Journal Pre-Proof: International Journal of Refrigeration

You might also like

Download as pdf or txt
Download as pdf or txt
You are on page 1of 20

Journal Pre-proof

Single-phase flow distribution in plate heat exchangers: experiments


and models

Wenzhe Li , Pega Hrnjak

PII: S0140-7007(21)00046-3
DOI: https://doi.org/10.1016/j.ijrefrig.2021.01.026
Reference: JIJR 5036

To appear in: International Journal of Refrigeration

Received date: 13 September 2020


Revised date: 10 January 2021
Accepted date: 27 January 2021

Please cite this article as: Wenzhe Li , Pega Hrnjak , Single-phase flow distribution in plate
heat exchangers: experiments and models, International Journal of Refrigeration (2021), doi:
https://doi.org/10.1016/j.ijrefrig.2021.01.026

This is a PDF file of an article that has undergone enhancements after acceptance, such as the addition
of a cover page and metadata, and formatting for readability, but it is not yet the definitive version of
record. This version will undergo additional copyediting, typesetting and review before it is published
in its final form, but we are providing this version to give early visibility of the article. Please note that,
during the production process, errors may be discovered which could affect the content, and all legal
disclaimers that apply to the journal pertain.

© 2021 Published by Elsevier Ltd.


Highlights:

 The single-phase flow distribution in brazed plate heat exchangers is

quantified.

 Two models are built to predict the flow distribution and experimentally

validated.

 The CFD-linked model links the 3-D header CFD with the in-channel DP

calculation.

 The mechanistic model uses 1-D conservation equations to obtain the DP

in headers.

1
Single-phase flow distribution in plate heat exchangers: experiments and

models

Wenzhe Lia, Pega Hrnjaka,b,*

a
Air Conditioning and Refrigeration Center,
University of Illinois at Urbana-Champaign, 1206 West Green Street, Urbana, IL, USA
liwz310@illinois.edu; pega@illinois.edu,Tel: +1-217-390-5278
b
Creative Thermal Solutions, 2209 Willow Road, Urbana, IL, USA

