You are on page 1of 21

Experimental Heat Transfer

A Journal of Thermal Energy Generation, Transport, Storage, and


Conversion

ISSN: (Print) (Online) Journal homepage: www.tandfonline.com/journals/ueht20

Visualization of direct contact heat transfer


process driven by continuous low-temperature
heat source and its performance characterization

Xihao Gao, Biao Li, Hui Sun, Jianxin Xu & Hua Wang

To cite this article: Xihao Gao, Biao Li, Hui Sun, Jianxin Xu & Hua Wang (2024) Visualization
of direct contact heat transfer process driven by continuous low-temperature heat source
and its performance characterization, Experimental Heat Transfer, 37:2, 142-161, DOI:
10.1080/08916152.2022.2105987

To link to this article: https://doi.org/10.1080/08916152.2022.2105987

Published online: 31 Jul 2022.

Submit your article to this journal

Article views: 107

View related articles

View Crossmark data

Citing articles: 4 View citing articles

Full Terms & Conditions of access and use can be found at


https://www.tandfonline.com/action/journalInformation?journalCode=ueht20
EXPERIMENTAL HEAT TRANSFER
2024, VOL. 37, NO. 2, 142–161
https://doi.org/10.1080/08916152.2022.2105987

Visualization of direct contact heat transfer process driven by


continuous low-temperature heat source and its performance
characterization
Xihao Gaoa,b,c, Biao Lia,b,c, Hui Suna,b,c,d, Jianxin Xua,b,c, and Hua Wanga,b,c
a
Engineering Research Center of Metallurgical Energy Conservation and Emission Reduction, Ministry of Education,
Kunming University of Science and Technology, Kunming, PR China; bNational Joint Engineering Research Center of
Energy Saving and Environmental Protection Technology in Metallurgy and Chemical Engineering Industry, Kunming
University of Science and Technology, Kunming, PR China; cFaculty of Metallurgical and Energy Engineering, Kunming
University of Science and Technology, Kunming, PR China; dFaculty of Science, Kunming University of Science and
Technology, Kunming, PR China

ABSTRACT ARTICLE HISTORY


In this study, the uniformity factor of the temperature difference field was Received 9 November 2021
introduced into the direct contact heat exchanger by establishing a trace Accepted 19 July 2022
element layer model for the heat exchanger. Furthermore, the uniformities of KEYWORDS
the temperature and temperature difference field of the direct contact heat Direct contact heat
exchanger operated using a continuous heat source were analyzed. The heat exchanger; uniformity of the
exchange height was proposed,and its effects on the volumetric heat trans­ temperature difference field;
fer coefficient and heat-exchanger efficiency were analyzed. The experimen­ heat exchange height;
tal results were consistent with the uniformity principle of the temperature efficiency
difference field.

Introduction
A direct contact heat exchanger (DCHE) is also known as a hybrid heat exchanger. Hot and cold fluids
in direct contact undergo heat exchange, which typically occurs between two fluids – one in the gas
phase and the other with a lower vaporization pressure in the liquid phase – that are easily separated
after heat exchange. For example, heat and mass transfer occurs between hot water and air that are in
direct contact to cool water in a water cooling tower. DCHEs are widely used in space heating,
refrigeration, power stations, chemical plants, and other fields [1]. Compared with other heat
exchangers, DCHEs have no internal inter-wall surfaces, larger heat-transfer areas, and negligible
heat transfer thermal resistance and can be operated using low-temperature heat sources, among other
advantages [2]. A large number of studies have been conducted on the intense heat transfer perfor­
mance of DCHEs [3–10]. Adding nanoparticles into a fluid to eliminate the flow boundary layer is an
effective method for improving the heat transfer efficiency [3, 6]. Furthermore, heat transfer between
fluids has been simulated and the heat transfer model, which is suitable for DCHEs, was found useful
for improving heat transfer [4, 5].
Heat transfer between fluids occurs at the interface and can be improved by increasing the
uniformity of two fluids through mixing [10–15]. Xiao et al. discussed the distribution of bubble
populations in DCHEs theoretically and experimentally by combining graphical phase processing
techniques [14]. Sun explored the homogeneity of bubble distributions in the dispersed phase.
Moreover, to solve the problem of significant differences in mixing times calculated by different
characterization methods, an average distance (AD) method was proposed to evaluate the mixing

CONTACT Jianxin Xu xujianxina@163.com Engineering Research Center of Metallurgical Energy Conservation and Emission
Reduction Ministry of Education, Kunming University of Science and Technology, Kunming 650093, PR China
© 2022 Taylor & Francis
EXPERIMENTAL HEAT TRANSFER 143

