You are on page 1of 12

International Journal of Refrigeration 126 (2021) 45–56

Contents lists available at ScienceDirect

International Journal of Refrigeration


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

Single-phase flow distribution in plate heat exchangers: Experiments


and models
Wenzhe Li a, Pega Hrnjak a,b,∗
a
Air Conditioning and Refrigeration Center, University of Illinois at Urbana-Champaign, 1206 West Green Street, Urbana, IL, United States
b
Creative Thermal Solutions, 2209 Willow Road, Urbana, IL, United States

a r t i c l e i n f o a b s t r a c t

Article history: Maldistribution of fluid among parallel channels is one of the main issues in applications of plate heat
Received 13 September 2020 exchangers. This paper presents an experimental and numerical investigation of the single-phase flow
Revised 10 January 2021
distribution in the brazed plate heat exchangers, but the results can be applied to plate-and-frame and
Accepted 27 January 2021
plate-and-shell designs. In the experiments, the pressure profile in the heat exchanger is measured by
Available online 29 January 2021
the probes inserted into the headers. The flow distribution is determined by the measured pressure drop
Keywords: across the channels and the developed in-channel friction factor correlation. The experimental results
Plate heat exchanger indicate that in a U-type brazed plate heat exchanger, the channel flow rate first increases for the first
Single-phase several channels near the heat exchanger entrance due to the sudden expansion of flow in the inlet
Pressure profile header. For the rest channels, the flow rate decreases with the distance away from the entrance/exit of
Flow distribution 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.
© 2021 Elsevier Ltd and IIR. All rights reserved.

Distribution de l’écoulement monophasique dans les échangeurs de chaleur à


plaques: expériences et modèles

Mots clés: Échangeurs de chaleur à plaques; Monophase; Profil de pression; Distribution de l’écoulement

1. Introduction In a PHE, the working fluid is distributed into the multiple par-
allel plate channels by the inlet header, which essentially consists
Plate heat exchangers (PHEs) have been extensively used in of a series of the entering ports on the plates. Then, the fluid trans-
the HVAC&R industry due to their advantages of high overall heat fers heat in the plate channels and is collected by the outlet header
transfer coefficients, low refrigerant inventory, and compactness. A to leave the heat exchanger. In such an arrangement, the flow rate
PHE consists of multiple thin metal plates that are stamped with of the fluid entering each channel is not equal because of different
wavy chevron or herringbone patterns. There are typically three pressure drops in the headers and that across the plate channels.
options of assembling the plates: Brazed (BPHE), Plate-and- Frame Such a maldistribution generally leads to a degradation of the heat
(P&F), and Plate-and-Shell (P&S). Each of these three options has exchanger performance (Rao et al., 2005).
subversions, but the essential designs are almost identical. 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

Corresponding author. pressure drops. In the case of two-phase flow, the flow regimes in
E-mail addresses: liwz310@illinois.edu (W. Li), pega@illinois.edu (P. Hrnjak). the inlet header, or quality at the inlet to each channel also impact

https://doi.org/10.1016/j.ijrefrig.2021.01.026
0140-7007/© 2021 Elsevier Ltd and IIR. All rights reserved.
W. Li and P. Hrnjak International Journal of Refrigeration 126 (2021) 45–56

on the experiments. Bajura and Jones (Bajura and Jones, 1976)