ABSTRACT
Maldistribution of fluid among parallel channels is one of the main issues in applications of plate heat
exchangers. This paper presents an experimental and numerical investigation of the single-phase flow
distribution in the brazed plate heat exchangers, but the results can be applied to plate-and-frame and
plate-and-shell designs. In the experiments, the pressure profile in the heat exchanger is measured by the probes
inserted into the headers. The flow distribution is determined by the measured pressure drop across the channels
and the developed in-channel friction factor correlation. The experimental results indicate that in a U-type
brazed plate heat exchanger, the channel flow rate first increases for the first several channels near the heat
exchanger entrance due to the sudden expansion of flow in the inlet header. For the rest channels, the flow rate
decreases with the distance away from the entrance/exit of the heat exchanger. Such a distribution profile is
associated with the axial momentum transfer in the inlet header. The influence of the total flow rate on the
distribution profile is trivial, but the maldistribution is more severe with an increased number of plates. Two
distribution models presented are developed based on the principle of equal total pressure drop for all flow paths.
Two models calculate the pressure profile in the headers by 3-D CFD modeling and 1-D mass and momentum
conservation equation. The experimental results validate the models.
Keywords
Plate heat exchanger; single-phase; pressure profile; flow distribution.
Nomenclature
A area, m2
b corrugation depth, mm
cp specific heat, J kg-1 K-1
d plate thickness, mm
D diameter, m
f Fanning frictional factor
L length or width, m
ṁ mass flow rate, kg s-1
ṗ momentum flow rate, kg m s-2
P pressure, kPa
Pc corrugation pitch, mm
R radius, m
Re Reynolds number
T temperature, oC
u velocity along the channel, m s-1
v axial velocity in the header, m s-1
2
Q heat transfer, rate, W
Greek Symbols
dynamic viscosity, Pa s-1
density, kg s-3
shear stress, Pa
φ chevron angle, o
Subscripts
avg average
b bulk
c cold
ch channel
h hot/header
i inlet/channel index
in inlet header
o outlet
out outlet header
1. INTRODUCTION
Plate heat exchangers (PHEs) have been extensively used in the HVAC&R industry due to their advantages of high overall heat
transfer coefficients, low refrigerant inventory, and compactness. A PHE consists of multiple thin metal plates that are stamped
with wavy chevron or herringbone patterns. There are typically three options of assembling the plates: Brazed (BPHE), Plate-and-
Frame (P&F), and Plate-and-Shell (P&S). Each of these three options has subversions, but the essential designs are almost
identical.
In a PHE, the working fluid is distributed into the multiple parallel plate channels by the inlet header, which essentially consists of
a series of the entering ports on the plates. Then, the fluid transfers heat in the plate channels and is collected by the outlet header
to leave the heat exchanger. In such an arrangement, the flow rate of the fluid entering each channel is not equal because of
different pressure drops in the headers and that across the plate channels. Such a maldistribution generally leads to a degradation
of the heat exchanger performance [1].
Generally, flow maldistribution exists in both single-phase flow and two-phase flow, but they essentially differ from each other. In
the case of single-phase flow, the distribution is only affected by pressure drops. In the case of two-phase flow, the flow regimes in
the inlet header, or quality at the inlet to each channel also impact the distribution. In this paper, we will focus on the distribution
in single-phase flow conditions for three reasons: (1) the single-phase flow distribution in one side of the heat exchanger affects
the two-phase flow distribution in the other side; (2) it is very relevant to a single-phase heat exchanger, but also to the two-phase
heat exchangers like in the condensers where single-phase flow at the inlet characterizes the applications; (3) it is also very
important at the exit of the direct-expansion evaporators. Eventually, in the papers that follow, we will present the case of the
direct expansion (DX) evaporator.
Many studies used modeling approaches to analyze the single-phase maldistribution issue in PHEs or other similar structures,
which are also featured with the header and parallel channels arrangement. Bajura and Jones [2] developed a 1-D analytical model
for the flow distribution in the manifold system. In this model, the authors applied the 1-D mass and momentum conservation
equations to the dividing and combining manifolds to correlate the flow rate and pressure profiles in the manifolds. The governing
equations of two manifolds were connected by the discharge equation, which described the relationship between the pressure
differential between the manifolds and the lateral flow rate. The 1-D assumption oversimplified the complex flow conditions in
the manifolds, so the authors proposed two coefficients to correct. One coefficient corrected the non-uniform velocity profile in
the manifolds, and the other one accounted for the axial momentum transfer to the lateral flow. These two correction coefficients
were assumed to be constant and selected based on the experiments. Bajura and Jones [2] also concluded that the proposed
3
correction coefficients should be varying in the neighborhood of the first few laterals, where the readjustment of the velocity
profile in the dividing manifold happened. By using their model, Bajura and Jones [2] obtained the general flow distribution
profile in the reverse and parallel manifold systems and concluded that the diameter of two manifolds and the lateral tubes, the
length of the manifold, and the lateral flow resistance are the key influential factors in this issue. Bassiouny and Martin [3, 4]
extended Bajura and Jones’ model into PHEs. They proposed a characteristic parameter to determine the flow distribution profile
in PHEs and also concluded that the area ratio between the intake and exhaust ducts needs to be correctly chosen to achieve a
uniform distribution. Wang et al. [5] explained the difficulty of applying the Bernoulli equation for the manifold or header is
identifying a relative streamline to conserve energy and estimate the frictional losses. The authors confirmed that the advantage of
using the 1-D momentum conservation equation is to avoid the necessity of knowing the detailed flow patterns. They also
developed an analytical model for the manifold pipe, which used non-constant friction factor and pressure recovery factors.
However, though the correction factors were changing along the manifold, their values were still coming from the experimental
data. Further, Wang [6, 7] obtained the analytical solution of the flow distribution in the U- and Z-type manifold, which included
both friction and inertial effects into consideration. The author also concluded that the previous models developed in [3, 4, 8, 9]
were the special cases of neglecting the friction or inertial effect.
The computational fluid dynamics (CFD) tool is also extensively used to study this issue. Muhana and Novog [10] applied the
CFD tool to simulate the single-phase water flow distribution and pressure gradient in a square inlet header with 4 round tubes.
The authors confirmed that the CFD tool predicted the distribution reasonably well when compared with the experimental
measurements. Tong et al. [11] numerically examined eight geometric strategies of improving the single-phase flow distribution in
a generic manifold system with ten channels. A 2-D computational model was developed in [11], and four promising strategies
were identified from the modeling results. Moreover, their simulation also indicated that the sudden enlargement of the flow area
at the manifold inlet, which is not uncommon in practice, generally reduces the flow rate through the channels adjacent to the inlet.
In some extreme cases, the flow reversals occurred in those channels. This phenomenon was also observed by Wang et al. [12] in
their 2-D CFD simulation of a compact parallel-flow heat exchanger with single-phase water flowing inside. Their calculation
demonstrated that the jet flow was formed due to the sudden enlargement of the flow cross-sectional area at the heat exchanger
inlet, yielding a smaller pressure difference for the first several tubes near the inlet and the flow rate was hereby reduced. Due to
the complexity of the plate geometry, it is very computational resource consuming to establish a full 3-D CFD model for a heat
exchanger with many channels. Such an attempt was completed by Tsai et al. [13], who developed a 3-D CFD model for a PHE
with two corrugated channels and simulated single-phase water flow inside. The computational results of [13] showed that the
maldistribution exists even in the heat exchanger with only two channels, though the difference was relatively small (1 %). The
authors also pointed out that 3-D modeling of PHEs with more channels requires more advanced development in computer
technology.
Experimental studies, addressing the issue of single-phase flow distribution in PHEs or similar structures, are also reported in the
open literature. Bajura and Jones [2] used a PVC fabricated manifold system with air flowing through it to validate their analytical
distribution model. Each header was instrumented with the pressure taps located midway between the branching tubes. The header
pressure profiles and the differential pressure between the headers were recorded and compared with the model predictions.
Tereda et al. [14] experimentally measured the pressure profile along the inlet and outlet header in a PHE with 25 channels. The
pressure measurements were accomplished by inserting a mandrel with five axial equally spaced through holes on it into the
headers. The flow rate through each channel is then calculated based on the measured differential pressure and the single-phase
friction factor correlation. Bobbili et al. [15] compared the hydraulic performance and flow distribution in two PHEs with 21 and
81 plates. In their tests, a mobile static pressure probe was placed into the header of the heat exchangers to measure the pressure
drop of different channels. Their experimental data indicated that the maldistribution was more severe in the heat exchanger with
more plates. Bobbili et al. [15] also verified the theoretical model proposed by Bassiouny and Martin [3] was capable of predicting
the flow distribution reasonably well.
The following conclusions can be drawn from the above literature review:
1. When applying 1-D mass and momentum conservation equation in headers to develop the distribution model, three effects need
to be corrected: (a) non-uniform axial velocity profile in headers; (b) axial momentum transfer into the channels; (c) readjustment
of the velocity profile near the entrance of heat exchangers.
4
2. CFD tool can capture the flow details in PHEs. However, the complexity of plate geometry brings difficulties in the full 3-D
simulation.
3. Single-phase flow distribution in PHEs is determined by the header induced pressure drop and can be experimentally quantified
by the pressure differential through channels.
This paper presents an experimental and numerical investigation of single-phase flow distribution in BPHEs. In the experiments,
the pressure profile in the heat exchanger is measured by the probes inserted into the headers, and the flow distribution is
determined by the measured pressure drop across the channels and the developed in-channel friction factor correlation. We
examined four BPHEs with identical plate geometry but different numbers of plates (4, 10, 50, and 100 plates) and developed two
distribution models based on the principle of equal total pressure drop for all flow paths. One model calculates the pressure drop
in the headers by 3-D CFD modeling while the other finds the pressure profile by 1-D mass and momentum conservation
equations. The models are validated by the experimental results.