homogeneity of two fluids [15]. The variables considered in a heat exchanger optimization model
are the volumetric heat transfer coefficient, which is the objective function of the model, flow
rates of the dispersed and continuous phases, nozzle distribution of the dispersed phase, and
height of the heat transfer oil. Furthermore, the introduction of field synergy, a bubble growth
model, and droplet swarm behavior into the model are frequently used to optimize DCHEs
[12, 16].
Meanwhile, temperature difference fields are widely used to solve several heat transfer optimization
problems in other industries* [17–19]. These problems include difficulties in cooling hot components,
single-stream string-wound heat transfer network problems, volume-to-point heat transfer problems,
and radiation heat transfer optimization [20, 21].
In recent years, the temperature and temperature difference fields inside heat exchangers have
been explored, and the principle of the uniformity of temperature difference fields was proposed by
Guo, who concluded that the smaller the uniformity coefficient of a temperature difference field (i.e.,
the more uniform a temperature difference field is), the more efficient is the heat exchanger [22].
However, most studies are based on numerical simulations or experimental investigations on the
ability of heat exchangers to measure the temperatures of hot and cold liquids [22–24]. It is
complicated to measure the temperature of hot and cold liquid in direct contact heat exchanger,
therefore the principle of the uniformity of temperature fields proposed by Zengyuan Guo is not
easily applied to DCHEs. Most studies on DCHEs have focused on the mixing effect of the dispersed
and continuous phases and the influence of the uniformity of a bubble distribution on the heat
transfer efficiency. Presently, there are few studies on the uniformity of temperature fields and
internal fluid temperature, which are the key factors affecting the internal fluid performance of heat
exchangers.
In this study, a new method is proposed to calculate the uniformity factor of temperature difference
fields and can describe the temperature distribution of a heat exchanger at a given height.The
proposed the uniformity factor of temperature difference field is based on the heat exchanger micro-
element layer model developed in this study, which is an index that considers the uniformity of hot
and cold fluid mixing and the heat exchanger efficiency. A fully transparent heat exchanger was also
designed to facilitate thermal imaging so that temperature recordings of the heat exchanger surface
can be obtained and the uniformity of the temperature difference field can be calculated. Numerical
processing of the thermal image allows the heat exchange height for each set of operating conditions to
be obtained. A high-speed camera was used to record the bubble growth and distribution during heat
transfer of the hot and cold fluids, the uniformity of the bubble distribution was calculated using
the AD method, and the factors affecting the performance of the heat exchanger under the continuous
operation of the heat source were investigated comprehensively by incorporating the uniformity factor
of the temperature difference field.

Experiment
Experimental platform
The DCHE test setup built in this study consists of piping and measurement systems (Figure 1a).

Piping system
(1) Continuous phase cycle: This cycle was composed of an Organic Rankine Cycle (ORC) direct-
contact evaporator, electric heater, and gear oil pump, and a seamless steel pipe connects the different
parts. The direct-contact evaporator is manufactured by Plexiglas, with a height of 1,200 mm, an outer
diameter of 1,500 mm, and a wall thickness of 8 mm.
(2) Disperse phase cycle: This cycle was mainly composed of an ORC direct-contact evaporator,
condenser, liquid storage tank, and centrifugal pump, and the different parts are connected by
a seamless steel pipe.
144 X. GAO ET AL.

Figure 1. Organic Rankine Cycle test platform. a) Schematic diagram of Organic Rankine Cycle system (1 – direct contact evaporator,
2 – electric heater, 3 – gear oil pump, 4 – condenser, 5 – reservoir, 6 – centrifugal pump, 7 – control cabinet, 8, 9, 10, 11 – valves, 12,
13, 14, 15 – pressure gauge, 16, 17, 18, 19 – temperature gauge, 20, 21, 22 – flow meter, 23 – high speed camera). b) Physical
diagram of Organic Rankine Cycle circulation system.
EXPERIMENTAL HEAT TRANSFER 145

Measurement system
The measurement system included a pressure gauge, temperature gauge, thermocouples, and tem­
perature tester. The ORC direct-contact evaporator distributed eight k-type thermocouples evenly
along the axial direction (Figure 1b). To measure the axial temperature inside the evaporator, the
thermocouples were placed 100 mm apart and calibrated, and the measured temperature data were
collected and transmitted to a computer using a temperature test patrol.

Experimental process
The continuous phase of this test was the heat transfer oil, which was pumped through a gear oil pump
and into an electric heater heated to a specific temperature (120°C, 110°C, or 100°C), after which valve 11
was opened. The continuous phase was then fed from the top into the direct-contact evaporator. After
reaching a height of 100 cm, valve 12 was opened. After heat exchange with the dispersed phase occurs,
the dispersed phased was discharged through the bottom and into an electric heater for reheating.
Temperature and pressure gauges were installed at the inlet and outlet of the continuous phase to
measure the pressure and temperature of the continuous phase at the inlet and outlet. The gear oil pump
and electric heater were connected to the control cabinet. Their powers were controlled such that the flow
rate of the continuous phase was kept constant and the required temperature was reached. The
dispersion phase made use of CF3-CH2-CHF2 liquid. The centrifugal pump and valves 9 and 10 were
turned on, and the liquid CF3-CH2-CHF2 in the storage tank was pumped out through the centrifugal
pump at the bottom of the DCHE and into the heat exchanger. Heat exchange with the continuous phase
in the heat exchanger occurred after the phase change, producing CF3-CH2-CHF2 gas. The gas was
discharged through the top exhaust pipe and into the condenser, where it was cooled into a liquid.
Finally, the CF3-CH2-CHF2 liquid was transported into the storage tank for the next cycle.
The test started with heating the continuous phase, controlling the electric heater power such that it
is heated to a specific temperature (90°C), opening valve 11, and closing valves 9, 10, and 12. The gear
oil pump was then opened to pump the heated continuous phase fluid into the direct-contact
evaporator. Valve 12 was opened when the level of the liquid in the evaporator reached a specific
height (100 cm), and the level of the liquid in the direct-contact evaporator was kept constant to
maintain a constant level. Meanwhile, the centrifugal pump and valves 9 and 10 were opened to
maintain the flow of the dispersed phase in the evaporator at a constant rate for heat exchange with the
continuous phase. The temperature and pressure gauges and flow meter were displayed and recorded
in real-time on the computer. The dispersed phase in the evaporator was excluded after the test. The
specific test conditions are shown in Table 1.

