′ ′
+ S9 · ICon

S6 CHWP(a)− 4 S9 CHWP(a)− 4
′ ′

4
CLWP(a) − S5 ·(1 − S8 ) · ICon
′ ′
+ S8 · ICon

S5 CHWP(a)− 5 S8 CHWP(a)− 5
′ ′

5
CLWP(e) − gt(S1 , 0) ·(gt(ICHWP(a)− 1 , 1) · 1 + le(ICHWP(a)− 1 , 1) · gt(ICHWP(a)− 1 , 0.5) · ICHWP(a)− 1 + le(ICHWP(a)− 1 , 0.5) · 0.5)
1
CLWP(e) − gt(S3 , 0) ·(gt(ICHWP(a)− 2 , 1) · 1 + le(ICHWP(a)− 2 , 1) · gt(ICHWP(a)− 2 , 0.5) · ICHWP(a)− 2 + le(ICHWP(a)− 2 , 0.5) · 0.5)
2

CLWP(a) − 1/2 and CLWP(e) − 1/2 only need to satisfy their own control process.

3.4. System modeling


Based on the control method and the connection of the cold source side system illustrated in Fig. 4, Case A, Case B and Case C are
modeled using TRNSYS software. Because Case C further performs variable flow control on CLWPs based on the Case B model, and the
Case B and Case C are both carried out on the basis of the actual baseline model Case A. To make the comparison criteria consistent, the
startup logic procedures of the three models and the components are built in the same way. Therefore, the final model Case C (Fig. 6) is
used as an example to present.

Fig. 6. TRNSYS model diagram of Case C.

10
W. Zhang et al. Case Studies in Thermal Engineering 43 (2023) 102820

In Fig. 6, the blue lines refer to the chilled water cycle, which are the loops for the chilled water supply and return. The red lines
connect the loops between the chillers and the c(tower)s. The pink lines are the glycol circulation that occurs between the chillers(d)
and the ice storage tank. To highlight the loops of the water supply and return pipelines in the system, the control and output lines are
uniformly represented by light gray dashed lines. Additionally, the work flow of all the equipment in the model is as detailed in the
corresponding explanation in Fig. 4. The collected real weather data is converted into the TMY2 format commonly used by TRNSYS
software and used as the input data of the c(tower)s to calculate the cooling water return temperature. The VFPs used in the system are
Type 110 from the TRNSYS component library, which is compatible with any flow rate between 0 and nominal. The ice storage
component is established based on the mathematical model of ice storage, and the exhaustive description and accuracy verification of
this component have been discussed in our previous work [18].

4. Results and discussion


Based on the above-mentioned variable flow method and the established three comparative models, the simulation results are
summarized in Table 5. Specially, it can be seen from the table that the cumulative cooling capacity of each model can cover the user
cooling load, which is a key prerequisite for subsequent analysis.

4.1. Energy consumption and CO2 emissions


The energy consumption and the CO2 emissions of the three models are summarized in Table 5. After further processing the data,
we found that the system consumption of the Case B and Case C are 9.50% and 14.15% lower than that of Case A, respectively, while the
pump energy savings are 10.20% and 28.35%, respectively. Among them, the CHWPs of the two variable flow models can save 31.95%
energy, while the CLWPs of Case B and Case C is 0.84% higher and 26.53% lower than the consumption of Case A. Regarding the
environmental impact, Case B and Case C reduce emissions by 9.45% and 14.18%, respectively, compared with Case A. To further
demonstrate the energy and environmental performance of the variable flow systems, a detailed explanation is provided next.
Fig. 7 presents the monthly system and pump energy consumption changes, all showing a pattern consistent with the cooling
capacity. Moreover, the monthly consumption of constant flow model is higher than that of the two new models. The Case C model
with variable flow control for both pumps has a lower monthly system and pump energy consumption.
To compare the proportion of the pump energy consumption to the total energy consumption more intuitively, Fig. 8 is drawn.
Fig. 8(a) illustrates that the CHWPs of the two variable flow models have a huge energy saving performance. Additionally, models
Case B and Case C, both of which carry out variable flow control on CHWPs, have almost equal CHWP consumption. In Fig. 8(b), Case C
exhibits the lowest monthly CLWP energy consumption among the three models. The monthly CLWP consumption of the Case B model
is slightly higher than that of the Case A model. This is because only the variable flow control on CHWPs, which may cause the cooling
water flow rate of the chiller to be slightly higher than what is currently required, resulting in higher CLWP consumption; however,
from the further processing of the data in Table 5, it is only 0.84% higher for the year. The above results comprehensively indicate that
variable flow systems have a significant role in promoting system energy savings. To continue the discussion of the above models from
the environmental perspective, Fig. 9 is plotted.
As can be seen from Fig. 9, the monthly emissions of the variable flow models are lower than that of Case A. This is consistent with
the previous results in Table 5, with the reductions of the two variable flow control models of 9.45% and 14.18%, respectively. This
demonstrates that the variable flow system has a positive role in reducing emissions and protecting the environment.