Nomenclature also concluded that the proposed correction coefficients should
be varying in the neighborhood of the first few laterals, where
A area, m2 the readjustment of the velocity profile in the dividing manifold
b corrugation depth, mm happened. By using their model, Bajura and Jones (Bajura and
cp specific heat, J kg−1 K − 1 Jones, 1976) obtained the general flow distribution profile in the
d plate thickness, mm reverse and parallel manifold systems and concluded that the di-
D diameter, m ameter of two manifolds and the lateral tubes, the length of the
f Fanning frictional factor manifold, and the lateral flow resistance are the key influential
L length or width, m factors in this issue. Bassiouny and Martin (Bassiouny and Mar-
m ˙ mass flow rate, kg s − 1 tin, 1984a; Bassiouny and Martin, 1984b) extended Bajura and
p˙ momentum flow rate, kg m s − 2 Jones’ model into PHEs. They proposed a characteristic parameter
P pressure, Pa to determine the flow distribution profile in PHEs and also con-
Pc corrugation pitch, mm cluded that the area ratio between the intake and exhaust ducts
R radius, m needs to be correctly chosen to achieve a uniform distribution.
Re Reynolds number Wang et al. (J.Y. Wang et al., 2011) explained the difficulty of apply-
T temperature, °C ing the Bernoulli equation for the manifold or header is identify-
u velocity along the channel, m s − 1 ing a relative streamline to conserve energy and estimate the fric-
v axial velocity in the header, m s − 1 tional losses. The authors confirmed that the advantage of using
Q heat transfer rate, W the 1-D momentum conservation equation is to avoid the neces-
sity of knowing the detailed flow patterns. They also developed an
Greek Symbols
analytical model for the manifold pipe, which used non-constant
μ dynamic viscosity, Pa s
friction factor and pressure recovery factors. However, though the
ρ density, kg m − 3
correction factors were changing along the manifold, their val-
τ shear stress, Pa
ues were still coming from the experimental data. Further, Wang
ϕ chevron angle, o
(Wang, 2008; Wang, 2010) obtained the analytical solution of the
Subscripts flow distribution in the U- and Z-type manifold, which included
avg average both friction and inertial effects into consideration. The author also
b bulk concluded that the previous models developed in (Bassiouny and
c cold Martin, 1984a; Bassiouny and Martin, 1984b; Kee et al., 2002;
ch channel Maharudrayya et al., 2005) were the special cases of neglecting the
h hot/header friction or inertial effect.
i inlet/channel index The computational fluid dynamics (CFD) tool is also exten-
in inlet header sively used to study this issue. Muhana and Novog (Muhana and
o outlet Novog, 2008) applied the CFD tool to simulate the single-phase
out outlet header 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. (Tong et al.,
the distribution. In this paper, we will focus on the distribution in 2009) numerically examined eight geometric strategies of improv-
single-phase flow conditions for three reasons: (1) the single-phase ing the single-phase flow distribution in a generic manifold sys-
flow distribution in one side of the heat exchanger affects the two- tem with ten channels. A 2-D computational model was developed
phase flow distribution in the other side; (2) it is very relevant in (Tong et al., 2009), and four promising strategies were identi-
to a single-phase heat exchanger, but also to the two-phase heat fied from the modeling results. Moreover, their simulation also in-
exchangers like in the condensers where single-phase flow at the dicated that the sudden enlargement of the flow area at the man-
inlet characterizes the applications; (3) it is also very important ifold inlet, which is not uncommon in practice, generally reduces
at the exit of the direct-expansion evaporators. Eventually, in the the flow rate through the channels adjacent to the inlet. In some
papers that follow, we will present the case of the direct expansion extreme cases, the flow reversals occurred in those channels. This
(DX) evaporator. phenomenon was also observed by Wang et al. (C. Wang et al.,
Many studies used modeling approaches to analyze the single- 2011) in their 2-D CFD simulation of a compact parallel-flow heat
phase maldistribution issue in PHEs or other similar structures, exchanger with single-phase water flowing inside. Their calculation
which are also featured with the header and parallel channels ar- demonstrated that the jet flow was formed due to the sudden en-
rangement. Bajura and Jones (Bajura and Jones, 1976) developed a largement of the flow cross-sectional area at the heat exchanger
1-D analytical model for the flow distribution in the manifold sys- inlet, yielding a smaller pressure difference for the first several
tem. In this model, the authors applied the 1-D mass and momen- tubes near the inlet and the flow rate was hereby reduced. Due
tum conservation equations to the dividing and combining mani- to the complexity of the plate geometry, it is very computational
folds to correlate the flow rate and pressure profiles in the man- resource consuming to establish a full 3-D CFD model for a heat
ifolds. The governing equations of two manifolds were connected exchanger with many channels. Such an attempt was completed by
by the discharge equation, which described the relationship be- Tsai et al. (Tsai et al., 2009), who developed a 3-D CFD model for
tween the pressure differential between the manifolds and the lat- a PHE with two corrugated channels and simulated single-phase
eral flow rate. The 1-D assumption oversimplified the complex flow water flow inside. The computational results of (Tsai et al., 2009)
conditions in the manifolds, so the authors proposed two coeffi- showed that the maldistribution exists even in the heat exchanger
cients to correct. One coefficient corrected the non-uniform veloc- with only two channels, though the difference was relatively small
ity profile in the manifolds, and the other one accounted for the (1%). The authors also pointed out that 3-D modeling of PHEs with
axial momentum transfer to the lateral flow. These two correc- more channels requires more advanced development in computer
tion coefficients were assumed to be constant and selected based technology.

46
W. Li and P. Hrnjak International Journal of Refrigeration 126 (2021) 45–56

Fig. 1. Experimental facility and the pressure and temperature probe.

Experimental studies, addressing the issue of single-phase flow 2 CFD tool can capture the flow details in PHEs. However, the
distribution in PHEs or similar structures, are also reported in the complexity of plate geometry brings difficulties in the full 3-D
open literature. Bajura and Jones (Bajura and Jones, 1976) used simulation.
a PVC fabricated manifold system with air flowing through it to 3 Single-phase flow distribution in PHEs is determined by the
validate their analytical distribution model. Each header was in- header induced pressure drop and can be experimentally quan-
strumented with the pressure taps located midway between the tified by the pressure differential through channels.
branching tubes. The header pressure profiles and the differential
This paper presents an experimental and numerical investiga-
pressure between the headers were recorded and compared with
tion of single-phase flow distribution in BPHEs. In the experiments,
the model predictions. Tereda et al. (Tereda et al., 2007) experi-
the pressure profile in the heat exchanger is measured by the
mentally measured the pressure profile along the inlet and out-
probes inserted into the headers, and the flow distribution is de-
let header in a PHE with 25 channels. The pressure measurements
termined by the measured pressure drop across the channels and
were accomplished by inserting a mandrel with five axial equally
the developed in-channel friction factor correlation. We examined
spaced through holes on it into the headers. The flow rate through
four BPHEs with identical plate geometry but different numbers of
each channel is then calculated based on the measured differen-
plates (4, 10, 50, and 100 plates) and developed two distribution
tial pressure and the single-phase friction factor correlation. Bob-
models based on the principle of equal total pressure drop for all
bili et al. (Bobbili et al., 2006) compared the hydraulic performance
flow paths. One model calculates the pressure drop in the headers
and flow distribution in two PHEs with 21 and 81 plates. In their
by 3-D CFD modeling while the other finds the pressure profile by
tests, a mobile static pressure probe was placed into the header
1-D mass and momentum conservation equations. The models are
of the heat exchangers to measure the pressure drop of different
validated by the experimental results.
channels. Their experimental data indicated that the maldistribu-
tion was more severe in the heat exchanger with more plates.
Bobbili et al. (Bobbili et al., 2006) also verified the theoretical 2. Experimental methods and results
model proposed by Bassiouny and Martin (Bassiouny and Martin,
1984a) was capable of predicting the flow distribution reasonably 2.1. The facility
well.
The following conclusions can be drawn from the above litera- The experimental facility is shown in Fig. 1. It consists of two
ture review: 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.
1 When applying 1-D mass and momentum conservation equa- Two centrifugal pumps, controlled by the variable frequency drives,
tion in headers to develop the distribution model, three effects are used to circulate water for two loops. Two immersion electrical
need to be corrected: (a) non-uniform axial velocity profile in heaters are in the tank for heating of the hot water loop; the tap
headers; (b) axial momentum transfer into the channels; (c) water provides cooling for the cold-water loop. Another pump is
readjustment of the velocity profile near the entrance of heat used to circulate the water in the tank to reduce the temperature
exchangers. stratification inside.