Infrared thermal camera and high-speed camera


In this study, a 100 μm/pixel infrared high-speed cooling thermal imaging camera (German Infrared
Professional IRBIS3) was used to measure the temperature distribution at the surface of a quartz tube.
Before experimentation, the camera was calibrated according to a procedure in the literature [25], and

Table 1. Experimental conditions.


Case Initial temperature(continuous phase/disperse phase) Flow rate(continuous phase/dispersed phase) Flow rate ratio
L1 100°C/20°C 0.72 kg·s−1/0.12 kg·s−1 6:1
L2 100°C/20°C 0.72 kg·s−1/0.18 kg·s−1 4:1
L3 100°C/20°C 0.72 kg·s−1/0.36 kg·s−1 2:1
L4 110°C/20°C 0.72 kg·s−1/0.12 kg·s−1 6:1
L5 110°C/20°C 0.72 kg·s−1/0.18 kg·s−1 4:1
L6 110°C/20°C 0.72 kg·s−1/0.36 kg·s−1 2:1
L7 120°C/20°C 0.72 kg·s−1/0.12 kg·s−1 6:1
L8 120°C/20°C 0.72 kg·s−1/0.18 kg·s−1 4:1
L9 120°C/20°C 0.72 kg·s−1/0.36 kg·s−1 2:1
146 X. GAO ET AL.

the imaging frequency was set to 10 frames per second (FPS). The temperature distribution under
different conditions is shown in Figure 4. The infrared high-speed cooling thermal imager had
a measurement accuracy of 0.1°C. However, considering the accuracy of the resistance temperature
detector used for the temperature calibration, the accuracy of the temperature measurement was 1°C.
During experimentation, a high-speed camera (5F01-M, Revealer, China) was used to record
images of the bubbles during direct-contact boiling. To do this, the spatial resolution and imaging
frequency were set to 50 μm/pixel and 10,000 FPS, respectively, and Figure 7 shows the distribution of
the bubbles during the experiment. The black area is the continuous phase, while the white area is the
dispersed phase. During the experiment, image sequences were recorded simultaneously using the
high-speed camera and thermal imager.

Uncertainty analysis
The raw data obtained using the thermal imaging camera reflects the intensity of the invisible infrared
energy emitted from the evaporator surface. The captured infrared intensity was then converted to
temperature using a calibration curve obtained using a blackbody simulator (HFY-500A, DETU,
China). It should be noted that the accuracy of this conversion was approximately 2°C. The tempera­
ture distribution obtained using the infrared thermographer was processed using MATLAB, and the
average temperature was calculated. The uncertainties of the measurement devices, including the
thermometers, flow meters, and thermocouples, are shown in Table 2. The measurements of the
dispersed phase mass flow were performed using mass balance after each operation. The total error in
the measurements, including the digital scale error, propane compatibility in the water, operation
time, and possibility of condensate discharge from the continuous phase, was calculated. The error of
the mass flow rate of the dispersed phase was estimated to be 11%, and the uncertainty of the heat
exchange between the fluid inside the heat exchanger and the ambient air, as well as the error of the
temperature at the outlet of the continuous phase, were estimated to be 5%. According to the error
chain theory [26], the accuracy of the measured data exceeded 95%.

Temperature distribution at the heat exchanger surface


Uniformity principle of temperature
In DCHEs, cold and hot fluids are mixed, making it impossible to measure and calculate the
temperatures of cold and hot liquids. In this study, a mixture of hot and cold fluids was considered
to further discuss temperature uniformity. A coordinate system was established at the surface of the
heat exchanger (Figure 2). The fluid temperature at a given coordinate point is Ti;j and was obtained
using the thermal imaging camera. The uniformity factor of temperature was calculated using the
following equation:

Table 2. Relative error of the measuring equipment.


Measuring equipment Measurement error %
Thermal imaging camera ±0.42
Thermocouple ±0.44
Thermometer ±0.51
Manometer ±0.23
Flow meters ±0.30
EXPERIMENTAL HEAT TRANSFER 147

Figure 2. Heat exchanger surface coordinates.

Uniformity principle of the temperature difference field


The temperature difference field plays a vital role in improving the heat transfer of heat exchangers.
Guo et al. proposed a method involving the uniformity factor of the temperature difference field [22].
It can be expressed as follows:

It reflects the degree of uniformity of the temperature difference between hot and cold fluids in a heat
exchanger. The heat exchanger was divided into M � N heat transfer units, which indicates the
temperature difference between fluids in the sub-heat exchanger units. M and N were divided into
vertical and horizontal directions. φ was obtained to analyze the efficiency of the heat exchanger, and
as it increases, a more optimal heat transfer efficiency is obtained. ΔTi;j is the logarithmic mean
temperature difference (LMTD). The LMTD is the logarithmic mean of the temperature difference
between hot and cold fluids in an exchanger. For a heat exchanger with a constant area and high heat
transfer coefficient, the larger the LMTD, the larger is the amount of heat transferred.