2. EXPERIMENTAL METHODS AND RESULTS


2.1 The facility
The experimental facility is shown in Fig. 1. It consists of two independent water loops: hot side and cold side. Two loops are
charged with the distilled water from the highest location while the system is held vacuum, thus no air pocket is trapped inside.
Two centrifugal pumps, controlled by the variable frequency drives, are used to circulate water for two loops. Two immersion
electrical heaters are in the tank for heating of the hot water loop; the tap water provides cooling for the cold-water loop. Another
pump is used to circulate the water in the tank to reduce the temperature stratification inside.
Coriolis type flow meters, absolute and differential pressure transducers, and type T (copper-constantan) thermocouples are
installed at locations as indicated in Fig 1. The sensors have been calibrated. Their range and uncertainty are listed in Table 1. The
static mixers are installed at the entrance and exit of the tested BPHE to guarantee a uniform temperature measurement by the
thermocouples.
Moreover, four temperature and pressure probes are placed into the inlet/outlet headers of the BPHEs to measure the temperature
change and pressure drop across the channels. The structure of the probe and its relative position in the tested BPHE is also shown
in Fig. 1. The probes are designed as inserting a lengthened thermocouple (outer diameter: 1.59 mm) into a stainless-steel tube
(outer diameter: 3.18 mm, inner diameter: 2.36 mm). The stainless-steel tube is blocked at the end, and a through-hole is drilled on
the side surface close to the end to measure the static pressure in the header. The tip of the thermocouple is bent into the channel
through the open hole of the stainless-steel tube, to record the inlet/outlet temperature of a channel. By moving the probes forward
and backward, the temperature change and pressure drop across different channels could be measured.
All the measurements are collected by an HP 75000 series B data logger. The control and monitoring of the data are accomplished
by VEE software.
The tested BPHEs are well insulated, with the heat loss calibrated so that the energy balance discrepancy (measured heat load
difference between hot and cold stream) is within ± 4 %.
The BPHEs with identical plate geometry but different numbers of plates are tested in the experiments. They all have the
1pass-1pass, U-type configuration. The schematic of the corrugated plate and its geometric parameters are given in Fig. 2 and
Table 2. It can be seen from Table 2 that the diameter of the feeding tube before the BPHEs is smaller than that of the header (very
typical in actual applications). Therefore, the influence of the sudden expansion in the flow area near the heat exchanger entrance
is also examined.
The inlet temperature of the hot and cold water is maintained at 50 oC, and 30 oC. Identical average channel mass flux is supplied
to both streams. For each test, the primary measurements of each flow stream are the total mass flow rate, the inlet/outlet
temperature of the BPHEs, and the temperature change /pressure drop across different channels.
Table 1: Measurement uncertainty
Measured Parameter Uncertainty
Temperature (T type) 0.1 (oC)
Absolute pressure (0 - 172 kPa) 0.11 % (full scale)
Differential pressure (0 - 7.5 kPa) 0.25 % (full scale)
5
Differential pressure (0 - 6.9 kPa) 0.25 % (full scale)
Mass flow rate 0.15 % (reading)
Calculated Parameter Uncertainty
Friction factor < 7.0 %

2.2 Data reduction


For the data reduction, fluid properties are calculated at the bulk-mean temperature given by:
𝑇𝑕,𝑖 + 𝑇𝑕,𝑜
𝑇𝑕,𝑏 = ; 𝑇𝑐,𝑏 = 𝑇𝑕,𝑏 − 𝐿𝑀𝑇𝐷 (1)
2
Where LMTD is calculated by:
𝐿𝑀𝑇𝐷 = [(𝑇𝑕,𝑖 − 𝑇𝑐,𝑜 ) − (𝑇𝑕,𝑜 − 𝑇𝑐,𝑖 )]/𝑙𝑛[(𝑇𝑕,𝑖 − 𝑇𝑐,𝑜 )/(𝑇𝑕,𝑜 − 𝑇𝑐,𝑖 )] (2)
The heat transfer rate is obtained from the average of hot side and cold side energy balance:
𝑄𝑕 = 𝑚̇𝑕 ∙ 𝑐𝑝 ∙ (𝑇𝑕,𝑖 − 𝑇𝑕,𝑜 );
𝑄𝑐 = 𝑚̇𝑐 ∙ 𝑐𝑝 ∙ (𝑇𝑐,𝑜 − 𝑇𝑐,𝑖 ); (3)
𝑄𝑎𝑣𝑔 = (𝑄𝑕 + 𝑄𝑐 )/2
The measured channel pressure drop in the 4-plate BPHE, which has only one hot channel in the middle of two cold channels, is
used to calculate the Fanning friction factor for a single channel and develop the single-phase water friction factor correlation. The
Fanning friction factor for a single channel is calculated through Eq. (4):
𝜌 ∙ 𝐷𝑒 ∙ ∆𝑃𝑐𝑕,𝑓𝑟 �
𝑓= , (4)
2 ∙ 𝐿𝑝 ∙ (𝑚̇𝑐𝑕 /𝐴𝑐𝑠 )2
In Eq. (4), 𝐷𝑒 = 2 ∙ 𝑏 is the channel effective diameter,and 𝐴𝑐𝑠 = 𝑏 ∙ 𝐿𝑤 is the flow area of the channel.
Then, the power-law equation, given in Eq. (5), is used to correlate the friction factor, 𝑓, with the channel Reynolds number, Re,
defined in Eq. (6). The constants C1 and C2 are obtained by curve-fitting. The developed friction factor correlation would be used
to determine the flow distribution among the channels in the BPHEs with 10, 50, and 100 plates.
𝑓 = 𝐶1 /𝑅𝑒 𝐶2 (5)
𝑚̇ch ∙ 𝐷𝑒
𝑅𝑒 = (6)
𝜇 ∙ 𝐴𝑐𝑠

6
Figure 1: Experimental facility and the pressure and temperature probe
Table 2: Plate geometry of BPHEs
Parameters Value
o
Chevron angle, φ, 60
Corrugation depth, b, mm 2.0
Corrugation pitch, Pc, mm 7.0
Plate thickness, d, mm 0.30
Port length, Lp, mm 172
Total length, Lv, mm 206
Port width, Lh, mm 42
Total width, Lw, mm 76
2
Heat transfer area per plate, Ap, m 0.014
Port diameter, Dp, mm 20.0
Figure 2: Schematic of the plate in BPHE Feeding tube diameter, Df, mm 16.0

2.3 Single-phase water friction factor


Fig. 3 gives the measured friction factor of single-phase water flow in a single channel of the tested BPHEs. For comparison, Fig.
3 also shows the friction factor data from Jin and Hrnjak [16], in which a different plate geometry was tested. The friction factor
correlation proposed in Focke et al. [17] is also plotted in the figure. As shown in Fig. 3, the friction factor data in Jin and Hrnjak
[16] present a clear transition from laminar to turbulent flow around the Re of 50 ~ 60. Shah and Sekulic [18] also concluded that
the transition to the turbulent flow in the channel of BPHE usually happens at a low Re (10 ~200). The tested Re in this study is
180 ~ 1600, and the measured friction factor presents a single trend, which is similar to the data in the turbulent region in Jin and
Hrnjak [16]. Therefore, the tested results of this study are in the early turbulent regime. The friction factor correlation from Focke
et al. [17] gives a more quantitatively similar result to the measured data in this study. The discrepancy in value may be attributed
to the difference in the plate geometry.
A friction factor correlation is curve-fitted by the measured data in Fig. 3 as the form of Eq. (5):
7
𝑓 = (4.33 ± 0.1317)/𝑅𝑒 (0.109±0.0051) (7)
Instead of a general correlation, Eq. (7) is only applicable to the specific plate geometry used in this study under the tested
conditions (Re=180~1500).

