Sie sind auf Seite 1von 8

Available online at www.sciencedirect.

com

ScienceDirect
Availableonline
Available onlineatatwww.sciencedirect.com
www.sciencedirect.com
Energy Procedia 00 (2017) 000–000

ScienceDirect
ScienceDirect
www.elsevier.com/locate/procedia

Energy
EnergyProcedia
Procedia129 (2017) 000–000
00 (2017) 451–458
www.elsevier.com/locate/procedia

IV International Seminar on ORC Power Systems, ORC2017


13-15 September 2017, Milano, Italy

Thermal performance of brazed


The 15th International metalfoam-plate
Symposium on District Heatingheat exchanger as
and Cooling
anthe
Assessing evaporator
feasibility foroforganic Rankine
using the cycle
heat demand-outdoor
temperature function
Daefor a long-term
Yeon district
Kim, Kyung Chun Kim*heat demand forecast
School of Mechanical Engineering,
a,b,c Pusan National
a University,a San 30, Jangjeon-dong,
b Geumjeong-gu, Busan
c 609-735, Republic ofc Korea
I. Andrić *, A. Pina , P. Ferrão , J. Fournier ., B. Lacarrière , O. Le Corre
a
IN+ Center for Innovation, Technology and Policy Research - Instituto Superior Técnico, Av. Rovisco Pais 1, 1049-001 Lisbon, Portugal
Abstract b
Veolia Recherche & Innovation, 291 Avenue Dreyfous Daniel, 78520 Limay, France
c
Département Systèmes Énergétiques et Environnement - IMT Atlantique, 4 rue Alfred Kastler, 44300 Nantes, France
Compact-sized organic Rankine cycle (ORC) power generators call for higher performance and down-sized heat exchangers. Heat
exchangers in ORC systems, especially evaporators contribute to a big portion of the system size as well as the capital cost, and
their price is also directly related to their size. Since the heat transfer area plays a direct role in the performance of heat exchangers,
high-porosity
Abstract metal foams are proposed for insertion into heat exchanger channels to enhance the heat transfer mechanism in
evaporators. Their high surface area to volume ratio makes them a good candidate for manufacturing high-performance heat
exchangers. The metalfoams
District heating networks are arecommonly
being considered to improve
addressed performance
in the literature whileofkeeping
as one the mosttheeffective
size of heat exchangers
solutions small. In this
for decreasing the
experimental study, the performance of a 100 kW class heat exchanger with the channels brazed
greenhouse gas emissions from the building sector. These systems require high investments which are returned through with nickel based metal foam
the and
heat
stainless steeltosheets
sales. Due was investigated.
the changed A hot water
climate conditions andloop was designed
building renovationfor heat input.heat
policies, The demand
cold sideinof the
the future
heat exchanger works
could decrease,
with R245fa as the working fluid.
prolonging the investment return period. The phase-change heat transfer experiments were performed with different mass flow rates
ranging from 0.32 to 0.51 kg/s while the operating pressure was at 10 to 13 bar with hot water inlet
The main scope of this paper is to assess the feasibility of using the heat demand – outdoor temperature function for heat demand temperature was 133℃.
Furthermore,
forecast. Thethedistrict
heat transfer performance
of Alvalade, located is compared
in Lisbonwith a commercial
(Portugal), plate as
was used heata exchanger
case study. (produced by Alfalaval,
The district AC120EQ)
is consisted of 665
which was custom-made
buildings that vary in both as a 100kW class evaporator.
construction period and In the comparison
typology. test, refrigerant
Three weather scenariosside has 0.5kg/s
(low, medium, of high)
mass flow rate atdistrict
and three 44℃
and heat source side has 1.23 kg/s mass flow rate at 132℃. Although the pressure drop in the metal
renovation scenarios were developed (shallow, intermediate, deep). To estimate the error, obtained heat demand values were foam plate heat exchanger is
increased compared to that of the conventional plate heat exchanger, increase of the
compared with results from a dynamic heat demand model, previously developed and validated by the authors. recovered waste heat from the heat source is
much higher showed
The results due to higher overall
that when onlyheat transfer
weather coefficient.
change As a result,
is considered, the the energy
margin density
of error of the
could be new heat exchanger
acceptable for some isapplications
about 2.5
times higherinthan
(the error that demand
annual of the conventional
was lower than plate20%
heat exchanger.
for all weather scenarios considered). However, after introducing renovation
©scenarios,
2017 The Authors. Published by Elsevier Ltd. (depending on the weather and renovation scenarios combination considered).
© 2017 Thethe
Peer-review under
error value
Authors. increased
Published
responsibility by
of
up to 59.5%
Elsevier Ltd. committee
scientific of the IV International Seminar onper
ORC Power Systems.
The value of
Peer-review slope
under coefficient of
responsibility the
increased on average
the scientific committee within theIV
of the range of 3.8% Seminar
International up to 8%on ORC decade,
Powerthat corresponds to the
Systems.
decrease in the number of heating hours of 22-139h during the heating season (depending on the combination of weather and
Keywords:
renovationMetalfoam;
scenariosOrganic RankineOn
considered). cycle;
theEvaporator;
other hand, Plate heat exchanger
function intercept increased for 7.8-12.7% per decade (depending on the
coupled scenarios). The values suggested could be used to modify the function parameters for the scenarios considered, and
improve the accuracy of heat demand estimations.

