You are on page 1of 17

See discussions, stats, and author profiles for this publication at: https://www.researchgate.

net/publication/372311427

A simplified cooling load calculation method based on equivalent heat


transfer coefficient for large space buildings with a stratified air-conditioning
system

Article in Energy and Buildings · July 2023


DOI: 10.1016/j.enbuild.2023.113370

CITATION READS

1 160

4 authors, including:

Haidong Wang
University of Shanghai for Science and Technology
38 PUBLICATIONS 669 CITATIONS

SEE PROFILE

All content following this page was uploaded by Haidong Wang on 13 July 2023.

The user has requested enhancement of the downloaded file.


A simplified cooling load calculation method based on equivalent
heat transfer coefficient for large space buildings with a stratified
air-conditioning system
Haidong Wang1,*, Jieke Zhu1,2, Yuwei Dai1 and Hao Hu1
1School of Environment and Architecture, University of Shanghai for Science and Technology

2Institute of Engineering Technology, Shanghai Installation Engineering Group Co. Ltd.

*Corresponding email: whd@usst.edu.cn


Abstract: The traditional method for calculating the cooling load of stratified air-conditioning (STRAC) systems is
complicated, particularly for interzonal heat transfer (IZHT). This study adopted experimental and computational fluid
dynamics (CFD) methods to analyze the IZHT at two model scales and proposed a novel cooling load calculation method
for STRAC systems. This new method, which is based on the equivalent heat transfer coefficient Kc, is called the
interzonal thermal resistance method. According to the CFD simulation results of the full-scale model, the equivalent
heat transfer coefficient Kc of the occupied and unoccupied zones ranged from 42.66 to 44.36W/(m2·℃) under the nozzle
sidewall air supply (NSWAS) system. Based on the magnitude of the coefficient, the difference between the cooling load
calculated by the newly developed and traditional methods was only 4.16%, which proves the reliability of the interzonal
thermal resistance method. Moreover, compared to the traditional method, the new method is more convenient. This
method provides an alternative for calculating the cooling load of STRAC systems in practical engineering applications
with a solid theoretical foundation.

Keywords: Large space building; Cooling load calculation; Nominal thermal resistance; CFD simulation; Stratified air-
conditioning systems

Nomenclature CLQenv building envelope cooling load (W)


ρ air density (kg/m3) CLQST cooling load of STRAC systems (W)
λ thermal conductivity (W/m·℃) Qrad radiative transfer heat (W)
w vertical air velocity (m/s) Qmig convection transfer heat (W)
cp specific heat capacity of air (J/kg·℃) Qtra inter-zone heat transfer (W)
tτ-ε hourly outdoor air temperature (℃) tdownzone characteristic air temperature of occupied zone (℃)
tn indoor air temperature (℃) tupzone characteristic air temperature of unoccupied zone (℃)
Δ site correction (℃) Kc equivalent heat transfer coefficient (W/m2·℃)
δ mesh grid thickness (m) R interzonal nominal thermal resistance (m2·℃/W)
μt turbulent dynamic viscosity (Pa·s)
Prt turbulent Prandtl number
𝑚̇ mass flow rate (kg/s) Subscripts
CL geometric ratio
CTs supply air temperature ratio s the variable of supply air
CV supply air velocity ratio r the variable of return air
Cq heat flux ratio e the variable of exhaust air
CLQtotal total cooling load of the building (W) i the variable at each grid node
CLQoc occupied zone cooling load (W)
CLQunoc unoccupied zone cooling load (W)

1
1. Introduction
Large-space buildings are characterized by high floor-to-ceiling heights and large indoor spaces. Stratified air-
conditioning (STRAC) systems only condition the lower parts of buildings (occupied zone), which can save energy [1-
4]. However, the adoption of STRAC systems causes heat to accumulate in the upper zone (unoccupied zone), which is
released to the occupied zone, contributing to the cooling load [5]. Accurate interzonal heat transfer (IZHT) assessment
is crucial for cooling load prediction in large-space buildings [6-8].
The conventional calculation of the cooling load follows the hypothesis that the indoor air is the same at all positions
in the room [9,10]. However, when using STRAC systems, the vertical thermal stratification is non-negligible [11];
therefore, this hypothesis cannot be directly adopted for the cooling load calculation of STRAC systems. Many related
studies found evident thermal stratification in the vertical direction of buildings with different forms of STRAC systems.
Yang et al. [12] analyzed the computational fluid dynamics (CFD) simulation results of a room with underfloor air
distribution (UFAD) systems (a STRAC system that supplies air at ground level) and found that when the indoor heat
source power and exhaust air volume changed, there was an obvious temperature stratification between the upper and
lower areas. Ryu et al. [13] also found stable temperature stratification in an office building with UFAD systems according
to the simulation results. Xu et al. [14] studied the indoor environment of a laboratory with a nozzle sidewall air supply
(NSWAS) system (a STRAC system that supplies air at the middle height of a room) when the parameters of the supply
air were changing and found that there was also a significant stratification between the upper and lower zones. The
temperature was uniform in the area below the nozzles and there was a significant temperature gradient in the upper zone.
Although various STRAC systems are used in practice, the so-called temperature stratification dose exists in different
systems. These studies further illustrate that the calculation of the cooling load of STRAC systems is completely different
from that of conventional systems.
The cooling load calculation method for STRAC systems has been a popular topic because of its significance and
unresolved issues in the design industry. Many manuals guide the calculation of the cooling load of STRAC systems in
practical engineering applications worldwide, and the most popular methods are published by ASHRAE and other official
institutions. Bauman et al. [15] proposed a guidebook for office buildings with a UFAD system. Because the proportions
of convective and radiant heat from different types of typical heat sources are different, the influence of various heat
sources on the occupied zone cooling load is distinct. Sodec et al. [16] published a guidebook for building a UFAD system
based on European architectural features. There are also design manuals for other types of STRAC systems [17-19].
However, the cooling load calculations in these guidebooks are complicated and cannot be used directly. Moreover, few
studies have been conducted on NSWAS.
Currently in China, the cooling load calculation of STRAC systems in large-space buildings usually employs the
method described in Practical Handbook of Heating and Air-conditioning Design [20], which was proposed for the
NSWAS system, following a pilot study of large-space building cooling load calculation by the China Academy of
Building Research (CABR) [21-23]. The calculation of IZHT by using this method is complicated: convective and
radiative cooling load transfer from unoccupied to occupied zone is calculated separately, heat transfer through internal
airflow is calculated according to an empirical nomogram, and the cooling load caused by radiation is calculated
according to an empirical ratio based on the radiative coefficient. Although the results obtained using the method in this
handbook are reliable, the complexity of the method usually makes it difficult for design engineers to apply, especially
that the radiative heat transfer involves the calculation of view factors of internal surfaces. Moreover, the application of
the traditional method in the handbook is limited, it can only be applied to NSWAS systems with central air supply and
lower air return in the building, and the empirical nomogram used to calculate IZHT was derived from the experimental
results of a factory with NSWAS system. For large space building with different flow pattern other than that, the method
will become unapplicable. Therefore, a reliable, widely applicable and simple cooling-load calculation method is required
for STRAC systems.