Modeling
A constant temperature difference between the cold and hot fluid, ΔT¼ const, was applied to the heat
exchanger for heating. Therefore, when the temperature difference between the two streams on both
sides of the heat exchanger is kept constant, the balanced counterflow heat exchanger reaches the
highest efficiency, which is consistent with the uniformity principle of the temperature difference field
[27]. Since the contact area between the hot and cold fluids in a direct-contact heat exchanger cannot
be determined, the heat transfer between the hot and cold fluids is represented as follows:
148 X. GAO ET AL.

For the heat exchange process designed in this study, a physical model was established to better
understand the longitudinal temperature distribution of the heat exchanger, as shown in Figure 3,
which divides the heat exchanger into m micro-element layers. The continuous phase in the ith micro-
element layer transferred heat to the dispersed phase and flowed to the ith+1 micro-element layer.
After absorbing heat from the continuous phase, the dispersed phase in the ith micro-element layer
partially vaporized and rose to the ith+1 micro-element layer. dq is the heat transfer involved in both
processes, for which the integration over the height is represented by the following equation:

The mathematical depiction of the heat conduction problem shown in Figure 3 is as follows:

where Φ is the heating power of the electric heating system, λ is the heat transfer coefficient between
the micro-element layers, h is the heat transfer coefficient between the heat transfer oil and refrigerant,
and tf is the initial temperature of the refrigerant.
After integrating Equation (5) twice, we obtained the following:

where the constants c1 and c2 are determined by two boundary conditions. Finally, the temperature
distribution was obtained as follows:

Figure 3. Schematic diagram of heat transfer model


EXPERIMENTAL HEAT TRANSFER 149

For the counterflow direct-contact evaporator, the temperature decreased from top to bottom. The
temperature of the continuous and dispersed phases within the heat transfer model proposed in this
study decreased gradually from top to bottom. This experiment aimed to determine the distribution of
fluid temperature in the vertical direction inside the heat exchanger. Therefore, the continuous and
dispersed phases were considered mixed fluids. The heat was transferred from the upper surface,
which has a higher temperature, to the lower surface, which has a lower temperature, and 2Q is the
total heat transfer. In terms of the boundary conditions, this process is similar to one-dimensional
steady-state heat conduction. 2dq is the heat transfer from the ith micro-element layer to the ith+1
micro-element layer.

Uniformity factor of the temperature difference field


Based on the heat exchange model established in the above section, the temperature of the heat
exchanger surface was obtained using the thermal imager, and the temperature difference between the
micro-element layers of each dx is the temperature difference ΔT of the heat exchanger surface,
according to the coordinate points shown in Figure 2. φ is calculated as described below.
Let the temperature of the coordinate point ði; jÞ be aij . Thus, the temperature difference between
a0;n and a1;n is calculated as follows:

Moreover, the temperature difference between ai 1;j and a1;n is calculated as follows:

The uniformity factor of the temperature field was expressed as follows:

The closer the value is to 1, the more uniform is the temperature difference field.

Heat exchange height of the heat exchanger


The dispersed phase fluid was charged from the bottom to absorb the heat of the continuous phase.
A study [28] showed the progress of evaporation along the direct-contact evaporator. f ðxÞ is a function
of dx
dz (x denotes the dispersed phase gasification ratio) and was obtained by fitting the experimental
data, yielding an equation for the temperature distribution between the dispersed and continuous
phases, along with the evaporator height, as follows:

When a bubble rises to a certain height, its temperature is close to that of the heat transfer oil;
therefore, the temperature of the upper region of a heat exchanger remains unchanged. This height
refers to the heat exchange height.
150 X. GAO ET AL.

The simultaneous derivation of Z on both sides yields the following:

The solution is as follows:

The heat exchange height under different working conditions can be derived from Equation (16) [29].
Figure 4 shows the thermal image of the entire heat exchanger after stabilization of the heat transfer
process under each experimental condition. In Figure 5, a region was calculated for which the
temperature uniformity was equal to 0.9998 (0.9998 before refrigerant addition). The height from
the bottom of the heat exchanger to this area was the actual height of the heat exchange.
As shown in Figure 6, Z1 represents the actual height of heat exchange and Z2 represents the
theoretical height of heat exchange. Both Z1 and Z2 increase with increases in the temperature and
flow rate due to the expansion of the refrigerant evaporation volume. Therefore, the actual height of
heat exchange under each experimental condition was always more important than the theoretical
height of heat exchange. At the same temperature, as the flow rate of the refrigerant increased, the
retention volume of refrigerant in the heat transfer oil increased; therefore, the difference between the
actual and theoretical heights of heat exchange height gradually decreased.
Figure 7 shows the bubble distribution map obtained using the high-speed camera. When
the dispersed phase first entered the heat exchanger, the bubble distribution in the lower
region of the heat exchanger was not uniform owing to the flow rate and incomplete gasifica­
tion of the dispersed phase. The dispersed and continuous phases were more uniformly mixed
in the upper region of the heat exchanger; therefore, the upper region of the image was
processed for analysis.
Figure 8 shows the distribution of the bubbles, and the distribution of bubble groups in the heat
exchange process under each condition was recorded using the high-speed camera at a frame rate of
100 frames/s. The images were binarized, and the centers of mass were identified using MATLAB.
The AD method (the method that calculates the average distance between any two bubble centers of
mass) was proposed [15] to determine the homogeneity of bubble distribution as a way to determine
the uniformity of mixing between the dispersed and continuous phases. The equation used is as
follows.