47
W. Li and P. Hrnjak International Journal of Refrigeration 126 (2021) 45–56

Table 1 Table 2
Measurement uncertainty. Plate geometry of BPHEs.

Measured Parameter Uncertainty Parameters Value

Temperature (T type) 0.1 ( °C) Chevron angle, ϕ , o 60


Absolute pressure (0 - 172 kPa) 0.11% (full scale) Corrugation depth, b, mm 2.0
Differential pressure (0 - 7.5 kPa) 0.25% (full scale) Corrugation pitch, Pc , mm 7.0
Differential pressure (0 - 6.9 kPa) 0.25% (full scale) Plate thickness, d, mm 0.30
Mass flow rate 0.15% (reading) Port length, Lp , mm 172
Total length, Lv , mm 206
Calculated Parameter Uncertainty Port width, Lh , mm 42
Total width, Lw , mm 76
Friction factor < 7.0% Heat transfer area per plate, Ap , m2 0.014
Port diameter, Dp , mm 20.0
Feeding tube diameter, Df , mm 16.0

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 °C, and 30 °C. Identical average channel mass flux is sup-
plied to both streams. For each test, the primary measurements of
each flow stream are the total mass flow rate, the inlet/outlet tem-
perature of the BPHEs, and the temperature change /pressure drop
across different channels.

2.2. Data reduction

For the data reduction, fluid properties are calculated at the


bulk-mean temperature given by:
Th,i + Th,o
Th,b = ; Tc,b = Th,b − LMT D (1)
2
Fig. 2. Schematic of the plate in BPHE.
Where LMTD is calculated by:
      
Coriolis type flow meters, absolute and differential pressure
LMT D = Th,i − Tc,o − Th,o − Tc,i /ln Th,i − Tc,o / Th,o − Tc,i
transducers, and type T (copper-constantan) thermocouples are in- (2)
stalled at locations as indicated in Fig. 1. The sensors have been
The heat transfer rate is obtained from the average of hot side
calibrated. Their range and uncertainty are listed in Table 1. The
and cold side energy balance:
static mixers are installed at the entrance and exit of the tested
 
BPHE to guarantee a uniform temperature measurement by the Qh = m˙ h · c p · Th,i − Th,o ;
thermocouples. Qc = m˙ c · c p · (Tc,o − Tc,i ); (3)
Moreover, four temperature and pressure probes are placed into Qavg = (Qh + Qc )/2
the inlet/outlet headers of the BPHEs to measure the temperature
change and pressure drop across the channels. The structure of the The measured channel pressure drop in the 4-plate BPHE,
probe and its relative position in the tested BPHE is also shown in which has only one hot channel in the middle of two cold chan-
Fig. 1. The probes are designed as inserting a lengthened thermo- nels, is used to calculate the Fanning friction factor for a single
couple (outer diameter: 1.59 mm) into a stainless-steel tube (outer channel and develop the single-phase water friction factor correla-
diameter: 3.18 mm, inner diameter: 2.36 mm). The stainless-steel tion. The Fanning friction factor for a single channel is calculated
tube is blocked at the end, and a through-hole is drilled on the through Eq. (4):
side surface close to the end to measure the static pressure in ρ · De · Pch,fri
the header. The tip of the thermocouple is bent into the channel f = 2
(4)
through the open hole of the stainless-steel tube, to record the in- 2 · L p · (m˙ ch /Acs )
let/outlet temperature of a channel. By moving the probes forward In Eq. (4), De = 2 · b is the channel effective diameter, and Acs =
and backward, the temperature change and pressure drop across b · Lw is the flow area of the channel.
different channels could be measured. Then, the power-law equation, given in Eq. (5), is used to cor-
All the measurements are collected by an HP 75,0 0 0 series B relate the friction factor, f , with the channel Reynolds number, Re,
data logger. The control and monitoring of the data are accom- defined in Eq. (6). The constants C1 and C2 are obtained by curve-
plished by VEE software. fitting. The developed friction factor correlation would be used to
The tested BPHEs are well insulated, with the heat loss cali- determine the flow distribution among the channels in the BPHEs
brated so that the energy balance discrepancy (measured heat load with 10, 50, and 100 plates.
difference between hot and cold stream) is within ± 4%.
The BPHEs with identical plate geometry but different numbers f = C1 /ReC2 (5)
of plates are tested in the experiments. They all have the 1pass-
1pass, U-type configuration. The schematic of the corrugated plate m˙ ch · De
Re = (6)
and its geometric parameters are given in Fig. 2 and Table 2. It can μ · Acs
48
W. Li and P. Hrnjak International Journal of Refrigeration 126 (2021) 45–56

Fig. 4. Pressure profiles and flow distribution in the 10-plate BPHE.

Fig. 3. BPHE friction factor.