2
Researchers have conducted few detailed studies on calculation methods for the cooling load of STRAC systems and
have proposed many new methods. Schiavon et al. [10] experimentally proposed a simplified cooling load calculation
for UFAD systems and defined the underfloor cooling load rate (UCLR) as the ratio of the cooling load of the UFAD
system to that of a well-mixed ventilation system. However, this method was derived from the experimental results of a
small office, and its application is limited. The theoretical model method is also commonly adopted to study STRAC
systems, and the block model method is widely used owing to its partitioning characteristics. The block model method is
often applied to predict the cooling load and evaluate the energy consumption. It divides the interior space vertically into
several well-mixed blocks [24]. Cai et al. [25] developed a synchronous solution model for large-space buildings with a
UFAD system based on a block model. Huang et al. [26] established a Block-Gebhart model to estimate the convective
heat transfer load of UFAD systems considering internal radiative heat transfer. However, because the block model
method considers the air parameters to be the same in each block, the calculation results are unreliable when the floor
area is large, causing significant horizontal differences.
The availability of the CFD simulation method and its widely validated accuracy make it a suitable tool for predicting
indoor environments and calculating the cooling load [27-31]. Xu et al. [32] proposed effective cooling load factors
(ECLFs) based on CFD simulations to calculate the cooling load caused by a heat source in STRAC systems. The heat
gain of the heat source multiplied by the corresponding ECLFs was equal to the effective heat gain. Cheng et al. [33]
verified the reliability of using the ECLFs to evaluate the cooling load of a STRAC system using the CFD method and
established a database of ECLFs with different heat sources. However, the method using the ECLFs only considers the
influence of the internal heat source on the interzonal transfer, while other factors, such as the heat gain due to the building
envelope, are not considered. Therefore, further verification is required before this method is used in practical engineering
applications.
The unoccupied zone transmits heat to the occupied zone via radiation and convection. Convective heat transfer
includes heat transfer caused by the air flow and temperature gradient. Tograri et al. [34] proposed the IZHT coefficient
Cb to characterize the intensity of convective heat transfer caused by the temperature gradient. Hu et al. [35] compared
the IZHT characteristics of two typical STRAC systems (floor level and NSWAS) using experimental and CFD simulation
methods. They obtained the corresponding Cb value, which provided a reference for calculating part of the IZHT of
STRAC systems. However, the Cb value only considers the heat transfer due to air temperature gradient, while the airflow
and radiative heat transfer part needs to be calculated separately. Therefore, a coefficient that considers all types of heat
transfer must be proposed so that the calculation of the IZHT can be simplified and easily used in practical engineering
applications.
To develop a simple and accurate calculation method for the cooling load of STRAC systems, this study employs a
microscopic calculation method based on CFD simulation results [5] to propose an interzonal thermal resistance model.
The equivalent heat transfer coefficient Kc values were obtained based on the results of a well-validated CFD simulation.
Kc is hypothesized to be the heat transfer across the stratified interlayer per unit time and per unit area and reflects the
intensity of the IZHT. In order to simplify the IZHT calculation for design engineers, Kc takes into consideration of both
convective and radiative heat transfer component and is supposed to be a composite coefficient. A STRAC system cooling
load calculation method based on Kc is proposed and demonstrated in this study.
Section 2 describes the specific details of the scaled-down CFD models based on an experiment and verifies their
accuracy. To make the research results applicable to real-scenario buildings, the scaled-down CFD models were extended
to full-scale CFD models based on the similarity principle, and the accuracy of the full-scale models was verified. In
Section 3, a novel cooling load calculation method for the IZHT based on equivalent heat transfer coefficient Kc values
is proposed, and the accuracy of this method is verified by comparing the cooling load calculated with the commonly
used method in Practical Handbook of Heating and Air-conditioning Design. A detailed discussion is presented in Section
4. Finally, Section 5 summarizes and concludes the paper.

3
2. Methodology

2.1. Description of experiments


A laboratory with a pitched roof was scaled down from an actual gymnasium with a geometrical scaling factor of 1:4.
The length and width of the room was 4.9 and 3.5 m, respectively. The top and bottom of the slope were 2.2 and 1.5 m
above the floor, respectively. There were no windows in the laboratory; therefore, the heat gain through the transparent
envelopes was not considered. In addition, this study did not consider the influences of indoor heat sources or infiltration.

2.1.1. Air-conditioning system

A laboratory equipped with the NSWAS system is shown in Fig. 1(a). There were 10 circular nozzles at a height of 1
m and the diameter of the nozzles was 4.35×10-2 m. The distance between the two adjacent nozzles was 0.48 m. A return
air outlet was installed at the lower central part of the western wall, below the plenum connected to the nozzles, with a
size of 0.35 × 0.35 m and its bottom was 0.64 m from the floor. Two exhaust air outlets with a size of 0.00125 × 0.35 m
(width × length) were installed in the middle of the ridge at a height of 2.2 m. An electrothermal film was patched on the
sidewalls and ceiling. The outdoor air was cooled by the air-handling unit and delivered to the plenum branched to the
ten terminal nozzles to supply conditioned air to the indoor area.

2.1.2. Specifications of instruments

There were 22 Tsic506 thermocouples vertically distributed along the central line of the laboratory at intervals of 100
mm, they were used to measure the air temperature along the central line of the laboratory. A Swa-300 hot-wire
anemometer was used to monitor the temperature and velocity of the air supply and exhaust. The return air volume was
measured using a SwemaFlow 65 air volume hood. The temperature and heat flux of the walls and ceilings were measured
using an HFP01 heat flux meter and a JTNT-A, respectively. The locations of the temperature and heat flux measurement
points are shown in Fig. 1(b). A Testo 512 differential pressure monitor was used to monitor the pressure difference
between the indoor and outdoor environments and ensure airflow balance. The technical specifications of the instruments
are listed in Table 1.

(a) Dimensions of the scaled-down laboratory. (b) Locations of measurement points.


Fig. 1. Configuration of the scaled laboratory.

Table 1 Specifications of instruments.