Figure 3: BPHE friction factor

2.4 Single-phase distribution in BPHEs


The water mass flow rate distribution in the BPHEs is obtained based on the measured pressure drop across the channels and the
developed in-channel friction factor correlation (Eq. (7)). By adopting this approach, the summation of the measured flow rate for
all channels (linear interpolation for the channels which are not measured directly) is within a difference of 5 % compared to the
total flow rate. To further reduce the error, a correction factor is applied to the measured channel flow rate, such that their
summation is equal to the total water mass flow rate. The pressure profiles along the inlet and outlet headers, and the mass flow
rate distribution of the cold stream in the BPHEs are shown in Fig. 4 to 6. The pressure profiles are given by the difference
between the pressure in the inlet/outlet header and that at the heat exchanger exit (Pexit). For each BPHE, three different average
channel mass flow rates are tested: 5, 7.5, and 10 g∙s-1 (corresponding channel Re ≈ 200 ~ 400). To compare the distributions at
different flow rates, the channel mass flow rates are plotted as the percentage of the average channel flow rate:

𝑚̇𝑐𝑕
𝑚̇𝑐𝑕 = (8)
𝑚̇𝑐𝑕,𝑎𝑣𝑔

Figure 4: Pressure profiles and flow distribution in the 10-plate BPHE

8
Figure 5: Pressure profiles and flow distribution in the 50-plate BPHE

Figure 6: Pressure profiles and flow distribution in the 100-plate BPHE


As shown in Fig. 4, the pressure change along the inlet and outlet headers of the 10-plate BPHE is almost negligible compared to
the pressure drop across the channels. This is because the tested 10-plate BPHE has a small number of channels, a short
inlet/outlet header, and a low total mass flow rate. Constant pressure in the inlet and outlet header gives an almost uniform flow
rate distribution among the channels. Moreover, different average channel mass flow rates seem to only change the absolute
pressure drop across the channels, while the shapes of the pressure profile in the headers are not affected. The distribution profiles
under three flow rates are almost overlapping.
In the 50-plate BPHE (Fig. 5), there is a significant pressure change in the headers. Following the flow direction, the static
pressure increases in the inlet header. This is because the flow rate in the inlet header decreases, which converts the dynamic
pressure to the static pressure. In the outlet header, the flow accelerates, and the static pressure is converted to the dynamic
pressure, which causes the static pressure to decrease along the flow direction. As mentioned above, the diameter of the feeding
tube (16 mm) is smaller than that of the headers (20 mm); therefore, there is a sudden expansion of the flow area near the entrance.
In the inlet header, such a flow expansion leads to a more significant pressure rise near the entrance than at the rear part. Moreover,
in the 50-plate BPHE, with a higher average channel mass flow rate, the conversion between the dynamic and static pressure is
more significant, so is the pressure change in the headers.
The header induced pressure drop significantly determines the channel flow rate distribution. In the 50-plate BPHE (Fig. 5), along
the flow direction in the inlet header, the channel flow rate first increases for the first several channels due to the entrance effect.
For the rest channels, the channel flow rate decreases with the distance away from the entrance/exit of the BPHE. This is mainly
because, in the inlet header, a portion of axial momentum is transferred into the channels; thus, the increase in the static pressure
9
rise is less than the decrease in the dynamic pressure. But in the outlet header, the flow coming from the channels contains little
axial momentum, and the increase in the dynamic pressure is all coming from the static pressure in the header. Consequently, the
static pressure change in the inlet header is smaller than that in the outlet header, and the channel pressure drop decreases with the
distance away from the entrance/exit of the BPHE, which drives less mass flow through the channels at the rear part. Besides, the
pressure loss due to the shear force in the headers further reduces the pressure change in the inlet header and increases that in the
outlet header. But the contribution of the shear force is not significant in BPHEs since the header flow only touches the edge of the
port of each metal plate.
Also, one could find in Fig. 5 the three non-dimensional distribution profiles at different flow rates are almost not distinguishable.
This is because, with a higher average mass flow rate, the pressure drop in the headers and that across the channels simultaneously
increase. These two effects compensate for each other, making the non-dimensional flow distribution unchanged. Similar results
were reported in other researches [9, 19]. Therefore, for all tested average channel flow rates, the maximum channel flow rate is
achieved in the 5th channel, and the minimum flow rate is in the last channel, and the difference between them is around 15 % of
the average channel flow rate.
Fig. 6 shows the experimental results in the 100-plate BPHE. Similar trends can be observed in the figure as the case of the
50-plate BPHE. In the 100-plate BPHE, the pressure change in the inlet and outlet header is much more significant than that in the
50-plate BPHE due to the higher total mass flow rate. The channel flow rate is also more mal-distributed in the 100-plate BPHE.
The difference between the maximum and minimum channel flow rate is around 66 % of the average channel flow rate.

3. CFD-LINKED DISTRIBUTION MODEL


A CFD-linked distribution model is developed to predict the flow rate distribution in the BPHEs. In this model, a compromise is
made between the complexity and accuracy of the simulation. The model adopts the 3-D CFD simulation for the flow in the
headers and connects two header CFD simulations by the in-channel pressure drop calculations (1-D). A similar approach was
adopted in Huang et al. [20] in microchannel heat exchangers.

3.1 CFD simulation of flow in the headers


The BPHE headers are simulated with a commercial CFD code Fluent. The 3-D meshes are generated for the headers in Gambit
software. Since the BPHE header consists of a series of ports on the plates, a small portion of the metal plate near the port, along
with the flow space in between, is included in the computational domain (Fig. 7) to consider the possible interaction between the
header flow and the metal plates. The feeding tube connected to the heat exchanger is also added. The generated meshes are
structured, consisting of only hexahedral cells. The grid independence tests have been conducted. Three different meshes with
370355, 1676241, 2242976 nodes are built for the inlet header with 49 channels (100-plate BPHE, hot side). Using these three
-1
meshes, the simulations are carried out at 𝑚̇𝑐𝑕,𝑎𝑣𝑔 =5 g s . The pressure profile in the header is used as the criterion since it is
crucial to the flow distribution. The results show that the calculated pressure profiles by these three meshes are almost the same,
which indicates the meshes are sufficiently fine for the tested case. To capture more flow details, the mesh with the largest number
of nodes is selected. For the headers with other numbers of channels, the same max cell size is kept in the mesh generations.
The generated mesh is then inputted to the Fluent for the flow simulation. In Fluent, the steady-state conservation equations for
the mass and momentum are solved using the finite volume method. The SIMPLEC scheme is used to solve for the
pressure-velocity coupling. The turbulence simulation is accomplished by the realizable k-ε model, which uses two transport
equations, one is for the turbulent kinetic energy, k, and the other for the rate of dissipation of turbulent kinetic energy, ε, to
describe turbulence.
In the calculations, the pressure inlet/outlet boundary conditions are set for the inlet/outlet of the heat exchanger. The velocity inlet
condition is used for the channel inlets/outlets, in which the fluid is assumed to flow in the vertical direction when entering/exiting
the channels. The velocity of the fluid entering/exiting the channels is calculated based on the mass flow rate. Since the mass flow
rate for each channel is unknown before the simulation, therefore, the iteration is needed to estimate a reasonable flow rate
distribution.