4.2. Economy and efficiency


Operators are often most concerned about the economy and efficiency of a system. Therefore, to evaluate the variable flow system
more comprehensively, this section will further explore these two aspects separately.
To clearly compare the economics of the three models, the monthly OCs and cooling load data obtained from the simulations are
plotted as shown in Fig. 10. The change trend of OCs is consistent with that of cooling load, and the monthly OCs of the variable flow
models are lower than that of constant flow model. Specially, the OCs of Case C is the lowest.
The OCRR data in Table 5 indicates that the two variable flow models, Case B and Case C, save 10.34% and 15.16% per year
compared with Case A model, respectively. In order to investigate the monthly changes of OCRR and efficiency of three models, Fig. 11
is drawn.
From Fig. 11, it can be concluded that the monthly OCRRCase C curves are all above OCRRCase B, that is, the monthly cost savings of
the model Case C are better than that of the model Case B that only performs variable flow on CHWPs. This is consistent with Fig. 10
that model Case C is more cost-effective.
Regarding the efficiency of the models, the right vertical axis of Fig. 11 provides the monthly operating efficiency changes of the
three models, which is similar to the OCRR trend. Additionally, in months when the load demand is not large, such as March and
November, the efficiency is high; and the months when the cold supply is large, such as July and August, the efficiency is relatively low.
During the hot season, the chillers need to work continuously to supply cold energy to users, which makes a significant increase in the
working hours and cooling capacity of chillers compared to normal periods. Therefore, the above reasons lead to a slight decrease in
efficiency. However, from the trend of the efficiency curves, it can be seen that the annual efficiency fluctuations are not significant.
Moreover, the monthly operating efficiency of the two variable flow models is higher than that of the constant flow model, and the
annual efficiency increased by 10.70% and 16.67%, respectively. In summary, the model Case C is a better cooling choice after multi-
objective comparison.

11
W. Zhang et al. Case Studies in Thermal Engineering 43 (2023) 102820

Table 5
The result statistics of the three models.

Items Case A Case B Case C

Energy consumptions (GWh) System 30.11 27.25 25.85


Pumps 7.16 6.43 5.13
CHWPs 2.41 1.64 1.64
CLWPs 4.75 4.79 3.49

Cooling load (GWh) 131.17 131.17 131.17


Cooling capacity (GWh) 146.44 146.52 146.52
CO2 emissions (million kg) 15.87 14.37 13.62
OCs (million RMB) 19.53 17.51 16.57
OCRR (%) – 10.34 15.16
η 4.86 5.38 5.67

Fig. 7. Monthly system and pump energy consumption.

Fig. 8. Monthly energy consumption of CHWPs and CLWPs.

5. Conclusions
An excellent solution for transitioning from a traditional cooling system to a sustainable system is the district cooling system (DCS),
the subject of this study. For this system, pumps account for a large proportion of energy consumption and have great potential for
energy saving. This study discussed the applicability of variable flow pumps (VFPs) in DCS and set corresponding variable flow
operation strategies to explore the performance of the new system. Based on the measured data and using the TRNSYS to establish

12
W. Zhang et al. Case Studies in Thermal Engineering 43 (2023) 102820

Fig. 9. Monthly CO2 emissions of the three models.