Instrument Parameter Range Resolution Accuracy
Tsic506 Air Temperature -10 to 60 ℃ 0.01℃ ±0.1 ℃
Swa-300 Air velocity 0.1 to 30 m/s 0.1 m/s ±3%
SwemaFlow 65 air volume hood Air volume 7 to 230 m3/h 1 m3/h ±4%
HFP01 heat flux meter Heat flux -2000 to 2000 W/m2 0.1 W/ m2 ±5%
JTNT-A Surface temperature -20 to 85 ℃ 0.1 ℃ ±0.3 ℃
Testo 512 differential pressure monitor Pressure 0 to 2 Pa 0.001 hPa 0.5% fs

4
2.1.3. Experimental procedure and results

Each experiment case ran for more than five hours to ensure steady-state condition was achieved, air flow and heat
balance was analyzed, then the heat flux, air flowrate and temperature was collected for analysis [5]. Because the exhaust
air that flows upward to purge accumulated heat is a key factor affecting the indoor thermal environment and STRAC
system cooling load [36-38], the exhaust air ratio was considered as a variable in the experiment. The air supply flowrate
of three experiment cases was kept equal, and the exhaust air duct was equipped with damper to control the exhaust air
flowrate. Table 2 lists the experimental results categorized by different ratios of exhaust air from 0 to 15%. These
experimental results are used to validate the CFD simulation results, and then extend the simulation cases to summarize
the recommended Kc value based on the validated CFD model.

Table 2 Experimental results of the scaled laboratory under three different exhaust air ratio conditions.
Parameter Unit Case 1 Case 2 Case 3
Envelope heat gain (occupied zone) W 155.37 170.67 174.82
Envelope heat gain (unoccupied zone) W 334.14 337.35 339.29
Supply air temperature ℃ 20.13 19.84 19.98
Return air temperature ℃ 26.81 26.56 26.68
Exhaust air temperature ℃ - 29.11 28.44
Supply air flowrate m3/h 215.30 215.77 216.58
Return air flowrate m3/h 217.00 201.00 187.00
Exhaust air flowrate m3/h 0 16.89 32.39
Exhaust air ratio - 0% 7.82% 14.96%

2.2. Similarity principle


Because the experiment was performed in a scaled laboratory, it was necessary that the original building and scaled-
down model were similar in order to make the results applicable to real-scenario buildings. The similarity between models
of different sizes was guaranteed according to the Reynolds number Re, Archimedes number Ar, and Prandtl number Pr
[39].
When the working fluids in the two models are the same (air in this case), Re and Ar cannot be equal simultaneously
[40]. Therefore, this study considered Ar as the main criterion, which was maintained equal for the different scale models.
The Re number was larger than 2400 to guarantee that it was in the self-similarity zone [41,42]. Similarly, a study by
Zhang et al. [43] concluded that the airflow cannot be considered Re-independent unless it is greater than 4000 at the air
outlet. To meet these requirements, Wang et al. [44] deduced the scaling ratio of each physical parameter of the scaled-
down and full-scale models, as presented in Table 3. The scaling ratio of the scaled-down model to the full-scale model
was 1:4, based on the geometric ratio of CL= 4. In this study, there was no indoor heat source or transparent envelopes,
and the heat storage of the envelopes was not considered.

Table 3 Similarity ratio of parameters deduced based on a geometric ratio of 4 [44].


Geometric ratio Supply air temperature ratio Supply air velocity ratio Heat ratio
CL CTs CV CQ

4 1 2 32

5
2.3. CFD simulation
This study proposed a nominal thermal resistance between the two zones of a large-space building to facilitate the
cooling load calculation. The experiment was performed in a scaled laboratory for validation. To extrapolate the model
to a real building and validate the results, the general-purpose CFD software PHOENICS [45] was employed to simulate
the indoor environment at the two model scales.

2.3.1. Numerical settings

The IMMERSOL model was employed to predict the radiative heat transfer of the walls and ceilings, and the
emissivity of the internal surfaces of the walls and ceilings was set to 0.93, according to color [46]. The Reynolds-
averaged Navier-Stokes equation (RANS) model has been extensively employed to simplify turbulent flows in engineering
[47-49]. The grid was generated directly in CFD software PHOENICS and structured grid was employed, and grids are
refined around nozzles, return air outlet and exhaust air outlets where large velocity gradient exists to ensure the accuracy
of the simulation. Our former research [5] demonstrated that the standard k-ω turbulence model can predict the indoor
environment of these experiments well. The control equation of the standard k-ω turbulence model as shown in Eq. (1)
and Eq. (2):
𝜕𝑘 𝜕𝑘 𝜕𝑢 𝜕 𝜕𝑘
𝜌 𝜕𝑡 + 𝜌𝑢𝑗 𝜕𝑥 = τ𝑖𝑗 𝜕𝑥 𝑖 − 𝛽 ∗ 𝜌𝑘𝜔 + 𝜕𝑥 [(𝜇 + 𝜎 ∗ 𝜇𝑡 ) 𝜕𝑥 ] (1)
𝑗 𝑗 𝑗 𝑖

𝜕𝜔 𝜕𝜔 𝜔 𝜕𝑢𝑖 𝜕 𝜕𝜔
𝜌 + 𝜌𝑢𝑗 = α τ𝑖𝑗 − 𝛽𝜌𝜔2 + [(𝜇 + 𝜎𝜇𝑡 ) 𝜕𝑥 ], (2)
𝜕𝑡 𝜕𝑥𝑖 𝑘 𝜕𝑥𝑗 𝜕𝑥𝑗 𝑗

where k is the turbulence kinetic energy and ω is the specific dissipation. μt is the turbulent viscosity. α, β, β*, σ and
σ* are all turbulence model coefficient and their value are 0.556, 0.075, 0.09, 0.5 and 0.5, respectively.
The simulation was carried out in a geometrically identical CFD scaled-down model with four different grid numbers
(0.1, 0.25, 0.5, 1, 2 and 4 million grids), and the results in Fig. 2 show that the grid-independent grid number for this
scaled-down model was 0.25 million. The selected number of grids not only met the accuracy requirements, but also
reduced the calculation time. The results were obtained after approximately 5000 iterations when the residual error was
less than 10-7 and all variables were stable, ensuring that the simulation results achieved convergence.
In the near-wall region, turbulent flow is not fully developed, and the flow is influenced by both turbulent fluctuation
and molecular viscosity. To describe the flow in the near-wall region more precisely, the RANS model introduces wall
function and a dimensionless value y+ [50]. The y+ value of the first grid layer at the wall is generally required to be
between 30 and 100 [51], with close to 30 being the most desirable. In the near-wall region, the standard wall function is
used in this paper, and the grid near the wall needs to fulfil the requirement of proper y+ value. According to the y+ value
of simulation results, it is found that the y+ value of the first layer of the wall mesh is around 35, which indicates that the
grid division is reasonable.

Fig. 2. Comparison of central line temperature Fig. 3. Schematic diagram of CFD grid model.
distribution by different grid resolutions.