10
Figure 7: Mesh generation and boundary conditions for the header CFD models

3.2 Equal total pressure drop for all flow paths


In a BPHE, any flow path, starting from the heat exchanger entrance, passing through the inlet header, plate channel, outlet header,
and eventually ending at the exit should have the same total pressure drop. The proposed CFD-linked model predicts the flow rate
distribution by iteratively adjusting the flow rate through each channel to achieve an equal total pressure drop for all flow paths.
Similar approaches were adopted by many works related to the flow distribution issues [21,22].
As shown in Fig. 8, for each flow path, the total pressure drop is divided into three components: the pressure drop in the
inlet/outlet header and that across the plate channel:
𝑖 𝑖

𝛥𝑃𝑝𝑎𝑡𝑕,𝑖 = ∑ 𝛥𝑃𝑖𝑛,𝑖 + 𝛥𝑃𝑐𝑕,𝑖 + ∑ 𝛥𝑃𝑜𝑢𝑡,𝑖 (9)


1 1

In Eq. (9), the pressure drop in the inlet/outlet header is obtained from the CFD simulation, and the pressure drop across the
channel includes: the local pressure drop of turning-in and turning-out, frictional and gravitational pressure drop:
𝛥𝑃𝑐𝑕,𝑖 = 𝛥𝑃𝑡𝑢𝑟𝑛−𝑖𝑛,𝑖 + 𝛥𝑃𝑓𝑟𝑖,𝑖 + 𝛥𝑃𝑔𝑟𝑎,𝑖 + 𝛥𝑃𝑡𝑢𝑟𝑛−𝑜𝑢𝑡,𝑖 (10)
The frictional pressure drop is calculated by the previously developed friction factor correlation (Eq. (7)), and the local pressure
loss of turning-in and turning-out are calculated by the correlations in Idel’chik [23] (Section VII, Diagram 7-21; Section VII,
Diagram 7-7).

Figure 8: Flow paths and total pressure drop calculation in BPHEs

11
3.3 Flow details in the headers of the BPHE
The CFD-linked distribution model predicts the single-phase flow distribution in the BPHE and simultaneously simulates the flow
field in the headers. The comparison between the predicted and measured flow rate distribution will be presented in section 4 of
this paper. Fig. 9 depicts the typical contours of the axial velocity and the pressure profile of the center plane in the inlet and outlet
header of the BPHE. The pressure profile is given by the pressure difference to the header inlet/outlet.
The axial velocity contour, as well as the local streamlines, in the inlet header, shows that there is a clear sudden expansion flow
near the entrance of the heat exchanger, induced by the size difference between the feeding tube and the header. The readjustment
of the velocity profile takes a distance of several channels. Due to this sudden expansion flow, the eddies are generated near the
inlet of the first several channels, and the axial velocity in this region is negative, as demonstrated by the streamlines. For the
channels after the sudden expansion region, the axial velocity at the channel inlet is positive, which means the axial momentum is
transferred to the channels. It is also observed from the streamlines that the fluid near the channel inlets branches out first and the
fluid closer to the center flows toward the boundary of the header and branches through downstream channels. The sudden
expansion near the entrance also leads to an abrupt pressure rise in this region, as shown in the pressure profile. At the
downstream of the expansion region, the pressure increase, caused by the conversion between the dynamic and static pressure, is
much more moderate.
The velocity profile in the outlet header, also given in Fig. 9, is more uniform than that in the inlet header. The vertical flow in the
channels, without too much axial momentum, joins the axial flow in the header, which makes the velocity of the fluid near the
channel outlets lower than that away from the channel outlets. The fluid also experiences a higher pressure change in the outlet
header than that in the inlet header (the sudden expansion region excluded) due to the contribution of the axial momentum transfer
and the shear force.

-1
Figure 9: Axial velocity and pressure profile in the headers (𝑚̇𝑐𝑕,𝑎𝑣𝑔 =10 g s , 50-plate BPHE, cold side)

4. MECHANISTIC DISTRIBUTION MODEL


As discussed in the literature review, three effects need to be corrected when developing a 1-D analytical distribution model. The
12
usual practice in the literature is applying empirical correction factors in the momentum conservation equation. Therefore, the
generality of the models is limited.
In this section, a mechanistic distribution model is developed, which accounts for these three effects mechanistically. This
mechanistic model also predicts the distribution by imposing the condition of equal total pressure drop for all flow paths. Instead
of calculating the header pressure profile by the CFD tool, the mechanistic model estimates the pressure change in the headers by
mass and momentum conservation equations.

4.1 Pressure change in the inlet header


The mass and momentum conservation equations are applied to a control volume in the inlet header, as shown in Fig. 10:

∫ 𝜌𝑣𝑖𝑛,𝑖 𝑑𝐴𝑕 = ∫ 𝜌𝑣𝑖𝑛,𝑖+1 𝑑𝐴𝑕 + ∫ 𝜌𝑢𝑐𝑕,𝑖 𝑑𝐴𝑐𝑕 (11)

2 2
𝑃𝑖𝑛,𝑖 𝐴𝑕 − 𝑃𝑖𝑛,𝑖+1 𝐴𝑕 − ∙ 𝐴𝑠𝑕𝑒𝑎𝑟 = ∫ 𝜌𝑣𝑖𝑛,𝑖+1 𝑑𝐴𝑕 − ∫ 𝜌𝑣𝑖𝑛,𝑖 𝑑𝐴𝑕 + ∫ 𝜌𝑢𝑐𝑕,𝑖 𝑣𝑐𝑕,𝑖 𝑑𝐴𝑐𝑕 (12)

In the momentum conservation Eq. (Eq. (12)), the left-hand side is the axial force exerting on the control volume, and the
right-hand side is the axial momentum flow rate leaving the control volume.