Fig. 10. Monthly cooling loads and OCs.

Fig. 11. Monthly OCRR and η

13
W. Zhang et al. Case Studies in Thermal Engineering 43 (2023) 102820

three models, namely the constant flow baseline model, variable flow model of chilled water pumps (CHWPs), and variable flow model
of CHWPs and cooling water pumps (CLWPs), and indicated by Case A, Case B and Case C, respectively. Then we developed the
complete control programs applicable to the above three models that can deal with the cooling load fluctuations in real time.
Moreover, the simulation results were further discussed and summarized as the following points.
1) A DCS located in the subtropics was selected as the case study. The cooling load pattern of the users is obtained by TRNSYS, and the
simulated value of the peak load is compared with the investigated value, and the error is 0.27%.
2) An in-depth analysis of the results illustrated that the Case B and Case C saved 9.50% and 14.15% energy of the system, respectively,
compared to constant flow model. Regarding pump energy consumption statistics, the two variable flow models were 10.20% and
28.35% more energy-efficient than the baseline model. Environmentally, the two newly built models reduced emissions by 9.45%
and 14.18% than that of previous model, respectively.
3) About the economy, Case B and Case C reduced the annual OCs by 10.34% and 15.16%, respectively, compared with the Case A, and
the monthly savings of Case C were better than that of Case B. Furthermore, the efficiency of the three models in months with low
cooling load was slightly higher than that in months with large cooling load, and the monthly efficiency of Case C was the highest
among the three models. Additionally, Case B and Case C improved annual efficiency by 10.70% and 16.67%, respectively. This
further demonstrated that the Case C model performed better in all of the above criteria. Additionally, the feasibility of variable
flow DCS was demonstrated.

Funding
This work was supported by the National Natural Science Foundation of China (51976027); the Project Supported by Science and
Technology Development Plan of Jilin Province (20210203200SF), China.

Author statement
Wei Zhang: Data collection, System modeling, Analysis, Writing - Original Draft.
Xu Jin: Language polishing.
Ling Zhang: Data curation.
Wenpeng Hong: Supervision and Writing.

Declaration of competing interest


The authors declare that they have no known competing financial interests or personal relationships that could have appeared to
influence the work reported in this paper.

Data availability

Data will be made available on request.

Acknowledgments
The authors would like to acknowledge the Zhuhai Hengqin Energy Development Co., Ltd. for the system parameter support.