© 2017 The Authors. Published by Elsevier Ltd.


Peer-review under responsibility of the Scientific Committee of The 15th International Symposium on District Heating and
* Corresponding author. Tel.: +82-51-510-1536; fax: +82-51-515-7866.
Cooling.
E-mail address: kckim@pusan.ac.kr
Keywords: Heat demand; Forecast; Climate change
1876-6102 © 2017 The Authors. Published by Elsevier Ltd.
Peer-review under responsibility of the scientific committee of the IV International Seminar on ORC Power Systems.

1876-6102 © 2017 The Authors. Published by Elsevier Ltd.


Peer-review under responsibility of the Scientific Committee of The 15th International Symposium on District Heating and Cooling.
1876-6102 © 2017 The Authors. Published by Elsevier Ltd.
Peer-review under responsibility of the scientific committee of the IV International Seminar on ORC Power Systems.
10.1016/j.egypro.2017.09.225
452 Dae Yeon Kim et al. / Energy Procedia 129 (2017) 451–458
2 Dae Yeon Kim and Kyung Chun Kim / Energy Procedia 00 (2017) 000–000

1. Introduction

The down-sizing of heat exchangers is one of the most important issues in the organic Rankine cycle (ORC) system.
High-porosity open-cell metal foams have gained attention as a heat transfer medium due to their excellent heat transfer
performance. It has a high surface to volume ratio (5600m2/m3) and has a low relative density due to its high porosity
(≥90%) [1-4]. Use of metal foams as a heat transfer medium is expected to develop a high performance compact heat
exchanger.
Calmidi et al.[4] conducted experimental and numerical studies on forced convection in highly porous aluminum
metal foams in a variety of porosities and pore densities using air as the fluid medium. Mancin et al.[5, 6] studied
liquid and flow boiling heat transfer in highly porous copper foams using R134a and R1234ze as working fluids and
electric heaters as heat sources. The effect of refrigerant mass flow rate, heat flux and steam quality on heat transfer
coefficient is shown, and dry out phenomenon and pressure drop characteristics are investigated. Abadi et al.[7, 8]
performed an experimental study on the single-phase heat transfer and pressure drop of metal foam filled tube and
metal foam filled plate heat exchanger with R245fa. Abadi et al.[9] also studied 10-kW heat exchangers with channels
filled with copper metal foam. R245fa is used as a working fluid and single-phase and two-phase heat transfer and
pressure drop are investigated. They have reported that metal foams increase the pressure drop of the channel but this
increases the overall heat transfer coefficient of the heat exchanger up to 2.3 times compared to simple channel heat
exchangers without metal foams. Kim et al.[10] proposed the use of metallic foams as plate heat exchangers for the
intermediated fluid vaporizer in an LNG regasification system. They designed and numerically analyzed the optimized
configuration of the metal foam plate heat exchangers and shell-and-tube heat exchangers to compare two heat
exchangers. They have reported that the metal foam plate heat exchangers have higher heat transfer performance and
lower pressure drop than the shell-and-tube heat exchangers through their design for a specific applications.
Plate heat exchangers have been extensively studied by many researchers [11-14], however, no comparative studies
have been done on the metal foam heat exchangers and plate heat exchangers. In this study, a brazed nickel foam plate
heat exchanger was designed and manufactured and the heat exchanger performance was compared to a commercial
brazed plate heat exchanger (BPHE) under the evaporator operating conditions of the ORC using R245fa and hot water.