Figure 10: Control volume in the inlet header


It should be mentioned that in BPHEs, the flow in the header only touches the edge of the metal plates, as Fig. 10 shows. The
shear force in Eq. (12) is due to the axial velocity gradient between the header flow and the branching-out channel flow. Therefore,
the shear force has a very limited influence on the flow in the header.
The axial momentum flow rates in Eq. (12) are calculated mechanistically by tracking the evolution of the velocity profile in the
inlet header. The evolution of the velocity profile is based on the fact that the fluid near the channel inlets tends to branch out first,
and the dynamic energy of this part of fluid would not be recovered in the header flow.
To implement this method, an initial velocity profile at the heat exchanger entrance is first assumed. For the fully developed
turbulent flow, the 1/7th power-law velocity profile is used (Eq. (13)). In Eq. (13), the origin of the Cartesian coordinate is at the
center of the cross-section.
1 ⁄7
√𝑥 2 + 𝑦 2
𝑣𝑖𝑛,1 = 1.22 ∙ 𝑣̅𝑖𝑛,1 ∙ (1 − ) (13)
𝑅

Fig. 11 demonstrates the procedure of estimating the downstream velocity profile based on the upstream information. For
generality, the velocity profile at the ith cross-section is assumed to be known. The momentum flow rate at this cross-section is
calculated by the numerical integration:
𝑥 2 +𝑦 2 =𝑅 2
2 (14)
𝑝̇𝑖 = ∫ 𝜌 ∙ 𝑣𝑖𝑛,𝑖 ∙ 𝑑𝑥𝑑𝑦
0
It is convenient to assume the branching is mainly in the upward direction (+y direction), and thus the fluid at the top of the header
tends to branch out first. An elliptical boundary is drawn on the velocity profile, as shown in Fig.11, with the assumption that the
flow above this boundary branches out through the ith channel and that below this boundary moves forward in the header. The
mass conservation requires the flow rate above the elliptical boundary equal to the presumed channel flow rate, 𝑚̇𝑐𝑕,𝑖 :

13
𝑥 2 +𝑦 2 =𝑅 2 ,𝑦>0
𝑚̇𝑐𝑕,𝑖 = ∫ 𝜌 ∙ 𝑣𝑖𝑛,𝑖 ∙ 𝑑𝑥𝑑𝑦 (15)
𝑦2
𝑥 2 + 2 =𝑅 2 ,𝑦>0
𝑟𝑖

𝑟𝑖 in the above equation is called the elliptical ratio, which determines the shape of the elliptical boundary. Since in the calculation
of the pressure profile, the channel flow rate 𝑚̇𝑐𝑕,𝑖 , is treated as a known variable, the elliptical ratio 𝑟𝑖 , can be solved
numerically from the Eq. (15). The axial momentum flow rate contained in the branching-out is calculated by:
𝑥 2 +𝑦 2 =𝑅 2 ,𝑦>0
2
𝑝̇𝑐𝑕,𝑖 = ∫ 𝜌 ∙ 𝑣𝑖𝑛,𝑖 ∙ 𝑑𝑥𝑑𝑦 (16)
𝑦2
𝑥 2 + 2 =𝑅 2 ,𝑦>0
𝑟𝑖

With the fluid above the elliptical boundary branching out, the remaining velocity profile below the boundary expands to fill the
entire cross-section of the header and form a new velocity profile at the (i+1)th cross-section. With the elliptical boundary, the
expansion of the velocity profile is linear in the vertical direction. In other words, to fill the entire cross-section, the remaining
2
velocity profile only needs to be stretched in the vertical direction by a ratio of , as shown in Fig. 11. The relationship
1+𝑟𝑖

between two corresponding points on the velocity profiles before and after the stretch is:
𝑥𝑖+1 = 𝑥𝑖 ;

1 − 𝑟𝑖 2𝑦𝑖 (17)
𝑦𝑖+1 = √𝑅2 − 𝑥𝑖2 ∙ + ;
1 + 𝑟𝑖 1 + 𝑟𝑖
In Eq. (17), (𝑥𝑖 , 𝑦𝑖 ) is a point on the velocity profile at ith cross-section; (𝑥𝑖+1 , 𝑦𝑖+1 ) is a corresponding point on the velocity profile
at (i+1)th cross-section. With the mass conservation. the velocity at the (i+1)th cross-section is given by:
(𝑟𝑖 + 1)
𝑣𝑖𝑛,𝑖+1 (𝑥𝑖+1 , 𝑦𝑖+1 ) = ∙ 𝑣𝑖𝑛,𝑖 (𝑥𝑖 , 𝑦𝑖 ) (18)
2
The momentum flow rate at (i+1)th cross-section is:
𝑥 2 +𝑦 2 =𝑅 2
2 (19)
𝑝̇𝑖+1 = ∫ 𝜌 ∙ 𝑣𝑖𝑛,𝑖+1 ∙ 𝑑𝑥𝑑𝑦
0
The pressure difference between the ith and (i+1)th cross-section is then calculated by the momentum conservation equation:
𝑃𝑖𝑛,𝑖 𝐴𝑕 − 𝑃𝑖𝑛,𝑖+1 𝐴𝑕 = 𝑝̇𝑖+1 − 𝑝̇𝑖 + 𝑝̇𝑐𝑕,𝑖 + ∙ 𝐴𝑠𝑕𝑒𝑎𝑟 (20)

Figure 11: Estimation of the velocity profile


This mechanistic model also takes the sudden expansion flow at the heat exchanger entrance into consideration. For the flow in an
abruptly expanding duct, the turbulent reattachment length, which is the distance between the expansion and reattachment location,
is about 7~9 step heights [24]. In BPHEs, one step height is the radius difference between the feeding tube and the header. The
14
reattachment should be easier with the fluid branching out through the channels. Therefore, the expansion length in this model is
selected to be the lower bound, 7 step heights. For the BPHEs used in this study, it is about 4 channels. For other BPHEs, it may
not be 4 channels, but can be calculated by the size of the feeding tube and the header. To simulate the readjustment of the velocity
profile, the diameter of the core flow is assumed to expand linearly from the diameter of the entrance to that of the header over the
expansion length. The control volume in the sudden expansion region is depicted in Fig. 12.