where ðx; yÞ and ðu; vÞ are the coordinate points of any two center-of-mass points in the coordinate
axes shown in Figure 8.
After transforming Equation (17), this method was theoretically and experimentally validated by
Sun et al. EðXÞ ¼ 0:5214 was used as a standard reference value to study the mixing uniformity of the
bubble population in DCHEs [30].
EXPERIMENTAL HEAT TRANSFER 151

Figure 4. Thermal image under L1-L9 experimental condition.


152 X. GAO ET AL.

Figure 4. Continued.

Results and discussion


Uniformities of temperature and the temperature difference field
Uniformity of temperature
Nine sets of experiments were conducted under different operating conditions. For the thermal images
generated using the thermal imager and MATLAB, each intercepted the same rectangular position and
established a coordinate system (Figure 2). The temperature uniformity was calculated using Equation
(1), and the experimental results are shown in Figure 9. The process of temperature uniformity with
time was divided into three.
Trend 1: At a flow rate ratio of 6:1, the uniformity factor of temperature gradually decreased with
time, and the temperature field became increasingly uneven. This is because the experiment was timed
from when the refrigerant was charged into the evaporator. At the start of the experiment, the
refrigerant had not yet fully exchanged heat with the heat transfer oil. The temperature recorded
using the thermal imaging camera was almost equal to that of the heat transfer oil before the experiment
started, that is, when the temperature at each point was approximately equal to the inlet temperature of
the heat transfer oil, leading to a high temperature uniformity at the start of the experiment. As heat
exchange between the refrigerant and heat transfer oil progressed, the temperature of the mixed fluid
inside the heat exchanger started to decrease, and this temperature change started at the bottom, with
EXPERIMENTAL HEAT TRANSFER 153

Figure 5. Schematic diagram of actual heat exchange height

Figure 6. Heat exchange height under each test condition

a flow rate ratio of 6:1, a relatively low refrigerant flow rate, and a low height of heat exchange. When
the initial temperature difference gradually increased, the rate of heat exchange between the two fluids
increased, and a downward trend in the temperature uniformity became more apparent.
Trend 2: At a flow rate ratio of 4:1, the uniformity factor of temperature first decreased and then
increased. After the refrigerant was charged into the evaporator, the temperature gradually decreased
starting at the bottom, a temperature difference between each micro-element layer gradually became
apparent, and the temperature uniformity gradually decreased. Contrastingly, when the flow rate ratio
was 6:1, the refrigerant flow rate was relatively large, and the height of heat exchange increased. The
temperature of each micro-element layer was relatively stable. With increases in the refrigerant flow
154 X. GAO ET AL.

Figure 7. Bubble map taken by the high-speed camera.

Figure 8. Distribution of bubbles after binarization process.

and temperature, the rate of heat exchange between refrigerant and heat transfer oil increased, and
mass transfer between the micro-element layers occurred more frequently. This led to a gradual
increase in the temperature uniformity and an increased uniformity factor of temperature.
Trend 3: At a flow rate ratio of 2:1, the uniformity factor of temperature gradually increased with
time. This was because the contact area between refrigerant and heat transfer oil increases with the
refrigerant flow rate . However, increases or decreases in the refrigerant flow rate increased the height
of heat exchange to a level higher than those at the previous two ratios. Therefore, the temperature
trend of the entire heat exchanger internal fluid temperature at the start of the experiment was higher
EXPERIMENTAL HEAT TRANSFER 155

Figure 9. Variation of uniformity factor of temperature in L1-L9 working condition.

at the upper region and lower at the bottom region. As the heat exchange process continues, the rate of
heat exchange between refrigerant and heat transfer oil increased, and mass transfer between the
micro-element layers occurred more frequently. Therefore, the points in the region tended to have the
same temperatures, and the uniformity factor of temperature gradually increased with time.

Uniformity of the temperature difference field and bubble distribution


In the same way as the analysis of the temperature uniformity, the experimental data were also
obtained from nine sets of experiments, one from each of L1–L9. The uniformity factor of the
temperature difference field was calculated using Equation (10), and the uniformity of the bubble
distribution was calculated using Equation (17).
The uniformity coefficient of the temperature difference field tended to stabilize with a gradual
increase in the test time in each group, indicating that the temperature difference field became
increasingly uniform as time proceeded (Figure 10). This is because there was no stable heat exchange
process in the heat exchanger at the start of the experiment. There was a large temperature difference
between each heat exchanger layer. The height of heat exchange gradually stabilized, and the
temperature difference between the heat exchanger layers gradually converged. The uniformity factor
of the temperature difference field gradually increased with an increase in the height of heat exchange.
Furthermore, the uniformity of the bubble distribution gradually converged to 0.5214 as the
experiment proceeded, which indicated that the bubble distribution was also becoming more uniform.
156 X. GAO ET AL.