6
2.3.2. Boundary conditions

The accuracy of the CFD simulation results is contingent on reliable boundary condition settings [52-54]. Based on
the results of the experiment, the floor was set as adiabatic, and the heat flux of the envelope was stable after
approximately 6 h. In this study, the cooling load was analyzed under steady-state conditions. The heat storage of the
envelope was not considered; therefore, the heat gained from the envelope was the cooling load caused by the envelope.
In the simulation of the full-scale model, the factors listed in Table 3 were employed to calculate the physical parameters
based on the experimental data from the scaled-down model. From the units of heat flux (W/m2), it can be derived that
the heat flux ratio Cq is obtained by dividing the heat ratio CQ =32 by the square of geometric ratio CL =4. Therefore, the
heat flux scale ratio Cq was 2 [35]. The heat fluxes of the walls and ceilings in the full-scale model were twice those in
the scaled-down model. All the other parameters of the full-scale model were calculated according to the scaling ratio,
i.e., in case 1, the supply air velocity and temperature were 4.14 m/s and 20.13 ℃ in the scaled-down model, respectively.
According to the velocity ratio, if CV is 2 and the supply air temperature ratio CTs is 1, the supply air velocity and
temperature in the corresponding full-scale model should be set to 8.28 m/s and 20.13 ℃, respectively. The boundary
conditions of the two models are listed in Table 4.

Table 4 CFD boundary conditions of the scaled-down and full-scale models.


Case 1 Case 2 Case 3
Boundary Parameter (unit)
SD FS SD FS SD FS
South wall (OC) 15.4 30.8 18.9 37.8 18.2 36.4
South wall (UNOC) 23.7 47.4 24.1 48.2 24.5 49.0
North wall (OC) 9.1 18.2 8.4 16.8 10.1 20.2
North wall (UNOC) 17.7 35.4 19.4 38.8 18.7 37.4
East wall (OC) Heat flux (W/m2) 4.4 8.8 5.1 10.2 4.8 9.6
East wall (UNOC) 3.4 6.8 3.4 6.8 3.4 6.8
West wall (OC) 9.7 19.4 9.6 19.2 9.5 19.0
West wall (UNOC) 5.3 10.6 5.1 10.2 5.1 10.2
Roof 11.7 23.4 11.7 23.4 11.8 23.6
Floor Adiabatic
Temperature (℃) 20.13 20.13 19.84 19.84 19.98 19.98
Supply air
Velocity (m/s) 4.14 8.28 4.15 8.30 4.17 8.34
Return air Velocity (m/s) Open Open 0.48 0.96 0.45 0.90
Exhaust air Velocity (m/s) - - Open Open Open Open
Note: SD = scaled down; FS = full scale; OC = occupied zone; UNOC = unoccupied zone.

2.4. CFD simulation validation

2.4.1. Determination of self-similarity

The Re number was calculated according to the CFD simulation results and is listed in Table 5. This demonstrates
that the Re number is much greater than the critical Re independence number of 4000 [43] for both the air supply inlet
and indoor space. Therefore, Ar was used as the primary criterion in this study.

Table 5 Results of Re number at the supply-air inlet and indoor space.


Case 1 Case 2 Case 3
Location
SD FS SD FS SD FS

Air supply inlet 11766 94131 11795 94359 11852 94813

Indoor space 36524 286961 36764 291443 37324 293790

Note: SD = scaled down; FS = ful scale.

7
2.4.2. Indoor thermal environment

The reliability of the CFD simulation results for the scaled-down model was validated using the tested vertical
temperature trend along the central line. As shown in Fig. 3, the largest relative error between the simulated temperature
and experimental results of the scaled-down model is 3.01%. This demonstrates that the indoor thermal environment of
the scaled-down model agreed well with the experimental results. The temperature trend along the central line of the full-
scale models was similar to that of the scaled-down models.
By analyzing the temperature distribution in Fig. 4, it can be seen that the jets from the nozzles divide the indoor
space into two different zones, a virtual interlayer at the height of the nozzles is considered as the stratified interlayer
from visual observation: the indoor air temperature above the interlayer is higher with greater thermal gradient, and below
the interlayer the temperature is lower and more uniform. The unoccupied zone and unoccupied zone are separated by
this virtual interlayer. The idea of occupied and unoccupied zone is to save cooling energy for STRAC system in real
world. Although the hypothesized interlayer was identified visually from temperature contour, it complies with the idea
of separating the two zones by installing air supply terminal at mid-height of the building. The indoor temperature
distributions of the two different scales are similar in the three cases. Therefore, we can further extrapolate the heat
transfer, as well as Kc and R to full-scale models according to the similarity principle.

Fig. 3. Temperature trend along the central line.

a) Scaled-down model for Case 1 (X=1.75 m) b) Full-scale model for Case 1 (X=7.00 m)

c) Scaled-down model for Case 2 (X=1.75 m) d) Full-scale model for Case 2 (X=7.00 m)

e) Scaled-down model for Case 3 (X=1.75 m) f) Full-scale model for Case 3 (X=7.00 m)
Fig. 4. Temperature distribution of the scaled-down and full-scale models under three different conditions.