References
[1] S. Shi, Y. Wang, Y. Wang, X. Feng, A new optimization method for cooling systems considering low-temperature waste heat utilization in a polysilicon industry,
Energy 238 (2022), 121800, https://doi.org/10.1016/j.energy.2021.121800.
[2] X.J. Luo, L.O. Oyedele, O.O. Olugbenga, A.O. Ajayi, 3D pattern identification approach for cooling load profiles in different buildings, J. Build. Eng. 31 (2020),
101339, https://doi.org/10.1016/j.jobe.2020.101339.
[3] China building energy consumption annual report, Member of Energy Consumption Statistics Committee of China Association of Building Energy Efficiency,
2020 (in Chinese).
[4] P. Jafarpur, U. Berardi, Effects of climate changes on building energy demand and thermal comfort in Canadian office buildings adopting different temperature
setpoints, J. Build. Eng. 42 (2021), 102725, https://doi.org/10.1016/j.jobe.2021.102725.
[5] T. Novosel, M. Grozdek, J. Domac, N. Duić, Spatial assessment of cooling demand and district cooling potential utilizing public data, Sustain. Cities Soc. 75
(2021), 103409, https://doi.org/10.1016/j.scs.2021.103409.
[6] Z. Zhuang, C.M. Hsieh, B. Wang, Evaluation of exhaust performance of cooling towers in a super high-rise building: a case study, Build. Simulat. 8 (2015)
179–188, https://doi.org/10.1007/s12273-014-0201-8.
[7] L. Zhang, Y. Li, H. Zhang, X. Xu, Z. Yang, W. Xu, A review of the potential of district heating system in Northern China, Appl. Therm. Eng. 188 (2021), 116605,
https://doi.org/10.1016/j.applthermaleng.2021.116605.
[8] M. Jangsten, T. Lindholm, J.O. Dalenbäck, Analysis of operational data from a district cooling system and its connected buildings, Energy 203 (2020), 117844,
https://doi.org/10.1016/j.energy.2020.117844.
[9] A. Inayat, H.H. Ang, M. Raza, B.A.A. Yousef, C. Ghenai, M. Ayoub, S.I.U.H. Gilani, Integration and simulation of solar energy with hot flue gas system for the
district cooling application, Case Stud. Therm. Eng. 19 (2020), 100620, https://doi.org/10.1016/j.csite.2020.100620.
[10] L. Zhang, J. Jing, M. Duan, M. Qian, D. Yan, X. Zhang, Analysis of district cooling system with chilled water thermal storage in hot summer and cold winter area
of China, Build. Simulat. 13 (2020) 349–361, https://doi.org/10.1007/s12273-019-0581-x.
[11] Z. He, T. Ding, Y. Liu, Z. Li, Analysis of a district heating system using waste heat in a distributed cooling data center, Appl. Therm. Eng. 141 (2018) 1131–1140,
https://doi.org/10.1016/j.applthermaleng.2018.06.036.
[12] K. Faraj, M. Khaled, J. Faraj, F. Hachem, C. Castelain, Phase change material thermal energy storage systems for cooling applications in buildings: a review,
Renew. Sustain. Energy Rev. 119 (2020), 109579, https://doi.org/10.1016/j.rser.2019.109579.