Nomenclature

BPHE Brazed plate heat exchanger


BFPHE Brazed foam plate heat exchanger
P Pressure [bar]
HTA Heat transfer area [m2]
Q Heat transfer rate [kW]
T Temperature [℃]
ΔTLMTD Log mean temperature difference [℃]
x vapour quality
g gravitational acceleration [m/s2]
ρ Density [kg/m3]
U Overall heat transfer coefficient [kW/m2K]
V Volume of heat exchanger [cm3]
L Distance between inlet and outlet port [m]
i Enthalpy[kJ/kg]
m Mass flow rate [kg/s]
f Darcy friction factor
Subscripts
m Mean value
i Inlet
o Outlet
l Saturated liquid state
v Saturated vapour state
h Hot stream
c Cold stream
Dae Yeon Kim et al. / Energy Procedia 129 (2017) 451–458 453
Dae Yeon Kim and Kyung Chun Kim / Energy Procedia 00 (2017) 000–000 3

2. Experimental setup

2.1. Test loop

As shown in Fig. 1, the test rig is composed of three different loops. The heat exchanger is installed in the refrigerant
loop and uses R245fa as the working fluid on the cold stream of the test section. The R245fa loop consists of a reservoir
tank before the pump with a height of 1.6m. The slurry pump (Hydra-Cell D/G-10) moves the refrigerant in this loop.
At the downstream of the pump, there is a gear type flowmeter (KEM-ZHA04) is located with a digital indicator. The
hot water loop of the test section is consists of a 115kW electric heater embedded boiler with temperature controller
and insulation. Pressure transducers and T-type thermocouples are located around the loop to measure the pressure
and temperature at inlet and outlet of every component during the experiment. The pressure control valve is installed
1m from the outlet port of the heat exchanger to control the operating pressure conditions. A data acquisition system
(National Instrument) is connected to these measurement devices. The customized LABVIEW program receives all
the flow rate, temperature and pressure values from the data acquisition system and the REFPROP 9.1 library converts
the measured information into the useful properties of water and R245fa at each point. The operating conditions are
listed in Table 1. In order to compare two different types of heat exchangers, this experiment was performed under
the same conditions of mass flow rate, operating pressure and inlet temperature for each hot and cold side. The mass
flow rate of hot water is constant at 1.23kg/s and the refrigerant is 0.3~0.5kg/s. The inlet temperature of hot stream
and cold stream is 132℃ and 44℃, respectively. The operating pressure of hot side is constant at 6.5 bar while the
operating pressure in the cold side is 10 to 13 bar. The superheated temperature is over 5℃ in all the experimental
cases.

Table 1. Test conditions.


Hot stream Cold stream
Fluid Water R245fa
Mass flow rate (kg/s) 1.23 0.3~0.5
Inlet temperature (℃) 132 44
Operating pressure (bar) ~6.5 10~13
Superheated temperature (℃) ≥5

Fig. 1. Schematic diagram of test loop.

2.2. Test section

Two types of heat exchangers are considered in this study. One is a brazed nickel foam-plate heat exchanger that
consists of 25PPI (pores per inch) nickel foam, stainless steel thin plate and channel frame. A total of 21 plates of
nickel foam and stainless steel sheet channels are brazed with the filler metal simultaneously. This is a vertical counter-
current type that the hot water flows from the top to the bottom and the refrigerant flows from the bottom to the top
454 Dae Yeon Kim et al. / Energy Procedia 129 (2017) 451–458
4 Dae Yeon Kim and Kyung Chun Kim / Energy Procedia 00 (2017) 000–000

in every other channel. Before performing the experiments, a hydrostatic test was conducted at 120 bar for 10 minutes
to ensure its durability. The shape and dimensions of the nickel foam-plate heat exchanger are shown in Fig. 2 and
Table 2. To compare the performance of the brazed nickel foam-plate heat exchanger, a commercial brazed plate heat
exchanger (Alfalaval-AC120EQ 70H) was used. The dimension of the BPHE is also listed in Table 2. The shape and
size of two heat exchangers are visibly compared as shown in Fig. 3.

Fig. 2. Design of metal foam heat exchanger channel and size comparison of the prototype of BFPHE and BPHE.

Fig. 3. Size comparison of the prototype of BFPHE and BPHE.