8
2.4.3. Cooling load and heat transfer

As mentioned above, the attenuation and delay of the cooling load due to the heat storage of the envelope were not
considered, and the heat gained from the envelope directly becomes the cooling load caused by the envelope. The total
cooling load of the building CLQtotal and the actual cooling load of the STRAC system CLQST can be expressed using
Eqs. (3) and (4), respectively [5]:
𝐶𝐿𝑄𝑡𝑜𝑡𝑎𝑙 = 𝑐𝑝 𝑚̇𝑟 𝑡𝑟 + 𝑐𝑝 𝑚̇𝑒 𝑡𝑒 − 𝑐𝑝 𝑚̇𝑠 𝑡𝑠 (3)
̅ − 𝑐𝑝 𝑚̇𝑠 𝑡𝑠 ,
𝐶𝐿𝑄𝑆𝑇 = 𝑐𝑝 𝑚̇𝑟 𝑡𝑟 + 𝑐𝑝 (𝑚̇𝑠 − 𝑚̇𝑟 )𝑡𝑠𝑡𝑟 (4)
where cp is the specific heat capacity of air and 𝑚̇𝑠 , 𝑚̇𝑟 , and 𝑚̇𝑒 (kg/s) are the mass flow rates of the supply, return,
and exhaust air, respectively. ts, tr, and te (℃) are the temperature of the supply, return, and exhaust air, respectively. tstr
(℃) is the average temperature of the stratified interlayer.
In the experiment, the Gebhart Radiant Heat Transfer Method [55] was employed to calculate the cooling load due
to the radiative transfer heat Qrad based on the wall temperature and view factor. For the CFD simulation results, the
radiant heat flux qQRZi of each grid cell through the stratified interlayer was extracted and Qrad was calculated using Eq.
(5):
𝑄𝑟𝑎𝑑 = ∑(−𝑞𝑄𝑅𝑍𝑖 𝐴𝑖 ), (5)
where Ai is the area of each grid cell i.
From the experimental results, the convection transfer heat Qmig was calculated based on energy conservation [5].
Qmig can be expressed by Eq. (6):
𝑄𝑚𝑖𝑔 = 𝐶𝐿𝑄𝑆𝑇 − 𝐶𝐿𝑄1𝑒𝑛𝑣 − 𝑄𝑟𝑎𝑑 , (6)
where CLQ1env is the occupied zone building envelope cooling load. In the CFD simulation results, Qmig was caused by
the mass flow on the stratified interface and the conduction heat transfer through the temperature gradient. Therefore,
Qmig can be expressed as the sum of these two causes and is calculated using the CFD microscopic calculation method
based on grid cells, as shown in Eq. (7) [5]:
∑(𝐴𝑖 𝑡𝑖 ) 𝜇 (𝑡𝑖+1 −𝑡𝑖 )
𝑄𝑚𝑖𝑔 = ∑ [−𝜌𝑐𝑝 𝑤𝑖 𝐴𝑖 (𝑡𝑖 − ∑ 𝐴𝑖
)] + ∑ [(𝜆 + 𝑐𝑝 𝑃𝑟𝑡 ) 𝐴𝑖 ], (7)
𝑡 𝛿

where wi, ti, and Ai are the vertical velocity, temperature, and area in the horizontal direction of each grid cell I,
respectively. ρ, δ, and λ are the air density, thickness of the grid, and the thermal conductivity, respectively. μt is the
turbulent dynamic viscosity and Prt is the turbulent Pr number [5].
The cooling load and IZHT were obtained from the experiment and CFD simulations. As shown in Table 6, the largest
errors of CLQtotal, CLQST, Qrad, and Qmig between the experimental and CFD simulation results in the scaled-down models
were -2.27%, -0.19%, 3.60%, and 2.71% for the three cases, respectively. In addition, the vertical temperature distribution
was similar (as shown in Fig. 4), and the reliability of the scaled-down models was further proven by similar cooling
loads.
With the boundary conditions of the full-scale model (heat flux of the envelope and supply air parameters) according
to the similarity principle, the simulation results of the scaled-down and full-scale models are listed in Table 7. The results
demonstrate that the cooling load, convection transfer heat, and radiative transfer heat of the full-scale models were
approximately 32 times higher than those of the scaled-down models. This is in good agreement with the scales in Table
3.
Table 6 Cooling load error of CFD (scaled-down models) and experiment.
Parameter Case 1 Case 2 Case 3
Total cooling load CLQtotal -1.48% -2.27% 0.24%
STRAC cooling load CLQST -0.19% 1.75% 1.31%
Radiative transfer heat Qrad -2.76% 2.56% 3.60%
Convection transfer heat Qmig 1.00% 2.71% 1.14%

9
Table 7 Cooling load and heat transfer under different scales.
Parameter (unit) Case 1 Case 2 Case 3
SD FS Ratio SD FS Ratio SD FS Ratio
Total cooling load
482.28 15497.61 1/32 503.36 16332.67 1/32 511.32 16687.35 1/33
CLQtotal(W)
STRAC cooling load
482.10 15492.81 1/32 489.29 15508.41 1/32 488.85 15520.58 1/32
CLQST(W)
Radiative transfer heat
110.74 3421.51 1/31 114.45 3554.71 1/31 113.88 3539.95 1/31
Qrad(W)
Convection transfer heat
215.92 7097.35 1/33 209.75 6621.04 1/32 201.79 6438.98 1/32
Qmig(W)
Note: SD = scaled-down; FS = full-scale.

3. Result

3.1. Cooling load calculation method based on interzonal nominal thermal resistance

3.1.1. Interzonal nominal thermal resistance R and equivalent heat transfer coefficient Kc

The cooling loads of large-space buildings can be estimated from two perspectives. From an energy conservation
perspective, the total cooling load of the building CLQtotal is the sum of the occupied zone cooling load CLQoc and the
unoccupied zone cooling load, CLQunoc, as shown in Eq. (8):
𝐶𝐿𝑄𝑡𝑜𝑡𝑎𝑙 = 𝐶𝐿𝑄𝑜𝑐 + 𝐶𝐿𝑄𝑢𝑛𝑜𝑐 . (8)
When the indoor thermal environment reaches steady-state conditions, the physical parameters in the occupied and
unoccupied zones remain constant. When the internal heat source and other factors are not considered, CLQoc and CLQunoc
are equal to the building envelope cooling load CLQenv of each zone and can be derived using the cooling load coefficient
method, as shown in Eq. (9):
𝐶𝐿𝑄𝑒𝑛𝑣 = 𝐾𝐹(𝑡𝜏−𝜀 − 𝑡𝑛 + 𝛥), (9)
where K is the heat transfer coefficient of the envelope, F is the area of the envelope, tτ-ε is the hourly outdoor temperature,
tn is indoor temperature, and Δ is the site correction.
tτ-ε and Δ are adopted according to the Practical Handbook of Heating and Air-conditioning Design [20]. K and A are
based on the actual engineering design parameters. For the occupied zone, tn can be determined in the range of 20 to 27℃
according to the room function. Therefore, it is easy to calculate CLQoc. However, the indoor temperature of the
unoccupied zone is not maintained at a certain value by STRAC systems. This parameter is key for calculating the
building-envelope cooling load in the unoccupied zone. If the indoor temperature of the unoccupied zone can be obtained,
CLQunoc can be calculated using Eq. (9) as CLQoc.
From the perspective of IZHT, as shown in Eq. (10), the cooling load of the STRAC system is the sum of CLQoc and
the cooling load caused by the IZHT Qtra:
𝐶𝐿𝑄𝑆𝑇 = 𝐶𝐿𝑄𝑜𝑐 + 𝑄𝑡𝑟𝑎 . (10)
CLQoc can be obtained using Eq. (9).

Fig. 5. Schematic of cooling load composition in a large space building.