14
W. Zhang et al. Case Studies in Thermal Engineering 43 (2023) 102820

[13] H. Lund, F. Arler, P.A. Østergaard, F. Hvelplund, D. Connolly, B.V. Mathiesen, P. Karnøe, Simulation versus optimisation: theoretical positions in energy system
modelling, Energies 10 (2017) 1–17, https://doi.org/10.3390/en10070840.
[14] K. Thakar, R. Patel, G. Patel, Techno-economic analysis of district cooling system: a case study, J. Clean. Prod. 313 (2021), 127812, https://doi.org/10.1016/j.
jclepro.2021.127812.
[15] J. Kang, S. Wang, C. Yan, A new distributed energy system configuration for cooling dominated districts and the performance assessment based on real site
measurements, Renew. Energy 131 (2019) 390–403, https://doi.org/10.1016/j.renene.2018.07.052.
[16] M. Jannatabadi, H.R. Rahbari, A. Arabkoohsar, District cooling systems in Iranian energy matrix, a techno-economic analysis of a reliable solution for a serious
challenge, Energy 214 (2021), 118914, https://doi.org/10.1016/j.energy.2020.118914.
[17] A. Alajmi, M. Zedan, Energy, cost, and environmental analysis of individuals and district cooling systems for a new residential city, Sustain. Cities Soc. 54
(2020), 101976, https://doi.org/10.1016/j.scs.2019.101976.
[18] W. Zhang, W. Hong, X. Jin, Research on performance and control strategy of multi-cold source district cooling system, Energy 239 (2022), 122057, https://doi.
org/10.1016/j.energy.2021.122057.
[19] Y. Wang, Z. Wang, Z. Wang, A stochastic load demand-oriented synergetic optimal control strategy for variable-speed pumps in residential district heating or
cooling systems, Energy Build. 238 (2021), 110853, https://doi.org/10.1016/j.enbuild.2021.110853.
[20] T. Zhao, J. Zhang, L. Ma, On-line optimization control method based on extreme value analysis for parallel variable-frequency hydraulic pumps in central air-
conditioning systems, Build. Environ. 47 (2012) 330–338, https://doi.org/10.1016/j.buildenv.2011.07.007.
[21] D. Lee, H.J. Jeong, H.Y. Ji, S.Y. Park, D.Y. Chung, C. Kang, K.M. Kim, D. Park, Peak load shifting control on hot water supplied from district heating using latent
heat storage system in apartment complex, Case Stud. Therm. Eng. 34 (2022), 101993, https://doi.org/10.1016/j.csite.2022.101993.
[22] A.K.S. Al-Sayyab, J. Navarro-Esbrí, A. Mota-Babiloni, Energy, exergy, and environmental (3E) analysis of a compound ejector-heat pump with low GWP
refrigerants for simultaneous data center cooling and district heating, Int. J. Refrig. 133 (2022) 61–72, https://doi.org/10.1016/j.ijrefrig.2021.09.036.
[23] X. Xu, S. You, X. Zheng, H. Li, A survey of district heating systems in the heating regions of northern China, Energy 77 (2014) 909–925, https://doi.org/
10.1016/j.energy.2014.09.078.
[24] H. Wang, H. Wang, Z. Haijian, T. Zhu, Optimization modeling for smart operation of multi-source district heating with distributed variable-speed pumps, Energy
138 (2017) 1247–1262, https://doi.org/10.1016/j.energy.2017.08.009.
[25] X. Sheng, D. Lin, Energy saving analyses on the reconstruction project in district heating system with distributed variable speed pumps, Appl. Therm. Eng. 101
(2016) 432–445, https://doi.org/10.1016/j.applthermaleng.2016.01.059.
[26] J. Gu, J. Wang, C. Qi, X. Yu, B. Sundén, Analysis of a hybrid control scheme in the district heating system with distributed variable speed pumps, Sustain. Cities
Soc. 48 (2019), 101591, https://doi.org/10.1016/j.scs.2019.101591.
[27] A. Yan, J. Zhao, Q. An, Y. Zhao, H. Li, Y.J. Huang, Hydraulic performance of a new district heating systems with distributed variable speed pumps, Appl. Energy
112 (2013) 876–885, https://doi.org/10.1016/j.apenergy.2013.06.031.
[28] H. Wang, H. Wang, T. Zhu, A new hydraulic regulation method on district heating system with distributed variable-speed pumps, Energy Convers. Manag. 147
(2017) 174–189, https://doi.org/10.1016/j.enconman.2017.03.059.
[29] S.A. Klein, et al., TRNSYS 18: A Transient System Simulation Program, Solar Energy Laboratory, University of Wisconsin, Madison, USA, 2017. Retrieved on
9.11.2022 from, http://sel.me.wisc.edu/trnsys.
[30] X.Z. Fu, Y.M. Xiao, Fluid Network for Transportation and Distribution [M], China Architecture & Building Press, Beijing, 2018, pp. 183–189.
[31] Ministry of Housing and Urban-Rural Development of the People’s Republic of China, Design Standard for Energy Efficiency of Public Buildings: GB 50189-2015
[S], China Architecture & Building Press, Beijing, 2015. Retrieved on 9.11.2022 from, http://www.jianbiaoku.com/webarbs/book/73810/1628137.shtml.
[32] Ministry of Housing and Urban-Rural Development of the People’s Republic of China, Design Standard for Energy Efficiency of Residential Buildings in Hot
Summer and Warm Winter Zone: JGJ 75-2012[S], China Architecture & Building Press, Beijing, 2012. Retrieved on 9.11.2022 from, http://www.jianbiaoku.
com/webarbs/book/66/595561.shtml.
[33] W. Gang, S. Wang, F. Xiao, D.C. Gao, District cooling systems: technology integration, system optimization, challenges and opportunities for applications,
Renew. Sustain. Energy Rev. 53 (2016) 253–264, https://doi.org/10.1016/j.rser.2015.08.051.
[34] Ministry of Housing and Urban-Rural Development of the People’s Republic of China, Standard for Building Carbon Emission Calculation: GB/T 51366-2019
(Article description-basic rules) [S], China Architecture & Building Press, Beijing, 2019, p. 48. Retrieved on 9.11.2022 from, http://www.jianbiaoku.com/
webarbs/book/135082/3995319.shtml.
[35] TRNSYS 18 Documentation, Mathematical Reference, 4, Solar Energy Laboratory, University of Wisconsin-Madison, 2017, pp. 337–340.

15

You might also like