Figure 10. Uniformity of temperature difference field and bubble distribution in L1-L9 working condition.

The homogeneity of bubble distribution is an indicator to evaluate the homogeneity of mixing


between the dispersed and continuous phases. An increase in the homogeneity of the bubble
distribution indicates that the mixing of the continuous and dispersed phases is becoming increasingly
uniform, heat exchange between the two fluids is becoming increasingly efficient, and the uniformity
of the temperature difference field is increasing, which is consistent with the experimental results
shown in Figure 10.

Performance of heat exchangers


In the literature [28], the effective volume of an evaporator was proposed instead of the areas in direct
contact, and a formula for heat exchange was obtained, as follows:

In direct-contact evaporators, latent heat dominates the phase change process such that the above
equation can be written as follows:

Therefore, the volumetric heat transfer coefficient is calculated as follows:


EXPERIMENTAL HEAT TRANSFER 157

In addition, this study also made use of the positive heat exchanger performance of the efficiency table.
The efficiency of the heat exchanger was defined as follows:

At different flow rate ratios, the volumetric heat transfer coefficient increased with time. At a flow rate
ratio of 4:1, the volumetric heat transfer coefficient reached a maximum (Figure 11). tThe temperature
difference field is the most uniform, indicating that the heat exchange between the two phases is more
efficient. Therefore, when optimizing DCHEs, optimization of the temperature difference field is
important for improving heat transfer.
From Equation (21), it is clear that the efficiency of heat exchanger is determined by a larger initial
temperature difference between the continuous and dispersed phases and actual temperature differ­
ence between the cold or hot fluids in the heat exchanger. The L1–L9 experimental results show that all
nine experiments had greater temperature differences that those between the inlet and outlet of the
dispersed phase and continuous phase. This was owing to the phase change of the dispersed phase
during the heat transfer process and the considerably lower flow rate of the dispersed phase compared
to that of the continuous phase. Therefore, when the initial temperature difference between the hot
and cold fluids was constant, the final temperature of the dispersed phase determined the efficiency of
the heat exchanger. Under different conditions, the efficiency of the heat exchanger increases with the
uniformity of the temperature difference field (Figure 11). The more uniformly the dispersed phase is

Figure 11. Variation of heat exchanger efficiency and volumetric heat transfer coefficient.
158 X. GAO ET AL.

Figure 12. Closed direct contact heat exchange platform.

mixed with the continuous phase, the more adequate is the heat exchange between the two. The above
experimental results are consistent with the principle of the uniformity of the temperature difference
field, thus, verifying the applicability of the principle of the uniformity of the temperature difference
field in direct-contact heat exchangers operated using continuous heat sources. When the initial
temperature difference between the dispersed and continuous phases was 60°C or 90°C and the flow
rate ratio was 4:1, the height of heat exchange was close to that of the heat exchanger when the bubble
rose to the liquid surface That is, when the temperature reached that of the continuous phase, the
temperature difference field was the most uniform, and the efficiency of the heat exchanger was the
highest. When the initial temperature difference between the dispersed and continuous phases was
120°C, owing to the liquid surface height, the working conditions L3, L6, and L9 did not reach
maximum heights of heat transfer, leading to a flow rate ratio of 6:1. At this flow rate ratio, the
dispersed phase of the inlet and outlet temperature difference was higher, the heat exchanger efficiency
was relatively high, and the overall heat exchanger efficiency decreased compared to the initial
temperature difference between the dispersed and continuous phases at 90°C.
To verify the applicability of the uniformity factor of the temperature difference field obtained from
Equation (10) under different experimental conditions, we conducted the same experiment using the
test stand shown in Figure 12, where the heat exchanger was adequately insulated and the longitudinal
temperature was measured using internally distributed thermocouples. The experimental results were
consistent with the principle of the uniformity of the temperature difference field.

Conclusion
In this study, the temperature difference field uniformity calculation method was improved and
applied to a DCHE. Moreover, the uniformities of temperature and the temperature difference field,
as well as the homogeneity of the bubble distribution under nine working conditions, were analyzed.
The conclusions are as follows:
(1) The heat exchange height is influenced by the initial temperature difference of the fluid flow.
The heat exchange height also affects the uniformities of temperature and temperature difference field.
(2) The heat exchange height affects the change in the temperature uniformity. The temperature
uniformity trend changes from decreasing to increasing as the height of heat exchange approaches that
of the heat exchanger.
EXPERIMENTAL HEAT TRANSFER 159

(3) The uniformities of the temperature difference field and bubble distribution were investigated in
a system operated using a continuous low-temperature heat source. Both the temperature difference
field and the bubble distribution tended to be uniform as time proceeded. The applicability of the AD
method under continuous low-temperature heating conditions was also verified.
(4) The volumetric heat transfer coefficient and efficiency of the heat exchanger were proportional
to the uniformity factor of the temperature difference field, indicating that the uniformity factor of the
temperature difference field calculated using the heat exchanger micro-element layer model was
consistent with the uniformity principle of the temperature difference field.