2.3. Single-phase water friction factor

Fig. 3 gives the measured friction factor of single-phase wa-


ter flow in a single channel of the tested BPHEs. For compari-
son, Fig. 3 also shows the friction factor data from Jin and Hrn-
jak (Jin and Hrnjak, 2017), in which a different plate geometry
was tested. The friction factor correlation proposed in Focke et al.
(Focke et al., 1985) is also plotted in the figure. As shown in Fig. 3,
the friction factor data in Jin and Hrnjak (Jin and Hrnjak, 2017)
present a clear transition from laminar to turbulent flow around
the Re of 50 ~ 60. Shah and Sekulic (Shah and Sekulic, 2002) 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 re-
gion in Jin and Hrnjak (Jin and Hrnjak, 2017). Therefore, the tested
results of this study are in the early turbulent regime. The fric- Fig. 5. Pressure profiles and flow distribution in the 50-plate BPHE.
tion factor correlation from Focke et al. (Focke et al., 1985) gives
a more quantitatively similar result to the measured data in this
study. The discrepancy in value may be attributed to the difference For each BPHE, three different average channel mass flow rates are
in the plate geometry. tested: 5, 7.5, and 10 g·s − 1 (corresponding channel Re ≈ 200 ~
A friction factor correlation is curve-fitted by the measured data 400). To compare the distributions at different flow rates, the chan-
in Fig. 3 as the form of Eq. (5): nel mass flow rates are plotted as the percentage of the average
channel flow rate:
f = (4.33 ± 0.1317)/Re(0.109±0.0051) (7)
m˙ ch
Instead of a general correlation, Eq. (7) is only applicable to the m˙ ∗ch = (8)
m˙ ch,avg
specific plate geometry used in this study under the tested condi-
tions (Re=180~1500). As shown in Fig. 4, the pressure change along the inlet and out-
let headers of the 10-plate BPHE is almost negligible compared to
2.4. Single-phase distribution in BPHEs the pressure drop across the channels. This is because the tested
10-plate BPHE has a small number of channels, a short inlet/outlet
The water mass flow rate distribution in the BPHEs is obtained header, and a low total mass flow rate. Constant pressure in the
based on the measured pressure drop across the channels and the inlet and outlet header gives an almost uniform flow rate distribu-
developed in-channel friction factor correlation (Eq. (7)). By adopt- tion among the channels. Moreover, different average channel mass
ing this approach, the summation of the measured flow rate for all flow rates seem to only change the absolute pressure drop across
channels (linear interpolation for the channels which are not mea- the channels, while the shapes of the pressure profile in the head-
sured directly) is within a difference of 5% compared to the total ers are not affected. The distribution profiles under three flow rates
flow rate. To further reduce the error, a correction factor is applied are almost overlapping.
to the measured channel flow rate, such that their summation is In the 50-plate BPHE (Fig. 5), there is a significant pressure
equal to the total water mass flow rate. The pressure profiles along change in the headers. Following the flow direction, the static pres-
the inlet and outlet headers, and the mass flow rate distribution of sure increases in the inlet header. This is because the flow rate
the cold stream in the BPHEs are shown in Fig. 4 to 6. The pres- in the inlet header decreases, which converts the dynamic pres-
sure profiles are given by the difference between the pressure in sure to the static pressure. In the outlet header, the flow accel-
the inlet/outlet header and that at the heat exchanger exit (Pexit ). erates, and the static pressure is converted to the dynamic pres-

49
W. Li and P. Hrnjak International Journal of Refrigeration 126 (2021) 45–56

sure, 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 dis-
tance away from the entrance/exit of the BPHE. This is mainly be-
cause, in the inlet header, a portion of axial momentum is trans-
ferred into the channels; thus, the increase in the static pressure
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
Fig. 6. Pressure profiles and flow distribution in the 100-plate BPHE.
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 the plates, a small portion of the metal plate near the port, along
less mass flow through the channels at the rear part. Besides, the with the flow space in between, is included in the computational
pressure loss due to the shear force in the headers further reduces domain (Fig. 7) to consider the possible interaction between the
the pressure change in the inlet header and increases that in the header flow and the metal plates. The feeding tube connected to
outlet header. But the contribution of the shear force is not signif- the heat exchanger is also added. The generated meshes are struc-
icant in BPHEs since the header flow only touches the edge of the tured, consisting of only hexahedral cells. The grid independence
port of each metal plate. tests have been conducted. Three different meshes with 370,355,
Also, one could find in Fig. 5 the three non-dimensional distri- 1,676,241, 2,242,976 nodes are built for the inlet header with 49
bution profiles at different flow rates are almost not distinguish- channels (100-plate BPHE, hot side). Using these three meshes, the
able. This is because, with a higher average mass flow rate, the simulations are carried out at m˙ ch,avg =5 g s − 1 . The pressure pro-
pressure drop in the headers and that across the channels simul- file in the header is used as the criterion since it is crucial to
taneously increase. These two effects compensate for each other, the flow distribution. The results show that the calculated pres-
making the non-dimensional flow distribution unchanged. Simi- sure profiles by these three meshes are almost the same, which
lar results were reported in other researches (Maharudrayya et al., indicates the meshes are sufficiently fine for the tested case. To
2005; Wang, 2011). Therefore, for all tested average channel flow capture more flow details, the mesh with the largest number of
rates, the maximum channel flow rate is achieved in the 5th chan- nodes is selected. For the headers with other numbers of channels,
nel, and the minimum flow rate is in the last channel, and the dif- the same max cell size is kept in the mesh generations.
ference between them is around 15% of the average channel flow The generated mesh is then inputted to the Fluent for the
rate. flow simulation. In Fluent, the steady-state conservation equations
Fig. 6 shows the experimental results in the 100-plate BPHE. for the mass and momentum are solved using the finite volume
Similar trends can be observed in the figure as the case of the 50- method. The SIMPLEC scheme is used to solve for the pressure-
plate BPHE. In the 100-plate BPHE, the pressure change in the inlet velocity coupling. The turbulence simulation is accomplished by
and outlet header is much more significant than that in the 50- the realizable k-ε model, which uses two transport equations, one
plate BPHE due to the higher total mass flow rate. The channel is for the turbulent kinetic energy, k, and the other for the rate of
flow rate is also more mal-distributed in the 100-plate BPHE. The dissipation of turbulent kinetic energy, ε , to describe turbulence.
difference between the maximum and minimum channel flow rate In the calculations, the pressure inlet/outlet boundary condi-
is around 66% of the average channel flow rate. tions are set for the inlet/outlet of the heat exchanger. The veloc-
ity inlet condition is used for the channel inlets/outlets, in which
3. CFD-linked distribution model the fluid is assumed to flow in the vertical direction when enter-
ing/exiting the channels. The velocity of the fluid entering/exiting
A CFD-linked distribution model is developed to predict the the channels is calculated based on the mass flow rate. Since the
flow rate distribution in the BPHEs. In this model, a compromise is mass flow rate for each channel is unknown before the simulation,
made between the complexity and accuracy of the simulation. The therefore, the iteration is needed to estimate a reasonable flow rate
model adopts the 3-D CFD simulation for the flow in the headers distribution.
and connects two header CFD simulations by the in-channel pres-
sure drop calculations (1-D). A similar approach was adopted in 3.2. Equal total pressure drop for all flow paths
Huang et al. (Huang et al., 2014) in microchannel heat exchangers.
In a BPHE, any flow path, starting from the heat exchanger
3.1. CFD simulation of flow in the headers entrance, passing through the inlet header, plate channel, outlet
header, and eventually ending at the exit should have the same
The BPHE headers are simulated with a commercial CFD code total pressure drop. The proposed CFD-linked model predicts the
Fluent. The 3-D meshes are generated for the headers in Gambit flow rate distribution by iteratively adjusting the flow rate through
software. Since the BPHE header consists of a series of ports on each channel to achieve an equal total pressure drop for all flow