Figure 12: Control volume in the sudden expansion region


With this assumption, when calculating the axial momentum flow rate at the ith cross-section by Eq. (14), and the elliptical ratio by
Eq. (15), the radius of the header, R, needs to be replaced by the radius of the core flow, R i.
As shown in the CFD simulation (Fig. 9), in the sudden expansion region, eddies are generated near the inlet of the channels, and
the fluid enters the channels with negative axial velocity. Therefore, Eq. (16) is not applicable in the expansion region, and the
momentum flow rate transfer through the channel is estimated by:
𝑝̇𝑐𝑕,𝑖 = 𝑚̇𝑐𝑕,𝑖 ∙ 𝑣̄ ch,i (21)
With: 𝑣̅𝑐𝑕,𝑖 ≈ −0.1 ∙ 𝑣̅𝑖𝑛,𝑖 [25].
When estimating the velocity profile at the (i+1)th cross-section in the sudden expansion region, in addition to the stretch, the
expansion of the core flow also needs to be considered. Eq. (17) is modified as:
𝑅𝑖+1
𝑥𝑖+1 = 𝑥𝑖 ∙ ;
𝑅𝑖
(22)
1 − 𝑟𝑖 2𝑦𝑖 𝑅𝑖+1
𝑦𝑖+1 = (√𝑅2 − 𝑥𝑖2 ∙ + )∙ ;
1 + 𝑟𝑖 1 + 𝑟𝑖 𝑅𝑖
The velocity profile at the (i+1)th cross-section is given by:
(𝑟𝑖 + 1) 𝑅𝑖 2
𝑣𝑖𝑛,𝑖+1 (𝑥𝑖+1 , 𝑦𝑖+1 ) = ∙ 𝑣𝑖𝑛,𝑖 (𝑥𝑖 , 𝑦𝑖 ) ∙ ( ) (23)
2 𝑅𝑖+1
The axial momentum flow rate at the (i+1)th cross-section is then calculated by Eq. (19) after replacing the radius of the header, R,
by the radius of the core flow, Ri+1. The corresponding pressure difference between the ith and (i+1)th cross-section is given by:
𝑃𝑖𝑛,𝑖 𝐴𝑕 − 𝑃𝑖𝑛,𝑖+1 𝐴𝑕 = 𝑝̇𝑖+1 − 𝑝̇𝑖 + 𝑝̇𝑐𝑕,𝑖 (24)

4.2 Pressure change in the outlet header


The control volume for the outlet header is shown in Fig. 13. Because the flow direction in the channels has been regulated by the
metal plates to be vertical, there is almost no axial momentum entering the outlet header. Besides, though there is a sudden
contraction of the flow area at the exit of the heat exchanger, the influence of this flow sudden contraction on the outlet header
flow is quite limited for the studied cases, as shown in the CFD results (Fig. 9). Moreover, there is no initial velocity profile for
the flow in the outlet header. Therefore, the flows in the outlet header should have more similarities, even with various geometries
and conditions. For simplicity, the 1-D flow approximation is made for the outlet header, and the averaged velocity is used to
calculate the pressure change:
𝜌𝐴𝑕 𝑣̄ 𝑜𝑢𝑡,𝑖 = 𝜌𝐴𝑕 𝑣̄ 𝑜𝑢𝑡,𝑖+1 + 𝑚̇𝑐𝑕,𝑖 (25)
2 2
𝑃𝑜𝑢𝑡,𝑖 𝐴𝑕 − 𝑃𝑜𝑢𝑡,𝑖+1 𝐴𝑕 = 𝛽𝑜𝑢𝑡 ∙ 𝜌𝐴𝑕 𝑣̄ 𝑜𝑢𝑡,𝑖+1 − 𝛽𝑜𝑢𝑡 ∙ 𝜌𝐴𝑕 𝑣̄ 𝑜𝑢𝑡,𝑖 − ∙ 𝐴𝑠𝑕𝑒𝑎𝑟 (26)
𝛽𝑜𝑢𝑡 in the above equation is to compensate for the non-uniformity of the axial velocity profile in the outlet header. 𝛽𝑜𝑢𝑡 ≈ 1.30
is suggested to be reasonably suitable for most cases ([2], [26]).

15
Figure 13: Control volume in the outlet header
With the pressure change in the headers calculated by the method introduced in section 4.1 and section 4.2, the flow rate
distribution can be predicted by imposing the condition of equal total pressure drop for all flow paths as discussed in section 3.2.

5. MODEL VALIDATIONS
The predictions of two newly proposed models are plotted and compared with the experimental results in Fig. 14. The modeling
results agree well with the experimental measurements. The deviations of the predictions by the mechanistic model are slightly
larger than those by the CFD-linked model due to more idealized simplifications used in the simulations. The predictions for the
100-plate BPHE are less accurate than those for the 50-plate BPHE since the flow in the headers is more complicated with a
higher mass flow rate in a larger heat exchanger. Models always underestimate the flow rate through the first several channels,
which indicates that further modification should focus on the simulation of the sudden expansion of the flow at the heat exchanger
entrance.
The highlight of the proposed mechanistic model is analytically tracking the evolution of the axial velocity profiles in the inlet
header. Fig. 15 compares the velocity profiles in the inlet header of a specific case, obtained by the mechanistic model and by the
CFD tool. It can be seen in Fig. 15, the mechanistic model is capable of simulating the “upward stretch” trend in the evolution of
the velocity profile. The basic shapes of the velocity contours given by the two methods are similar. However, one could also
observe that some flow physics are not captured by the mechanistic model, like the recirculating flow at the bottom of the header.
In conclusion, the mechanistic model has the advantage of easy application to various heat exchanger geometry and operating
conditions while scarifying the accuracy due to the 1-D simplifications.

50-plate BPHE, 𝑚̇𝑐𝑕,𝑎𝑣𝑔 =5 g s-1 100-plate BPHE, 𝑚̇𝑐𝑕,𝑎𝑣𝑔 =5 g s-1

16
50-plate BPHE, 𝑚̇𝑐𝑕,𝑎𝑣𝑔 =7.5 g s-1 100-plate BPHE, 𝑚̇𝑐𝑕,𝑎𝑣𝑔 =7.5 g s-1

50-plate BPHE, 𝑚̇𝑐𝑕,𝑎𝑣𝑔 =10 g s-1 100-plate BPHE, 𝑚̇𝑐𝑕,𝑎𝑣𝑔 =10 g s-1

Figure 14: Model predictions vs. experiments

-1
Figure 15: Axial velocity profiles in the inlet header (50-plate BPHE, 𝑚̇𝑐𝑕,𝑎𝑣𝑔 =10 g s )