Nomenculture

DCHE Direct contact heat exchanger


f Uniformity factor of temperature
Ti;j The temperature at any point on the surface of the heat exchange (°C)
Tav Average temperature of heat exchanger surface (°C)
φ Uniformity factor of temperature difference field (°C)
ΔT Temperature difference
tc Temperature of continuous phase (°C)
td Temperature of disperse phase (°C)
tc;i Inlet temperature of continuous phase (°C)
tc;o Dispersion phase exit temperature (°C)
td;i Inlet temperature of disperse phase (°C)
td;o Dispersion phase exit temperature (°C)
EðxÞ Bubble uniformity calculated by AD method (°C)
hv Volumetric heat exchange (°C)
Vr volume of the continuous phase of the evaporator
ε Efficiency (°C)
Δtm Logarithmic mean temperature difference between hot and cold fluids (°C)
hfg Latent heat of evaporation (°C)
mc Continuous phase mass flow rate (kg/s)
md Disperse phase mass flow rate (kg/s)
Uo Disperse phase mass flow rate (kg/s)
kc thermal conductivity of the continuous phase (W/m·K)
kv velocity factor
Ac condenser cross-sectional area (m2)
w volume fraction
Z height (m)
f ðzÞ functions on Z
a0 initial radius of two-phase bubble (m)
Cp;c specific heat of continuous phase (kJ/kg·K)
Cp;d specific heat of dispersed phase (kJ/kg·K)

Acknowledgments
We wish to thank the referees for their numerous detailed questions and constructive criticism, which greatly improved
the presentation.

Disclosure statement
No potential conflict of interest was reported by the author(s).

Funding
This work was supported by the National Natural Science Foundation of China [52166004].
160 X. GAO ET AL.

References
[1] B. Li, et al. “Experimental investigation of bubble group and temperature distribution uniformity in the direct
contact boiling heat transfer process[J],” Exp. Therm. Fluid Sci, vol. 134, pp. 110620, 2022. DOI: 10.1016/j.
expthermflusci.2022.110620.
[2] T. Lemenand, et al. “Turbulent direct-contact heat transfer between two immiscible fluids[J],” Int. J. Therm. Sci,
vol. 49, no. 10, pp. 1886–1898, 2010. DOI: 10.1016/j.ijthermalsci.2010.05.014.
[3] P. A. H, et al. “An updated review on application of nanofluids in heat exchangers for saving energy[J],” Energy
Convers. Manage, vol. 198, pp. 111886, 2019.
[4] M. Ghasemi, et al. “A numerical study on thermal analysis and cooling flow fields effect on PEMFC
performance[J],” Int. J. Hydrogen Energy, vol. 42, no. 38, pp. 24319–24337, 2017. DOI: 10.1016/j.ijhydene.2017.
08.036.
[5] Z. Y. Guo, H. Y. Zhu, and X. G. Liang, “Entransy—A physical quantity describing heat transfer ability[J],” Int.
J. Heat Mass Transf, vol. 50, no. 13–14, pp.2545–2556, 2007. DOI: 10.1016/j.ijheatmasstransfer.2006.11.034.
[6] C. Yadav and R. R. Sahoo, “Effect of nano-enhanced PCM on the thermal performance of a designed cylindrical
thermal energy storage system[J],” Exp. Heat Transfer, vol. 34, no. 4, pp.356–375, 2021. DOI: 10.1080/08916152.
2020.1751744.
[7] D. D. L, et al. “Comparison of direct and indirect contact heat exchange to improve recovery of bio-oil[J],” Appl.
Energy, vol. 251, pp. 113346, 2019. DOI: 10.1016/j.apenergy.2019.113346.
[8] A. Vaisi, et al. “Experimental examination of condensation heat transfer enhancement with different perforated
tube inserts[J],” Exp. Heat Transfer, pp. 1–27, 2021. DOI:10.1080/08916152.2021.1991510.
[9] B. A. S, et al. “Measuring the average volumetric heat transfer coefficient of a liquid–liquid–vapour direct contact
heat exchanger[J],” Appl. Therm. Eng, vol. 103, pp. 47–55, 2016. DOI: 10.1016/j.applthermaleng.2016.04.067.
[10] O. S. Al-Yahia, H. J. Yoon, and D. Jo, “Experimental study of bubble flow behavior during flow instability under
uniform and non-uniform transverse heat distribution[J],” Nucl. Eng. Technol, vol. 52, no. 12, pp.2771–2788,
2020. DOI: 10.1016/j.net.2020.05.025.
[11] R. Prakash, S. K. Majumder, and A. Singh, “Bubble size distribution and specific bubble interfacial area in two–
phase microstructured dense bubbling bed[J],” Chem. Eng. Res. Des, vol. 156, pp. 108–130, 2020. DOI: 10.1016/j.
cherd.2020.01.032.
[12] Q. Wang, et al. “A novel method for measuring spatial uniformity of irregular boiling bubbles in a direct contact
heat exchanger[J],” Int. J. Energy Res, vol. 44, no. 11, pp. 8823–8840, 2020. DOI: 10.1002/er.5577.
[13] J. Xu, et al. “Accurate estimation of mixing time in a direct contact boiling heat transfer process using statistical
methods[J],” Int. Commun. Heat Mass Transf, vol. 75, pp. 162–168, 2016. DOI: 10.1016/j.icheatmasstransfer.
2016.04.012.
[14] Q. Xiao, et al. “Extraction and evolution of bubbles attributes in a two-phase direct contact evaporator[J],” Int.
J. Heat Mass Transf, vol. 124, pp. 761–768, 2018. DOI: 10.1016/j.ijheatmasstransfer.2018.04.002.
[15] H. Sun, et al. “Evolution and quantification of distribution uniformity of bubbles using computational
geometry[J],” Chem. Eng. Sci, vol. 247, pp. 116910, 2022. DOI: 10.1016/j.ces.2021.116910.
[16] H. Cai, et al. “Numerical study on uniformity of temperature difference field in a spiral tube heat exchanger[J],”
Appl. Therm. Eng, vol. 190, pp. 116798, 2021. DOI: 10.1016/j.applthermaleng.2021.116798.
[17] Q. Xiao, et al. “Non-uniformity quantification of temperature and concentration fields by statistical measure
and image analysis[J],” Appl. Therm. Eng, vol. 124, pp. 1134–1141, 2017. DOI: 10.1016/j.applthermaleng.2017.
06.073.
[18] Y. Syaiful, et al. “Effect of attack angle of concave and convex winglet vortex generators on heat transfer and
pressure drop in evaporator fin and tube heat exchanger with field synergy principle using numerical
simulation[J],” Syst. Rev. Pharm, vol. 11, no. 12, pp. 873–887, 2020.
[19] X. Lan, et al. “Experimental and numerical study on the temperature uniformity of a variable density alternating
obliquely truncated microchannel[J],” Int. J. Heat Mass Transf, vol. 176, pp. 121440, 2021. DOI: 10.1016/j.
ijheatmasstransfer.2021.121440.
[20] M. Y. T, et al. “Numerical study on temperature uniformity in a novel mini-channel heat sink with different flow
field configurations[J],” Int. J. Heat Mass Transf, vol. 85, pp. 147–157, 2015. DOI: 10.1016/j.ijheatmasstransfer.
2015.01.093.
[21] M. Halimi, A. El Amrani, and C. Messaoudi, “New experimental investigation of the circumferential temperature
uniformity for a PTC absorber[J],” Energy., vol. 234, pp. 121288, 2021.
[22] G. Z. Y, et al., “Principle of uniformity of temperature difference field in heat exchanger[J],” Sci. China (Series E:
Technol. Sci.)., vol. 68-75, pp. 1, 1996.
[23] Z.-Y. Guo, S.-Q. Zhou, L. Zhi-Xin, and L.-G. Chen, “Theoretical analysis and experimental confirmation of the
uniformity principle of temperature difference field in heat exchanger[J],” Int. J. Heat Mass Transf, vol. 45, no. 10,
pp.2119–2127, 2002. DOI: 10.1016/S0017-9310(01)00297-6.
EXPERIMENTAL HEAT TRANSFER 161