Table 2. Geometric characteristics of tested heat exchangers


Foam Plate BPHE (AC120EQ)
Type Brazed Brazed
Plate width, W(mm) 100 192
Channel height (mm) 5 2.35
Number of channels (ea) 21 70
Horizontal distance between port centers (mm) 75 92
Vertical distance between port centers (mm) 496 519
Inlet/outlet ports diameter (mm) 20/20 40/40
Entrance area of channel (cm2) 0.5 1.58
Heat transfer area (m2) 0.826 6.98
Volume (cm )3
7,455 20,731

2.3. Data reduction

The experimental measurements were made according to the following procedure. The physical properties of water
and R245fa were calculated according to the averages of the inlet and outlet bulk fluid temperature and pressure. The
heat transfer rate between the hot and cold stream in the test section was obtained from the properties of hot water
side.
Dae Yeon Kim et al. / Energy Procedia 129 (2017) 451–458 455
Dae Yeon Kim and Kyung Chun Kim / Energy Procedia 00 (2017) 000–000 5


=Q m h (ih ,i − ih ,o ) (1)

in which ṁ and i are the mass flow rate and enthalpy, respectively. The overall heat transfer coefficient was calculated
by the following equation

Q (2)
U=
HTA * ∆TLMTD

where HTA is the heat transfer area. The log mean temperature difference ΔTLMTD is determined from

(Th ,i − Tc ,o ) − (Th ,o − Tc ,i ) (3)


∆TLMTD =
(T − T )
ln h ,i c ,o
(Th ,o − Tc ,i )

The outlet vapor quality is calculated based on the heat duty of hot side and the latent heat at the pressure measured
at the outlet port of the test section.

Q − Qsp Q icold ,in − il ,cold ,out (4)


xout = 
= 

mcold ilv ,cold ,out mcold ilv ,cold ,out ilv ,cold ,out

The total pressure drop between the inlet and outlet of the refrigerant side is measured by the pressure transducers.
The total pressure drop across the heat exchangers consists of four components; static pressure drop due to the vertical
configuration of heat exchangers, pressure drops due to acceleration of the refrigerant, pressure drops due to
inlet/outlet manifolds and frictional pressure drop through the main core of plate channel. The frictional pressure drop
can be obtained by removing other pressure loss factors from this total pressure drop:

∆Pf ,TP = ∆Ptotal − ∆Pstatic − ∆Pmomentum − ∆Pport (5)

The static pressure drop is:

∆Pstatic =g ⋅ ρ m ⋅ L (6)

where ρm is the mean density of the two-phase flow.

−1
x 1 − xmean  (7)
ρ m  mean +
=
 ρv ρl 

The acceleration pressure drop defined in terms of refrigerant mass flux is given by the momentum difference at the
inlet and outlet of the channel:

 1 1 (8)
∆Pmomentum= G 2 ⋅  −  ⋅ ∆x
 ρ v ρl 

Δx is the vapor quality difference between the inlet and outlet of the channel. The port pressure drop is calculated
by:
456 Dae Yeon Kim et al. / Energy Procedia 129 (2017) 451–458
6 Dae Yeon Kim and Kyung Chun Kim / Energy Procedia 00 (2017) 000–000

 G2  (9)
∆Pport =
1.5  
 2 ⋅ ρm 

Finally, the two-phase Darcy friction factor for evaporation of R245fa was obtained by the following relation:

∆Pf ,TP ⋅ Dh ⋅ ρ m (10)


fTP =
1 2
⋅G ⋅ L
2

2.4. Experimental uncertainties

An experimental error analysis was performed according to the procedure outlined by Moffat [15]. The T-type
thermocouples have an accuracy of ±0.05K. The flowmeter has an accuracy of ±0.5% and the pressure transducers
have an accuracy of ±0.01bar. Based on these errors, an uncertainty of ±7.6~13.7% exists in calculated values of
overall heat transfer coefficient. Table 3 summarizes the uncertainties in this experiment. The properties of the
refrigerant and hot water were calculated by REFPROP 9.1.

Table 3. Uncertainties of different parameters in the experiment.