50
W. Li and P. Hrnjak International Journal of Refrigeration 126 (2021) 45–56

Fig. 7. Mesh generation and boundary conditions for the header CFD models.

Fig. 8. Flow paths and total pressure drop calculation in BPHEs.

paths. Similar approaches were adopted by many works related to


the flow distribution issues (Tuo et al., 2012; Li and Hrnjak, 2021).
As shown in Fig. 8, for each flow path, the total pressure drop
is divided into three components: the pressure drop in the in-
let/outlet header and that across the plate channel:

i 
i
Ppath,i = Pin,i + Pch,i + Pout,i (9)
1 1

In Eq. (9), the pressure drop in the inlet/outlet header is ob-


tained from the CFD simulation, and the pressure drop across
− 1
the channel includes: the local pressure drop of turning-in and Fig. 9. Axial velocity and pressure profile in the headers (m˙ ch,avg =10 g s , 50-
turning-out, frictional and gravitational pressure drop: plate BPHE, cold side).

Pch,i = Pturn−in,i + Pf ri,i + Pgra,i + Pt urn−out ,i (10)


The frictional pressure drop is calculated by the previously de- axial velocity and the pressure profile of the center plane in the
veloped friction factor correlation (Eq. (7)), and the local pressure inlet and outlet header of the BPHE. The pressure profile is given
loss of turning-in and turning-out are calculated by the correla- by the pressure difference to the header inlet/outlet.
tions in Idel’chik (Idel’chik, 1994) (Section VII, Diagram 7–21; Sec- The axial velocity contour, as well as the local streamlines, in
tion VII, Diagram 7–7). the inlet header, shows that there is a clear sudden expansion flow
near the entrance of the heat exchanger, induced by the size differ-
3.3. Flow details in the headers of the BPHE ence between the feeding tube and the header. The readjustment
of the velocity profile takes a distance of several channels. Due to
The CFD-linked distribution model predicts the single-phase this sudden expansion flow, the eddies are generated near the in-
flow distribution in the BPHE and simultaneously simulates the let of the first several channels, and the axial velocity in this region
flow field in the headers. The comparison between the pre- is negative, as demonstrated by the streamlines. For the channels
dicted and measured flow rate distribution will be presented in after the sudden expansion region, the axial velocity at the chan-
Section 4 of this paper. Fig. 9 depicts the typical contours of the nel inlet is positive, which means the axial momentum is trans-

51
W. Li and P. Hrnjak International Journal of Refrigeration 126 (2021) 45–56

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 mech-
anistically 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 recov-
ered in the header flow.
To implement this method, an initial velocity profile at the heat
exchanger entrance is first assumed. For the fully developed turbu-
lent flow, the 1/7th power-law velocity profile is used (Eq. (13)). In
Fig. 10. Control volume in the inlet header. Eq. (13), the origin of the Cartesian coordinate is at the center of
the cross-section.
 1/7
ferred to the channels. It is also observed from the streamlines x2 + y2
that the fluid near the channel inlets branches out first and the
vin,1 = 1.22 · v̄in,1 · 1 − (13)
R
fluid closer to the center flows toward the boundary of the header
and branches through downstream channels. The sudden expan- Fig. 11 demonstrates the procedure of estimating the down-
sion near the entrance also leads to an abrupt pressure rise in this stream velocity profile based on the upstream information. For
region, as shown in the pressure profile. At the downstream of the generality, the velocity profile at the ith cross-section is assumed
expansion region, the pressure increase, caused by the conversion to be known. The momentum flow rate at this cross-section is cal-
between the dynamic and static pressure, is much more moderate. culated by the numerical integration:
The velocity profile in the outlet header, also given in Fig. 9,  x2 +y2 =R2
is more uniform than that in the inlet header. The vertical flow p˙ i = ρ · v2in,i · dxdy (14)
in the channels, without too much axial momentum, joins the ax- 0