[24] R. H. Fu and X. Zhang, “A rigorous proof for the uniformity principle of temperature difference field in heat
exchanger with deductive method[J],” Int. J. Heat Mass Transf, vol. 114, pp. 135–144, 2017. DOI: 10.1016/j.
ijheatmasstransfer.2017.06.012.
[25] A. Richenderfer, et al. “Investigation of subcooled flow boiling and CHF using high-resolution diagnostics[J],”
Exp. Therm. Fluid Sci, vol. 99, pp. 35–58, 2018. DOI: 10.1016/j.expthermflusci.2018.07.017.
[26] J. Kline, “Describing uncertainty in single sample experiments[J],” Mech. Engin, vol. 75, pp. 3–8, 1953.
[27] Z. Y. Guo and Y. C. Hua, “Comments on the statement that the temperature difference field uniformity principle
is a duplicate of the principle, ΔT/T= const, for balanced counter-flow heat exchangers[J],” Int. J. Heat Mass
Transf, vol. 127, pp. 1343–1346, 2018. DOI: 10.1016/j.ijheatmasstransfer.2018.07.108.
[28] S. Sideman, G. Hirsch, and Y. Gat, “Direct contact heat transfer with change of phase: effect of the initial drop size
in three-phase heat exchangers[J],” AIChE J, vol. 11, no. 6, pp.1081–1087, 1965. DOI: 10.1002/aic.690110622.
[29] A. S. Baqir, H. B. Mahood, and A. H. Sayer, “Temperature distribution measurements and modelling of a
liquid-liquid-vapour spray column direct contact heat exchanger[J],” Appl. Therm. Eng, vol. 139, pp. 542–551,
2018. DOI: 10.1016/j.applthermaleng.2018.04.128.
[30] M. W. Abdulrahman, “Experimental studies of direct contact heat transfer in a slurry bubble column at high gas
temperature of a helium–water–alumina system[J],” Appl. Therm. Eng, vol. 91, pp. 515–524, 2015. DOI: 10.1016/
j.applthermaleng.2015.08.050.

You might also like