Parameter Uncertainty
Mass flow rate ±5%
Temperature ±0.6%
Pressure ±1.7%
Heat duty ±5.2%
Overall heat transfer coefficient ±7.6~13.7%

3. Results and discussion

As shown in Table 4, the outlet quality on refrigerant side decreases with increasing mass flow rate of refrigerant
due to the constant flow rate and temperature of hot water stream. The BFPHE (brazed foam plate heat exchanger)
has an outlet quality of 1.2 to 1.1 and the BPHE has an outlet quality of 2.0 to 1.5. The superheated temperature at the
refrigerant stream outlet decreases as the mass flow rate increases. The BFPHE indicates 40℃ to 4.5℃ but the BPHE
overheat at 46℃ to 30℃. As the mass flow rate increases, the heat transfer rate increases from 75kW to 108kW for
BFPHE and from 80 to 126kW for BPHE. BPHE has a heat duty of 6.0 to 11.6% higher than BFPHE.

Table 4. Operating pressure, outlet vapor quality and super heat temperature of the refrigerant side

Mass flow rate (kg/s) 0.31 0.33 0.36 0.39 0.40 0.42 0.44 0.46 0.49
Operating pressure (bar) 9.8 10.4 11.0 11.3 11.8 12.2 12.3 12.6 13.0
Outlet quality (-) 1.2 1.2 1.2 1.2 1.2 1.2 1.1 1.1 1.1
BFPHE
Super heat (℃) 41.3 34.8 33.7 25.2 22.8 20.5 14.5 7.6 4.5
Heat duty (kW) 76.5 81.2 87.4 90.9 95.9 99.6 101.9 104.5 108.4
Operating pressure (bar) 9.8 10.3 11.1 11.6 11.9 12.3 12.3 12.7 12.8
Outlet quality (-) 2.0 1.9 1.9 1.8 1.8 1.7 1.7 1.6 1.5
BPHE
Super heat (℃) 46.0 43.3 40.3 39.1 37.3 35.4 35.2 31.5 31.8
Heat duty (kW) 82.3 86.3 92.1 96.2 100.9 105.2 110.5 113.2 120.6
Dae Yeon Kim et al. / Energy Procedia 129 (2017) 451–458 457
Dae Yeon Kim and Kyung Chun Kim / Energy Procedia 00 (2017) 000–000 7

(a) (b)

Fig. 4. Comparison of (a) overall heat transfer coefficient and (b) the frictional pressure drop
(a) (b)

Fig. 5. Comparison of (a) Darcy friction factor and (b) energy density of two heat exchangers.
The overall heat transfer coefficient for two different types of heat exchangers is shown in Fig. 4(a). The overall
heat transfer coefficient of BFPHE is about 2,510 to 2,770 W/m2K and the BPHE is about 850 to1460 W/m2K. The
overall heat transfer coefficient of BFPHE is 82~198% higher than that of BPHE. The total pressure drop is measured
between the inlet and outlet port of the test section by a pressure transducer according to the mass flow rate of
refrigerant side. The operating pressure increases from 10 to 13 bar as the flow rate of refrigerant increases, as shown
in the Table 4. As shown in Fig. 4(b), the frictional pressure drop is calculated based on the equation 5. As shown in
Table 4, as the mass flow rate increases, the outlet quality on the refrigerant side decreases, and as the flow rate
increases, the liquid flow area increases in the two-phase flow region along the test section. Thus, the frictional
pressure drop is scarcely increased as the mass flow rate increases due to the different flow regime and increased
operating pressure in the test section. Because the vapor flow at the same mass flux conditions has a higher pressure
drop than the liquid flow due to the effect of the fluid velocity more effective than the fluid density and the higher the
operating pressure, the lower the pressure loss in the test section. This causes the frictional pressure drop of the BFPHE
to be reduced even in the higher mass flow region. The frictional pressure drop of BFPHE is about 112 to 100 kPa,
but the friction pressure drop of BPHE is 10 to 19 kPa. Because the inlet area of the heat exchanger is different, as
shown in Table 2, the flow velocity through the BFPHE channel is three times faster than BPHE and the pressure loss
is also relatively high.
Darcy friction factor of two different heat exchangers was obtained using equation 10 and compared as shown in
Fig. 5(a). As the mass flow rate increases, the friction factor of the BFPHE decreases from 89 to 51 while the friction
factor of BPHE increases from 21 to 34. The lower vapor quality reduces the friction factor of BFPHE, while the
friction factor of BPHE increases somewhat. The friction factor of BFPHE is 320% higher than BPHE in the low mass
flow region, but the gap of friction factor between two heat exchangers is reduced to 57% in the high mass flow region.
458 Dae Yeon Kim et al. / Energy Procedia 129 (2017) 451–458
8 Dae Yeon Kim and Kyung Chun Kim / Energy Procedia 00 (2017) 000–000

Another parameter for comparing different sizes and different types of heat exchangers is the energy density,
defined as a heat duty per unit volume of the heat exchanger. As shown in Fig. 5(b), the energy density of BFPHE is
2.5 times higher than the energy density of BPHE.