ial flow in the header, which makes the velocity of the fluid near It is convenient to assume the branching is mainly in the up-
the channel outlets lower than that away from the channel outlets. ward direction (+y direction), and thus the fluid at the top of the
The fluid also experiences a higher pressure change in the outlet header tends to branch out first. An elliptical boundary is drawn on
header than that in the inlet header (the sudden expansion region the velocity profile, as shown in Fig. 11, with the assumption that
excluded) due to the contribution of the axial momentum transfer the flow above this boundary branches out through the ith chan-
and the shear force. nel and that below this boundary moves forward in the header.
The mass conservation requires the flow rate above the elliptical
4. Mechanistic distribution model boundary equal to the presumed channel flow rate, m˙ ch,i :
 x2 +y2 =R2 ,y>0
As discussed in the literature review, three effects need to be m˙ ch,i = 2
ρ · vin,i · dxdy (15)
corrected when developing a 1-D analytical distribution model. x2 + y2 =R2 ,y>0
r
i
The usual practice in the literature is applying empirical correc-
tion factors in the momentum conservation equation. Therefore, ri in the above equation is called the elliptical ratio, which deter-
the generality of the models is limited. mines the shape of the elliptical boundary. Since in the calculation
In this section, a mechanistic distribution model is developed, of the pressure profile, the channel flow rate m˙ ch,i , is treated as
which accounts for these three effects mechanistically. This mech- a known variable, the elliptical ratio ri , can be solved numerically
anistic model also predicts the distribution by imposing the con- from the Eq. (15). The axial momentum flow rate contained in the
dition of equal total pressure drop for all flow paths. Instead of branching-out is calculated by:
calculating the header pressure profile by the CFD tool, the mecha-  x2 +y2 =R2 ,y>0

nistic model estimates the pressure change in the headers by mass p˙ ch,i = 2
ρ · v2in,i · dxdy (16)
x2 + y2 =R2 ,y>0
and momentum conservation equations. r
i

With the fluid above the elliptical boundary branching out, the
4.1. Pressure change in the inlet header remaining velocity profile below the boundary expands to fill the
entire cross-section of the header and form a new velocity profile
The mass and momentum conservation equations are applied at the (i + 1)th cross-section. With the elliptical boundary, the ex-
to a control volume in the inlet header, as shown in Fig. 10: pansion of the velocity profile is linear in the vertical direction. In
  
other words, to fill the entire cross-section, the remaining veloc-
ρvin,i dAh = ρvin,i+1 dAh + ρ uch,i dAch (11) ity profile only needs to be stretched in the vertical direction by
a ratio of 1+2r , as shown in Fig. 11. The relationship between two
  i
corresponding points on the velocity profiles before and after the
Pin,i Ah − Pin,i+1 Ah − τ · Ashear = ρv 2
in,i+1 d Ah − ρv 2
in,i d Ah stretch is:

1 − ri 2yi
+ ρ uch,i vch,i dAch (12) yi+1 = R2 − x2i · + ; (17)
1 + ri 1 + ri
In the momentum conservation Eq. (Eq. (12)), the left-hand side In Eq. (17), (xi , yi ) is a point on the velocity profile at ith cross-
is the axial force exerting on the control volume, and the right- section; (xi+1 , yi+1 ) is a corresponding point on the velocity profile
hand side is the axial momentum flow rate leaving the control vol- at (i + 1)th cross-section. With the mass conservation. the velocity
ume. at the (i + 1)th cross-section is given by:
It should be mentioned that in BPHEs, the flow in the header ( ri + 1 )
only touches the edge of the metal plates, as Fig. 10 shows. The vin,i+1 (xi+1 , yi+1 ) = · vin,i (xi , yi ) (18)
2
52
W. Li and P. Hrnjak International Journal of Refrigeration 126 (2021) 45–56

Fig. 11. Estimation of the velocity profile.

Fig. 13. Control volume in the outlet header.


Fig. 12. Control volume in the sudden expansion region.

the momentum flow rate transfer through the channel is estimated


The momentum flow rate at (i + 1)th cross-section is: by:
 x2 +y2 =R2 p˙ ch,i = m˙ ch,i · vch,i (21)
p˙ i+1 = ρ · v2in,i+1 · dxdy (19)
0 With: v̄ch,i ≈ −0.1 · v̄in,i (Devenport and Sutton, 1993).
The pressure difference between the and (i + ith
cross- 1)th When estimating the velocity profile at the (i + 1)th cross-
section is then calculated by the momentum conservation equa- section in the sudden expansion region, in addition to the stretch,
tion: the expansion of the core flow also needs to be considered.
Eq. (17) is modified as:
Pin,i Ah − Pin,i+1 Ah = p˙ i+1 − p˙ i + p˙ ch,i + τ · Ashear (20) Ri+1
xi+1 = xi · Ri
;
This mechanistic model also takes the sudden expansion flow 1−ri 2yi Ri+1 (22)
yi+1 = R2 − x2i · 1+ri
+ 1+ri
· Ri
;
at the heat exchanger entrance into consideration. For the flow
in an abruptly expanding duct, the turbulent reattachment length,
The velocity profile at the (i + 1)th cross-section is given by:
which is the distance between the expansion and reattachment
 2
location, is about 7~9 step heights (Morrison et al., 1988). In ( ri + 1 ) Ri
BPHEs, one step height is the radius difference between the feed- vin,i+1 (xi+1 , yi+1 ) = · vin,i (xi , yi ) · (23)
2 Ri+1
ing tube and the header. The reattachment should be easier with
the fluid branching out through the channels. Therefore, the expan- The axial momentum flow rate at the (i + 1)th cross-section is
sion length in this model is selected to be the lower bound, 7 step then calculated by Eq. (19) after replacing the radius of the header,
heights. For the BPHEs used in this study, it is about 4 channels. R, by the radius of the core flow, Ri+1 . The corresponding pressure
For other BPHEs, it may not be 4 channels, but can be calculated difference between the ith and (i + 1)th cross-section is given by:
by the size of the feeding tube and the header. To simulate the
readjustment of the velocity profile, the diameter of the core flow Pin,i Ah − Pin,i+1 Ah = p˙ i+1 − p˙ i + p˙ ch,i (24)
is assumed to expand linearly from the diameter of the entrance to
that of the header over the expansion length. The control volume 4.2. Pressure change in the outlet header
in the sudden expansion region is depicted in Fig. 12.
With this assumption, when calculating the axial momentum The control volume for the outlet header is shown in Fig. 13.
flow rate at the ith cross-section by Eq. (14), and the elliptical ratio Because the flow direction in the channels has been regulated by
by Eq. (15), the radius of the header, R, needs to be replaced by the metal plates to be vertical, there is almost no axial momen-
the radius of the core flow, Ri . tum entering the outlet header. Besides, though there is a sudden
As shown in the CFD simulation (Fig. 9), in the sudden ex- contraction of the flow area at the exit of the heat exchanger, the
pansion region, eddies are generated near the inlet of the chan- influence of this flow sudden contraction on the outlet header flow
nels, and the fluid enters the channels with negative axial velocity. is quite limited for the studied cases, as shown in the CFD results
Therefore, Eq. (16) is not applicable in the expansion region, and (Fig. 9). Moreover, there is no initial velocity profile for the flow in