10
A schematic diagram of the above two perspectives, which discuss the building envelope cooling load CLQenv and
cooling load caused by the IZHT Qtra is shown in Fig. 5. In this study, an interzonal thermal resistance method was
proposed to calculate Qtra. Thermal stratification in the vertical direction results in a hypothetical IZHT resistance. The
hypothetical heat transfer resistance is defined as the interzonal nominal thermal resistance, R. According to the
theoretical equation for heat transfer, the interzonal nominal thermal resistance R can be expressed using Eq. (11):
𝑡𝑢𝑝𝑧𝑜𝑛𝑒 −𝑡𝑑𝑜𝑤𝑛𝑧𝑜𝑛𝑒
𝑅= 𝑄𝑡𝑟𝑎
∙ 𝐴. (11)

tupzone and tdownzone are the characteristic temperatures of the unoccupied and occupied zones, respectively, and A is the
stratified interlayer area.
For the convenience of analysis, the equivalent heat transfer coefficient Kc is derived as shown in Eq. (12):
1 𝑄𝑡𝑟𝑎
𝐾𝑐 = 𝑅 = (𝑡 , (12)
𝑢𝑝𝑧𝑜𝑛𝑒 −𝑡𝑑𝑜𝑤𝑛𝑧𝑜𝑛𝑒 )∙𝐴

where Kc is the heat flux across the stratified interface and reflects the intensity of the interzone heat transfer.
According to the above analysis, CLQunoc and Qtra can be expressed using Eqs. (13) and (14):
𝐶𝐿𝑄𝑢𝑛𝑜𝑐 = 𝐾2 𝐹2 (𝑡𝜏−𝜀 − 𝑡𝑢𝑝𝑧𝑜𝑛𝑒 + 𝛥) (13)
𝑄𝑡𝑟𝑎 = 𝐾𝑐 𝐴(𝑡𝑢𝑝𝑧𝑜𝑛𝑒 − 𝑡𝑑𝑜𝑤𝑛𝑧𝑜𝑛𝑒 ). (14)
When the unoccupied zone reaches thermal balance, the heat gained by the unoccupied zone is transferred to the
occupied zone and becomes part of the cooling load of the STRAC system. Therefore, CLQunoc is equal to Qtra and the
cooling load of the STRAC system QST is equal to the total cooling load of the building Qtotal when the indoor thermal
environment is stable. In combination with Eqs. (13) and (14), if Kc is known, Qtra and tupzone can be obtained. Therefore,
the key to calculating the cooling load is determining the correct Kc value. After obtaining reference range of Kc value for
different conditions, IZHT of practical engineering can be simply calculated by selecting the correct Kc value and
substituting it into Eqs. (13) and (14), which is much more convenient than the method in handbook. Since it is difficult
to accurately calculate Kc from experimentation, CFD simulations that can provide enough details to accurately calculate
it was employed in this research.

3.1.2. Calculation of the equivalent heat transfer coefficient Kc

The IZHT Qtra comprises both Qmig and Qrad. Therefore, Eq. (12) can be deduced using Eq. (15):
𝑄𝑚𝑖𝑔 +𝑄𝑟𝑎𝑑
𝐾𝑐 = (𝑡 . (15)
𝑢𝑝𝑧𝑜𝑛𝑒 −𝑡𝑑𝑜𝑤𝑛𝑧𝑜𝑛𝑒 )∙𝐴

For CFD simulation results, tupzone and tdownzone are the average air temperature of all grids in the occupied zone of the
CFD numerical models and the average air temperature of all grids in the unoccupied zone of the CFD numerical models,
respectively. According to the CFD simulation results, Kc was calculated, and the values are listed in Table 8. In the full-
scale and scaled-down models, the Kc ranges were from 42.66 to 44.36 W/(m2·℃) and 21.25 to 22.25 W/(m2·℃),
respectively. The Kc values of the full-scale models were nearly twice those of the scaled-down models. It was found that
when the exhaust ratio was 0 to 15%, the exhaust ratio had a minor influence on the Kc value.

Table 8 Equivalent heat transfer coefficient Kc of nozzle sidewall air-supply system.


Parameter (unit) Case 1 Case 2 Case 3
SD FS SD FS SD FS
Interzonal heat transfer (W) 348.23 11323.76 342.43 11002.34 336.78 10800.49
Characteristic temperature of unoccupied zone tupzone (℃) 28.03 28.05 27.73 27.65 27.78 27.73
Characteristic temperature of occupied zone tdownzone (℃) 26.93 26.94 26.71 26.61 26.76 26.69
tupzone — tdownzone (℃) 1.10 1.11 1.02 1.04 1.02 1.04
Equivalent heat transfer coefficient Kc (W/m2·℃) 21.25 42.66 22.52 44.36 22.20 43.30
Note: SD = scaled-down; FS = full-scale.

11
3.2. Verification of interzonal thermal resistance method

The nominal thermal resistance, although deduced from a steady-state condition study, was assumed to be compliant
with the transient heat transfer process of the cooling load. Therefore, if we follow the conventional cooling load
calculation method for the building envelope, that is, the cooling load factor method, for both the occupied and
unoccupied zones, the hypothesis of employing the nominal thermal resistance is valid.
To verify the accuracy of the interzonal thermal resistance method, this section uses a workshop explicitly described
in the Practical Handbook of Heating and Air-conditioning Design [20] as an example. The calculation results were
compared with the detailed cooling load calculated using the method described in the Practical Handbook of Heating
and Air-conditioning Design [20].
The size of the workshop was 24×12×14.6 m and the height of the occupied zone was 2.7 m. The indoor air
temperature of the occupied zone tn was 27 ℃ and the outdoor air dry-bulb temperature tw was 35.2 ℃. There was neither
a heat source in the workshop nor exhaust air. In addition, the cooling load caused by solar radiation through the windows
was not considered [20]. The boundary conditions of the workshop were consistent with the research premises of this
study.

Table 9 Thermal parameters of envelopes.


Heat transfer coefficient K (W/m2∙℃) Emissivity ε
Roof 1.02 0.88
Wall 2.04 0.92
Window 6.40 0.94
Floor 0.47 0.88

For the same demonstration building, the cooling load of the STRAC system calculated using the method in the
Practical Handbook of Heating and Air-conditioning Design [20] was 43023 W, and the cooling load calculated using
the interzonal thermal resistance method proposed in this paper was 44812 W. The error was 4.16%, which reflected the
reliability of the interzonal thermal resistance method.
4. Discussion
This study was a preliminary investigation of the STRAC system cooling load calculation method based on the IZHT.
The hypothetical thermal resistance between the unoccupied and occupied zones, considering convective, conductive,
and radiative heat transfer, provides new insight into the heat transfer process inside a large-space building with STRAC
systems.
However, because of the complexity of the problem addressed in this study, many parameters were not investigated,
such as the height of the occupied zone, the internal heat source, and the shape of the building (which affects the radiative
heat transfer by different view factors of the internal surface). Among the variables influencing the IZHT coefficient, the
type of air distribution system has a distinct impact on the magnitude of the interzonal heat migration [35]. According to
our previous study [35], the NSWAS and floor-level air supply systems cause magnitude differences in the nominal
thermal resistance.
When only part of the heat transfer (the convective heat transfer Qmig caused by the temperature gradient Qcond) was
considered, the IZHT coefficient Cb was capable of evaluating the intensity of conductive heat transfer caused by the
temperature gradient Qcond between regions, the other parts of the IZHT (convective heat transfer caused by the mass
exchange Qconv and radiative heat transfer Qrad) need to be calculated separately. A smaller Cb value indicated a larger
thermal resistance. The major factor that influenced thermal resistance was the air distribution system type [35], i.e., the
floor level air supply system had a much greater thermal resistance value, where the Cb value was approximately 4 W/m2·℃
and the Cb value of the NSWAS system was approximately 12 W/m2·℃. In addition, the Cb value of the NSWAS system