4. Conclusion

A prototype of brazed nickel foam plate heat exchanger considering heat recovery system in ORC system was
experimentally investigated. The heat transfer and pressure drop performance of the BFPHE is compared to the
performance of commercial brazed plate heat exchanger (Alfalaval-AC120EQ) under the same operating conditions
depending on the mass flow rate of refrigerant. The volume of BFPHE is 1/3 times and the flow area is narrower than
1/3 of BPHE. Therefore, the frictional pressure drop was more than 5 times higher, but the friction factor was 50%
higher in the high mass flow region. The heat duty was 10% higher in BPHE but the energy density of BFPHE is
2.5times better than BPHE, which is expected to reduce the size of the ORC evaporator.

Acknowledgements

This research was supported by the National Research Foundation of Korea (NRF) with a grant funded by the
Korean government (MSIP) through the Global Core Research Center for Ships & Offshore Plants (GCRC-SOP, No.
2011-0030013). This research was also supported by Global PH.D Fellowship Program through the National Research
Foundation of Korea (NRF) funded by the Ministry of Education (NRF-2014H1A2A1022186).

References

[1] Kim SY, Paek JW, Kang BH. Flow and heat transfer correlations for porous fin in a plate-fin heat exchanger. J Heat Trans-T Asme.
2000;122:572-8.
[2] Dukhan N. Correlations for the pressure drop for flow through metal foam. Exp Fluids. 2006;41:665-72.
[3] Bhattacharya A, Calmidi VV, Mahajan RL. Thermophysical properties of high porosity metal foams. Int J Heat Mass Tran. 2002;45:1017-31.
[4] Calmidi VV, Mahajan RL. Forced convection in high porosity metal foams. J Heat Trans-T Asme. 2000;122:557-65.
[5] Mancin S, Diani A, Doretti L, Rossetto L. R134a and R1234ze(E) liquid and flow boiling heat transfer in a high porosity copper foam. Int J
Heat Mass Tran. 2014;74:77-87.
[6] Diani A, Mancin S, Doretti L, Rossetto L. Low-GWP refrigerants flow boiling heat transfer in a 5 PPI copper foam. Int J Multiphas Flow.
2015;76:111-21.
[7] Abadi GB, Moon C, Kim KC. Experimental study on single-phase heat transfer and pressure drop of refrigerants in a plate heat exchanger with
metal-foam-filled channels. Appl Therm Eng. 2016;102:423-31.
[8] Abadi GB, Kim KC. Experimental heat transfer and pressure drop in a metal-foam-filled tube heat exchanger. Exp Therm Fluid Sci. 2017;82:42-
9.
[9] Abadi GB, Kim DY, Yoon SY, Kim KC. Thermal performance of a 10-kW phase-change plate heat exchanger with metal foam filled channels.
Appl Therm Eng. 2016;99:790-801.
[10] Kim DY, Sung TH, Kim KC. Application of metal foam heat exchangers for a high-performance quefied natural gas regasification system.
Energy. 2016;105:57-69.
[11] Amalfi RL, Vakili-Farahani F, Thome JR. Flow boiling and frictional pressure gradients in plate heat exchangers. Part 2: Comparison of
literature methods to database and new prediction methods. Int J Refrig. 2016;61:185-203.
[12] Hsieh YY, Chiang LJ, Lin TF. Subcooled flow boiling heat transfer of R-134a and the associated bubble characteristics in a vertical plate heat
exchanger. Int J Heat Mass Tran. 2002;45:1791-806.
[13] Khan TS, Khan MS, Chyu MC, Ayub ZH. Experimental investigation of evaporation heat transfer and pressure drop of ammonia in a 60
degrees chevron plate heat exchanger. Int J Refrig. 2012;35:336-48.
[14] Lee H, Li S, Hwang Y, Radermacher R, Chun HH. Experimental investigations on flow boiling heat transfer in plate heat exchanger at low
mass flux condition. Appl Therm Eng. 2013;61:408-15.
[15] Moffat RJ. Using Uncertainty Analysis in the Planning of an Experiment. J Fluid Eng-T Asme. 1985;107:173-8.

Das könnte Ihnen auch gefallen