53
W. Li and P. Hrnjak International Journal of Refrigeration 126 (2021) 45–56

the outlet header. Therefore, the flows in the outlet header should βout in the above equation is to compensate for the non-
have more similarities, even with various geometries and condi- uniformity of the axial velocity profile in the outlet header. βout ≈
tions. For simplicity, the 1-D flow approximation is made for the 1.30 is suggested to be reasonably suitable for most cases (Bajura
outlet header, and the averaged velocity is used to calculate the and Jones, 1976; Bajura, 1971).
pressure change: With the pressure change in the headers calculated by the
ρ Ah v̄out,i = ρ Ah v̄out,i+1 + m˙ ch,i (25) 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.
Pout,i Ah − Pout,i+1 Ah = βout · ρ Ah v̄2out,i+1 − βout · ρ Ah v̄2out,i − τ · Ashear
(26)

Fig. 14. Model predictions vs. experiments. (For interpretation of the references to color in this figure legend, the reader is referred to the web version of this article.)

54
W. Li and P. Hrnjak International Journal of Refrigeration 126 (2021) 45–56

− 1
Fig. 15. Axial velocity profiles in the inlet header (50-plate BPHE, m˙ ch,avg =10 g s ).

5. Model validations 2 In the BPHEs, the influence of the average channel flow rate on
the distribution profile is trivial, but the maldistribution is more
The predictions of two newly proposed models are plotted and severe as the number of plates increases.
compared with the experimental results in Fig. 14. The modeling 3 The proposed CFD-linked model simulates the flow in the head-
results agree well with the experimental measurements. The de- ers by the 3-D CFD modeling and calculates the pressure drop
viations of the predictions by the mechanistic model are slightly across the channel by the 1-D correlations. The flow distribu-
larger than those by the CFD-linked model due to more ideal- tion is predicted by imposing the condition of equal total pres-
ized simplifications used in the simulations. The predictions for the sure drop for all flow paths.
100-plate BPHE are less accurate than those for the 50-plate BPHE 4 The mechanistic model uses 1-D mass and momentum conser-
since the flow in the headers is more complicated with a higher vation equations to calculate the pressure change in the head-
mass flow rate in a larger heat exchanger. Models always underes- ers. The effects of the non-uniform velocity profile, sudden
timate the flow rate through the first several channels, which in- expansion at the entrance, and axial momentum transfer are
dicates that further modification should focus on the simulation of mechanistically considered by tracking the evolution of the ax-
the sudden expansion of the flow at the heat exchanger entrance. ial velocity profile in the inlet header.
The highlight of the proposed mechanistic model is analytically 5 Compared with the experimental results, the predictions by
tracking the evolution of the axial velocity profiles in the inlet the two models demonstrate a reasonably good agreement. The
header. Fig. 15 compares the velocity profiles in the inlet header CFD-linked model has slightly higher accuracy, but the mech-
of a specific case, obtained by the mechanistic model and by the anistic model has the advantage of easy application to various
CFD tool. It can be seen in Fig. 15, the mechanistic model is ca- heat exchanger geometry and operating conditions.
pable of simulating the “upward stretch” trend in the evolution of
the velocity profile. The basic shapes of the velocity contours given Declaration of Competing Interest
by the two methods are similar. However, one could also observe
that some flow physics are not captured by the mechanistic model, The authors declare that they have no known competing finan-
like the recirculating flow at the bottom of the header. In conclu- cial interests or personal relationships that could have appeared to
sion, the mechanistic model has the advantage of easy application influence the work reported in this paper.
to various heat exchanger geometry and operating conditions while
scarifying the accuracy due to the 1-D simplifications.
Acknowledgments