12
was unaffected by the exhaust air ratio; when the exhaust ratio was 0–15% and the Cb value ranged from 11.8 to 12.3
W/m2·℃. Mainly, the innovation of the study is to propose an equivalent heat transfer coefficient Kc that considers all
types of heat transfer, makes it easy to use in practical engineering applications.
In certain cases, the concept of Kc may be ineffective. For example, a precondition for using Kc was that a completely
stratified interlayer should be stably formed. When the supply air velocity was too high, the stratified interlayer between
the two zones was destroyed owing to the strong airflow movement near the nozzle [56]. When the supply air velocity
was too low, the supply air was discharged from the return air outlet immediately before being sent a distance away,
making it difficult to form a completely stratified interlayer [57] [58]. Therefore, for a regularly sized large-space building
with a NSWAS system, when neither the internal heat source nor any other factors mentioned above are considered and
a stable and completely stratified interlayer is formed, the recommended Kc value can be applied to a rough estimation
of the cooling load of such a similar system.

5. Conclusion
In this study, a simplified cooling load calculation method for large-space buildings with STRAC systems was
developed through CFD simulation. The reliability of the CFD scaled-down models was verified by the experimental
results, and the CFD models were scaled up to the actual building size according to the similarity principle. The
similarities in the indoor thermal environment and cooling load at different scales were compared. A model of interzonal
thermal resistance was proposed, and a nominal thermal resistance value was recommended. The conclusions are
summarized as follows:
(1) The equivalent heat transfer coefficient Kc and nominal thermal resistance R were proposed and obtained. The
results of the CFD simulation showed that Kc of the full-scale models with the nozzle sidewall air supply system ranged
from 42.66 to 44.36 W/(m2·℃).
(2) The error between the IZHT calculated using the CFD microcosmic and heat balance methods in the experiment
was within 4%. This reflects the reliability of the CFD microscopic calculation method to calculate the IZHT.
(3) The error between the occupied zone cooling load calculated by the interzonal thermal resistance method and that
calculated according to the Practical Handbook of Heating and Air-conditioning Design was only 4.16%. This indicates
that the interzonal thermal resistance method is a reliable and simplified cooling load calculation method. For similar
cases especially large-space buildings with sidewall nozzles, the cooling load can be calculated through the simplified
method using an interzonal nominal thermal resistance of 42.66 to 44.36 W/(m2·℃).

Acknowledgment
This research is supported by National Science Foundation of China under the grant number of 52278116.
References
[1] Huang, Y., Niu, J. L. A review of the advance of HVAC technologies as witnessed in ENB publications in the period from 1987 to 2014,
Energy and Buildings 130 (2016) 33-45.

[2] Cheng, Y., et al. Investigation on the thermal comfort and energy efficiency of stratified air distribution systems. Energy for Sustainable
Development, 28 (2015) 1-9.

[3] Gordon, R. L., Bagheri, H. M. Verification of stratified air condition design. ASHRAE Report 388RP (1986).

[4] Bing, W., et al. Energy saving potentials of all cold air distribution system with stratified air conditioning in large space building, Energy
Sustainability, 43192 (2008) 261-266.

[5] Wang, H., et al. On the calculation of heat migration in thermally stratified environment of large space building with sidewall nozzle air-
supply, Building and Environment, 147 (2019) 221-230.

[6] Causone, F., et al. Experimental evaluation of heat transfer coefficients between radiant ceiling and room. Energy Build, 41 (6) (2009)
13
622-628.

[7] Rhee, K. N., et al. A 50-year review of basic and applied research in radiant heating and cooling systems for the built environment.
Building and Environment, 91 (2015) 166-190.

[8] Cai, N., Huang, C. A study of cooling load calculation of stratified air conditioning system for large space based on the simultaneously
solving model. Applied Mechanics and Materials, 672 (2014) 1755-1761.

[9] Huang, C., Wang, X. Discussion of design method and optimization on airflow distribution in a large-space building with stratified air-
conditioning system. ASHRAE Transactions, 115 (2) (2009) 345-350.

[10] Schiavon, S., et al. Simplified calculation method for design cooling loads in underfloor air distribution (UFAD) systems. Energy and
Buildings, 43 (2-3) (2011) 517-528.

[11] Lam, J. C., Chan, A. L. CFD analysis and energy simulation of a gymnasium. Building and Environment, 36 (3) (2001) 351-358.

[12] Yang, X., et al. Heat transfer between occupied and unoccupied zone in large space building with floor-level side wall air-supply system,
Building Simulation 13(6) (2020) 1221-1233.

[13] Ryu, S. W., Park, D. Y. Effect of blind angles on thermal decay in the UFAD system in summer, Applied Thermal Engineering, 215
(2022) 118927.

[14] Xu, Y., et al. Study of convective heat transfer load induced by nozzle air supply in large spaces with thermal stratification based on
Block-Gebhart model, Sustainable Cities and Society 50 (2019) 101669.

[15] F. Bauman , T. Webster , C. Benedek , Cooling airflow design calculations for UFAD, ASHRAE J. 49 (2007) 36–44 .

[16] Sodec F, Craig R. The underfloor air supply system-the European experience. ASHRAE Transact 1990;96:690e5.

[17] Sandberg M. Displacement ventilation systems in office rooms. ASHRAE Transact 1989;95:1041e9.

[18] E. Mundt , P.V. Nielsen , K. Hagström , J. Railio , Displacement ventilation in non-industrial premises, in: H. Skistad (Ed.), REHVA
Guidebook, No. 1, 2002 .

[19] Q.Y. Chen , L. Glicksman , System Performance Evaluation and Design Guidelines for Displacement Ventilation, ASHRAE, Atlanta,
GA, 2003 .