6. SUMMARY AND CONCLUSIONS


The single-phase flow distribution in the BPHEs is investigated experimentally and numerically. The pressure profile in the
17
BPHEs is measured by the probes inserted into the headers, and the flow distribution is determined by the measured pressure drop
across the channels and the developed in-channel friction factor correlation. Two distribution models are developed based on the
principle of equal total pressure drop for all flow paths.
The main results in this study are summarized as the following:
1. Along the flow direction in the inlet header, the channel flow rate first increases for the first several channels due to the
sudden expansion of the flow area at the heat exchanger entrance. For the rest channels, the flow rate generally decreases
with the distance away from the entrance/exit of the BPHE. This is because the inlet header has a relatively smaller
pressure change, which is associated with the axial momentum transfer to the channels.
2. In the BPHEs, the influence of the average channel flow rate on the distribution profile is trivial, but the maldistribution
is more severe as the number of plates increases.
3. The proposed CFD-linked model simulates the flow in the headers by the 3-D CFD modeling and calculates the pressure
drop across the channel by the 1-D correlations. The flow distribution is predicted by imposing the condition of equal
total pressure drop for all flow paths.
4. The mechanistic model uses 1-D mass and momentum conservation equations to calculate the pressure change in the
headers. The effects of the non-uniform velocity profile, sudden expansion at the entrance, and axial momentum transfer
are mechanistically considered by tracking the evolution of the axial velocity profile in the inlet header.
5. Compared with the experimental results, the predictions by the two models demonstrate a reasonably good agreement.
The CFD-linked model has slightly higher accuracy, but the mechanistic model has the advantage of easy application to
various heat exchanger geometry and operating conditions.

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.

ACKNOWLEDGMENTS
The authors thankfully acknowledge the support provided by the Air Conditioning and Refrigeration Center at the University of
Illinois at Urbana-Champaign and the Creative Thermal Solutions, Inc.

REFERENCES
[1] B.P. Rao, B. Sunden, S.K. Das, An experimental and theoretical investigation of the effect of flow
maldistribution on the thermal performance of plate heat exchangers, J. Heat Transfer, 127(3), (2005),
332-343.
[2] R.A. Bajura, E.H. Jones Jr, Flow distribution manifolds, J. Fluids Eng. Trans. ASME, 98(4), (1976),
654-665.
[3] M.K. Bassiouny, H. Martin, Flow distribution and pressure drop in plate heat exchangers-I: U-type
arrangement, Chemical Engineering Science, 39(4), (1984a), 693-700.
[4] M.K. Bassiouny, H. Martin, Flow distribution and pressure drop in plate heat exchangers-II: Z-type
arrangement, Chemical Engineering Science, 39(4), (1984b), 701-704.
[5] J.Y. Wang, Z.L. Gao, G.H. Gan, D.D. Wu, Analytical solution of flow coefficients for a uniformly
distributed porous channel, Chem. Eng. J., 84 (1), (2011), 1–6.
[6] J.Y. Wang, Pressure drop and flow distribution in parallel-channel of configurations of fuel cell stacks:
U-type arrangement, Int. J. Hydrogen Energy, 33(21), (2008), 6339–6350.
[7] J.Y. Wang, Pressure drop and flow distribution in parallel-channel of configurations of fuel cell stacks:
Z-type arrangement, Int. J. Hydrogen Energy, 35, (2010), 5498–5550.
[8] R.J. Kee, P. Korada, K. Walters, M. Pavol, A generalized model of the flow distribution in channel
18
networks of planar fuel cells, J. Power Sources, 109, (2002), 148-159.
[9] S. Maharudrayya, S. Jayanti, A.P. Deshpande, Flow distribution and pressure drop in parallel-channel
configurations of planar fuel cells, J. Power Sources, 144, (2005), 94–106.
[10] A. Muhana, D.R. Novog, Validation of Fluent for prediction of flow distribution and pressure gradients
in a multi-branch header under low flow conditions, Proceedings of the 16th International Conference on
Nuclear Engineering, Orlando, (2008), paper 48128.
[11] J.C.K. Tong, E.M. Sparrow, J.P. Abraham, Geometric strategies for attainment of identical outflows
through all of the exit ports of a distribution manifold in a manifold system, Applied Thermal Engineering,
29, (2009), 3552–3560.
[12] C. Wang, K. Yang, J. Tsai, I.Y. Chen, Characteristics of flow distribution in compact parallel flow heat
exchangers, part I: Typical inlet header, Applied Thermal Engineering, 31, (2011), 3226-3234.
[13] Y. Tsai, F. Liu, P, Shen, Investigations of the pressure drop and flow distribution in a chevron-type plate
heat exchanger, International Communications in Heat and Mass Transfer, 36, (2009), 574–578.
[14] F.A. Tereda, N. Srihari, S.K. Das, B. Sunden, Experimental investigation on port-to-channel flow
maldistribution in plate heat exchangers, Heat Transfer Engineering, 28(5), (2007), 435-443.
[15] P.R. Bobbili, B. Sunden, S.K. Das, An experimental investigation of the port flow maldistribution in
small and large plate package heat exchangers, Applied Thermal Engineering, 26, (2006), 1919–1926.
[16] S. Jin, P. Hrnjak, Effect of end plates on heat transfer of plate heat exchanger. Int. J. Heat and Mass
Transfer, 108, (2017), 740–748.
[17] W.W. Focke, J. Zachariades, I. Olivier, The effect of the corrugation inclination angle on the
thermohydraulic performance of plate heat exchangers. Int. J. Heat and Mass Transfer, 28(8), (1985),
1469-1479.
[18] R.K. Shah, D.P. Sekulic, Fundamentals of Heat Transfer Design, first ed., Wiley, 2002.
[19] J. Wang, Theory of flow distribution in manifolds, Chem. Eng. J. 168, (2011), 1331–1345.
[20] L. Huang, M.S. Lee, K. Saleh, V. Aute, R. Radermacher, A computational fluid dynamics and
effectiveness-NTU based co-simulation approach for flow mal-distribution analysis in microchannel heat
exchanger headers, Applied Thermal Engineering, 65, (2014), 447-457.
[21] H. Tuo, A. Bielskus, P. Hrnjak, Experimentally validated model of refrigerant distribution in a parallel
microchannel evaporator, SAE Int. J. Mater. Manf., 5(2), (2012), 365-374.
[22] J. Li, P. Hrnjak, An experimentally validated model for microchannel condensers with separation
circuitry, Applied Thermal Engineering, ATE-D-20-00859.
[23] I.E. Idel'chik, Handbook of hydraulic resistance, CRC Press, 1994.
[24] G.L. Morrison, G.B. Tatterson, M.W. Long, Three-dimensional laser velocimeter investigation of
turbulent, incompressible flow in an axisymmetric sudden expansion. Journal of Propulsion and Power, 4(6),
(1988), 533-540.
[25] W.J. Devenport, E.P. Sutton, An experimental study of two flows through an axisymmetric sudden
expansion, Experiments in Fluids, 14, (1993), 423-432.
[26] R.A. Bajura, A model for flow distribution in manifolds, J. Eng. Power, 93(1), (1971), 7-12.

19

You might also like