Summary and conclusions The authors thankfully acknowledge the support provided by
the Air Conditioning and Refrigeration Center at the University of
The single-phase flow distribution in the BPHEs is investigated Illinois at Urbana-Champaign and the Creative Thermal Solutions,
experimentally and numerically. The pressure profile in the BPHEs Inc.
is measured by the probes inserted into the headers, and the flow
distribution is determined by the measured pressure drop across References
the channels and the developed in-channel friction factor correla-
Bajura, R.A., 1971. A model for flow distribution in manifolds. J. Eng. Power 93 (1),
tion. Two distribution models are developed based on the principle 7–12.
of equal total pressure drop for all flow paths. Bajura, R.A., Jones Jr, E.H., 1976. Flow distribution manifolds. J. Fluids Eng. Trans.
The main results in this study are summarized as the following: ASME 98 (4), 654–665.
Bassiouny, M.K., Martin, H., 1984a. Flow distribution and pressure drop in plate heat
exchangers-I: u-type arrangement. Chem. Eng. Sci. 39 (4), 693–700.
1 Along the flow direction in the inlet header, the channel flow Bassiouny, M.K., Martin, H., 1984b. Flow distribution and pressure drop in plate heat
rate first increases for the first several channels due to the sud- exchangers-II: z-type arrangement. Chem. Eng. Sci. 39 (4), 701–704.
Bobbili, P.R., Sunden, B., Das, S.K., 2006. An experimental investigation of the port
den expansion of the flow area at the heat exchanger entrance.
flow maldistribution in small and large plate package heat exchangers. Appl.
For the rest channels, the flow rate generally decreases with the Therm. Eng. 26, 1919–1926.
distance away from the entrance/exit of the BPHE. This is be- Devenport, W.J., Sutton, E.P., 1993. An experimental study of two flows through an
cause the inlet header has a relatively smaller pressure change, axisymmetric sudden expansion. Exp. Fluids 14, 423–432.
Focke, W.W., Zachariades, J., Olivier, I., 1985. The effect of the corrugation inclination
which is associated with the axial momentum transfer to the angle on the thermohydraulic performance of plate heat exchangers. Int. J. Heat
channels. Mass Transf. 28 (8), 1469–1479.

55
W. Li and P. Hrnjak International Journal of Refrigeration 126 (2021) 45–56

Huang, L., Lee, M.S., Saleh, K., Aute, V., Radermacher, R., 2014. A computational Shah, R.K., Sekulic, D.P., 2002. Fundamentals of Heat Transfer Design. Wiley first ed.
fluid dynamics and effectiveness-NTU based co-simulation approach for flow Tereda, F.A., Srihari, N., Das, S.K., Sunden, B., 2007. Experimental investigation on
mal-distribution analysis in microchannel heat exchanger headers. Appl. Therm. port-to-channel flow maldistribution in plate heat exchangers. Heat Transf. Eng.
Eng. 65, 447–457. 28 (5), 435–443.
Idel’chik, I.E., 1994. Handbook of Hydraulic Resistance. CRC Press. Tong, J.C.K., Sparrow, E.M., Abraham, J.P., 2009. Geometric strategies for attainment
Jin, S., Hrnjak, P., 2017. Effect of end plates on heat transfer of plate heat exchanger. of identical outflows through all of the exit ports of a distribution manifold in
Int. J. Heat Mass Transf. 108, 740–748. a manifold system. Appl. Therm. Eng. 29, 3552–3560.
Kee, R.J., Korada, P., Walters, K., Pavol, M., 2002. A generalized model of the flow Tsai, Y., Liu, F., P, Shen, 2009. Investigations of the pressure drop and flow distri-
distribution in channel networks of planar fuel cells. J. Power Sources 109, bution in a chevron-type plate heat exchanger. Int. Commun. Heat Mass Transf.
148–159. 36, 574–578.
Li, J., Hrnjak, P., 2021. An experimentally validated model for microchan- Tuo, H., Bielskus, A., Hrnjak, P., 2012. Experimentally validated model of refrigerant
nel condensers with separation circuitry. Appl. Therm. Eng. 183 (Part I) distribution in a parallel microchannel evaporator. SAE Int. J. Mater. Manf. 5 (2),
https://doi.org/10.1016/j.applthermaleng.2020.116114. 365–374.
Maharudrayya, S., Jayanti, S., Deshpande, A.P., 2005. Flow distribution and pressure Wang, C., Yang, K., Tsai, J., Chen, I.Y., 2011a. Characteristics of flow distribution in
drop in parallel-channel configurations of planar fuel cells. J. Power Sources 144, compact parallel flow heat exchangers, part I: typical inlet header. Appl. Therm.
94–106. Eng. 31, 3226–3234.
Morrison, G.L., Tatterson, G.B., Long, M.W., 1988. Three-dimensional laser velocime- Wang, J., 2011. Theory of flow distribution in manifolds. Chem. Eng. J. 168,
ter investigation of turbulent, incompressible flow in an axisymmetric sudden 1331–1345.
expansion. J. Propul. Power 4 (6), 533–540. Wang, J.Y., 2008. Pressure drop and flow distribution in parallel-channel of config-
Muhana, A., Novog, D.R., 2008. Validation of Fluent for prediction of flow distribu- urations of fuel cell stacks: u-type arrangement. Int. J. Hydrog. Energy 33 (21),
tion and pressure gradients in a multi-branch header under low flow conditions. 6339–6350.
In: Proceedings of the 16th International Conference on Nuclear Engineering. Wang, J.Y., 2010. Pressure drop and flow distribution in parallel-channel of con-
Orlando paper 48128. figurations of fuel cell stacks: z-type arrangement. Int. J. Hydrog. Energy 35,
Rao, B.P., Sunden, B., Das, S.K., 2005. An experimental and theoretical investigation 5498–5550.
of the effect of flow maldistribution on the thermal performance of plate heat Wang, J.Y., Gao, Z.L., Gan, G.H., Wu, D.D., 2011b. Analytical solution of flow coeffi-
exchangers. J. Heat Transf. 127 (3), 332–343. cients for a uniformly distributed porous channel. Chem. Eng. J. 84 (1), 1–6.

56

You might also like