[20] Lu, Y., Practical handbook of heating and air conditioning design, 2 nd edn, Beijing: China Architecture and Building Press, 2007. (in
Chinese)

[21] Zou, Y., et al. Investigation on cooling load calculation of stratified air conditioning for large space industrial plants, Journal of
Refrigeration (4) (1983) 49-56. (in Chinese)

[22] Zou, Y., et al. Research on calculation method of heat transfer load of stratified air conditioning, Heating Ventilating & Air Conditioning
(1983)11-18. (in Chinese)

[23] Zou, Y., et al., Study on calculation method of air distribution in stratified air conditioning, Journal of Refrigeration (4) (1983) 1–6. (in
Chinese)

[24] Wang, X., et al. Mathematical modeling and experimental study on vertical temperature distribution of hybrid ventilation in an atrium
building, Energy and buildings 41 (9) (2009) 907-914.

[25] Cai, N., et al. Study on a Simultaneously Solving Model for Stratified Air Conditioning under Low Sidewall Air Supply System in a
Large Space Building, Journal of Refrigeration 32 (3) (2011).

[26] Huang C., et al. Study of indoor thermal environment and stratified air-conditioning load with low-sidewall air supply for large space
based on Block-Gebhart model, Building and Environment, 147 (2019) 495-505.

[27] Cheong, K. W. D., et al. Thermal comfort study of an air-conditioned lecture theatre in the tropics, Building and Environment, 38 (1)
(2003) 63-73.

[28] Gan, G., Riffat, S. B. CFD modelling of air flow and thermal performance of an atrium integrated with photovoltaics, Building and
14
Environment, 39 (7) (2004) 735-748.

[29] Stamou, A., & Katsiris, I. Verification of a CFD model for indoor airflow and heat transfer, Building and Environment, (41) (2006)
1171–1181.

[30] Papakonstantinou, K. A., et al. Computational analysis of thermal comfort: the case of the archaeological museum of Athens, Applied
Mathematical Modelling, 24 (7) (2000) 477-494.

[31] Chen, Q. Ventilation performance prediction for buildings: A method overview and recent applications, Building and environment 44
(4) (2009) 848-858.

[32] Xu, H., et al. A method to generate effective cooling load factors for stratified air distribution systems using a floor-level air supply,
HVAC&R Research 15 (5) (2009) 915-930.

[33] Cheng, Y., et al. Cooling load calculation methods in spaces with stratified air: A brief review and numerical investigation, Energy and
Buildings (165) (2018) 47-55.

[34] Togari S., et al. A simplified model for predicting vertical temperature distribution in a large space. ASHARE Transactions, 99 (1) (1993)
84-99.

[35] Hu, H., et al. Investigation of inter-zonal heat transfer in large space buildings based on similarity: Comparison of two stratified air-
conditioning systems, Energy and Buildings (254) (2022) 111602.

[36] Li, D., et al. Numerical analysis on thermal performance of naturally ventilated roofs with different influencing parameters, Sustainable
Cities and Society (22) (2016) 86-93.

[37] Awad, A. S., et al. An experimental study of stratified flow in enclosures, Applied Thermal Engineering 28 (17-18) (2008) 2150-2158.

[38] Huang, C., et al. Measurements of indoor thermal environment and energy analysis in a large space building in typical seasons, Building
and Environment 42 (5) (2007) 1869-1877.

[39] Walker, C., et al. Reduced-scale building model and numerical investigations to buoyancy-driven natural ventilation, Energy and
Buildings 43 (9) (2011) 2404-2413.

[40] Li, A., et al. Reduced-scale model study of ventilation for large space of generatrix floor in HOHHOT underground hydropower station,
Energy and buildings 43 (4) (2011) 1003-1010.

[41] Linden, P. F. The fluid mechanics of natural ventilation, Annual review of fluid mechanics 31 (1) (1999) 201-238.

[42] Fan, C., Large space air conditioning design abroad, J HVAC 4 (1996).

[43] Zhang, Z., Xu, F., Experimental estimation of self-similarity in model experiment of top supply air system in large-space buildings,
HV&AC 4 (1) (2010) 115-118.

[44] Wang, H., et al. Study on Inter-zonal Heat Transfer for Large Space Buildings with Floor-level Side Wall Air-supply System Based on
Similarity Theory, Journal of Refrigeration, 43 (02) (2022) 89-96. (in Chinese)

[45] Ludwig, J. C., Mortimor, S., PHOENICSVR Reference Guide, London: CHAM (2010).

[46] Lu, L., et al. Calculation Method of Convection and Radiation Heat of Building Inner Surface Based on Gebhart Radiant Model, Building
Science, 33 (2) (2017) 72-76. (in Chinese)

[47] Walters, D. K., Cokljat, D. A three-equation eddy-viscosity model for Reynolds-averaged Navier–Stokes simulations of transitional flow,
Journal of fluids engineering 130 (12) (2008).

[48] Lau, J., Niu, J. L. Measurement and CFD simulation of the temperature stratification in an atrium using a floor level air supply method,
Indoor and Built Environment 12 (4) (2003) 265-280.

[49] Wilcox, D. C. Turbulence modeling for CFD, Vol. 2, DCW Industries, La Canada, CA, (1998).

[50] Zhang, L., et al. Comparison between CFD multi-dimensional simulations and towing tank tests of a standard ship model, Ship and
Ocean Engineering 49 (1) (2020) 48-51. (in Chinese)
15
[51] Niklas, K., Pruszko, H. Full scale CFD seakeeping simulations for case study ship redesigned from V-shaped bulbous bow to X-bow
hull form, Applied Ocean Research 89 (2019) 188-201.

[52] Kato, S., et al. CFD analysis of flow and temperature fields in atrium with ceiling height of 130 m, American Society of Heating,
Refrigerating and Air-Conditioning Engineers, Inc., Atlanta, GA (United States), 1995.

[53] Chow, W. K., Fung, W. Y., Numerical studies on the indoor air flow in the occupied zone of ventilated and air-conditioned space, Building
and Environment 31 (4) (1996) 319-344.

[54] Kato, S., et al. Chained analysis of wind tunnel test and CFD on cross ventilation of large-scale market building, Journal of wind
engineering and industrial aerodynamics (67) (1997) 573-587.

[55] Gebhart, B. A new method for calculating radiant transfers, ASHRAE Trans (65) (1959) 321-332.

[56] Tian, X., Cheng, Y., Liu, J., et al. Evaluation of sidewall air supply with the stratified indoor environment in a consultation room,
Sustainable Cities and Society (75) (2021) 103328.

[57] Wang, H., et al. Investigation of the quasi-periodic airflow oscillation and ventilation performance of an enclosed environment with a
jet air supply, Energy and Buildings (265) (2022) 112104.

[58] Wang, H., et al. Load characteristics of stratified air conditioning systems with nozzle air supply in large space buildings. Journal of HV
&AC (52) (4) (2022)138-145. (in Chinese)

16

View publication stats

You might also like