You are on page 1of 8

Available online at www.sciencedirect.

com
ScienceDirect
ScienceDirect
Energy Procedia 00 (2018) 000–000
Availableonline
Available onlineatatwww.sciencedirect.com
www.sciencedirect.com
Energy Procedia 00 (2018) 000–000 www.elsevier.com/locate/procedia

ScienceDirect
ScienceDirect
www.elsevier.com/locate/procedia

Energy
EnergyProcedia
Procedia160 (2019) 000–000
00 (2017) 499–506
www.elsevier.com/locate/procedia
2nd International Conference on Energy and Power, ICEP2018, 13–15 December 2018,
Sydney,
2nd International Conference on Energy Australia
and Power, ICEP2018, 13–15 December 2018,
Sydney, Australia
Moisture diffusion measurement and evaluation for porous
Moisture
The diffusion measurement
15th International
membranes Symposium onand
used in enthalpy Districtevaluation
Heating and for
exchangers porous
Cooling
membranes used in enthalpy exchangers
AssessingAhmed
the feasibility
K. Albdoora,bof usingMa
, Zhenjun thea,*,heat
Paul demand-outdoor
Coopera
temperatureSustainable
function for
AhmedBuildings
K. Albdoor
a a long-term
Research
a,b
, Zhenjun Ma district
a,
*, Paul
Centre, University of Wollongong, heat
Cooper
2522, demand
a
NSW, Australia forecast
b
Department of Mechanical Techniques, Al-Nasiriyah Technical Institute, Southern Technical University, 64001, Thi-Qar, Iraq
a
Sustainable Buildings Research Centre, University of Wollongong, 2522, NSW, Australia
a,b,c a a b c c
b
I. Andrić *, A. Pina , P. Ferrão , J. Fournier ., B. Lacarrière , O. Le Corre
Department of Mechanical Techniques, Al-Nasiriyah Technical Institute, Southern Technical University, 64001, Thi-Qar, Iraq

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
Abstract c
Département Systèmes Énergétiques et Environnement - IMT Atlantique, 4 rue Alfred Kastler, 44300 Nantes, France
Air-to-air membrane enthalpy exchangers using semi-permeable membranes are widely used in building ventilation systems to
pre-condition
Air-to-air membranethe supply air byexchangers
enthalpy exchangingusing energy with the exhaust
semi-permeable air stream.are
membranes Moisture
widely diffusivity is one ventilation
used in building of the mostsystems
importantto
properties of porous membranes. This property has a significant influence on the design and performance
pre-condition the supply air by exchanging energy with the exhaust air stream. Moisture diffusivity is one of the most important of membrane enthalpy
Abstract
exchangers.
properties of In this membranes.
porous study, moisture Thisdiffusion
property has resistances of five
a significant porousonmembranes
influence the design were measured, and
and performance the effectsenthalpy
of membrane of the
membrane
exchangers. pore size
In this on the moisture
study, moisture diffusivity
diffusion were evaluated
resistances under different test conditions. The five porous membranes tested
District heating networks are commonly addressed in theofliterature
five porous membranes
as one of the most were measured,
effective and the
solutions for effects of
decreasing the
the
included:
membrane i) two
pore PVDF
size on (Polyvinylidene
the moisture difluoride)
diffusivity were membranes
evaluated with
under mean pore
different diameters
test of
conditions. 0.22
The µm
five and 0.45
porous µm respectively,
membranes tested
greenhouse gas emissions from the building sector. These systems require high investments which are returned through the heat
ii)sales.
two Nylon
included: Due tomembranes
i) two PVDF
the withclimate
0.1 µmconditions
(Polyvinylidene
changed and 0.45 µm
difluoride) and pore sizes with
membranes
building respectively,
mean pore
renovation anddiameters
iii) aheat
policies, PESof (Polyethersulfone)
0.22 µm
demand in and
the 0.45 membrane
futureµm with a
respectively,
could decrease,
0.1 µm pore
ii)prolonging
two Nylonthe size. A
membranes theoretical
investmentwith return model
0.1period. to predict the effectiveness of a crossflow membrane enthalpy
µm and 0.45 µm pore sizes respectively, and iii) a PES (Polyethersulfone) membrane with exchanger was alsoa
developed
0.1
Theµm main with
porescope respect
size.of A to the
thistheoreticallatent
paper is tomodel heat
assesstotransfer.
thepredict The experimental
the of
feasibility effectiveness results
using the heat showed
of demand
a crossflow that the PVDF
membrane
– outdoor membrane
enthalpy
temperature with
exchanger
function a mean
for heatwas pore
also
demand
diameter
forecast.ofThe
developed 0.45district
with µm outperformed
respect of
to the latentthe
Alvalade, heatothers
located in Lisbon
in
transfer. terms of(Portugal),
the moisture
The experimental wasdiffusivity.
used
results as aThe
showed caseresults
that study.
the from
PVDF the
Themembranetheoretical
district iswith model
a meanof
consisted the
ofpore
665
enthalpy
buildings exchanger
that vary agreed
in bothwell with those
construction from
period the
and experiments
typology. reported
Three in
weatherthe literature.
scenarios The
(low,
diameter of 0.45 µm outperformed the others in terms of the moisture diffusivity. The results from the theoretical model of the latent
medium, effectiveness
high) and was found
three to
district
berenovation
insensitive scenarios
to the were
outdoor developed
air conditions.(shallow, intermediate, deep). To estimate the error,
enthalpy exchanger agreed well with those from the experiments reported in the literature. The latent effectiveness was found toobtained heat demand values were
becompared
insensitivewithto results from air
the outdoor a dynamic
conditions.heat demand model, previously developed and validated by the authors.
The results showed that when only weather change is considered, the margin of error could be acceptable for some applications
©(the
2019
2018 TheinAuthors.
error annual Published
demand was by Elsevier
lower than Ltd.20% for all weather scenarios considered). However, after introducing renovation
This is an
scenarios, open
the access
error articleincreased
value under thetheupCC BY-NC-ND license (https://creativecommons.org/licenses/by-nc-nd/4.0/)
This is
© 2018 The
Selection
an open
and
access
Authors. article
Published
peer-review
under
under Elsevier 59.5%
CCto
byresponsibility Ltd. of (depending
BY-NC-ND license on
the scientific
scientific
the weather and renovation scenarios combination considered).
(https://creativecommons.org/licenses/by-nc-nd/4.0/)
committee
The value
Selection
This is an and of slope coefficient
peer-review under increased
responsibilityon average
of the within the range of of
committee 3.8%
of the 2nd
2ndtoInternational
the up
open access article under the CC BY-NC-ND license (https://creativecommons.org/licenses/by-nc-nd/4.0/) 8% per decade,
International Conference on
on Energy
Energytoand
that corresponds
Conference the
and
Power,
decrease
Power, ICEP2018.
in the
ICEP2018. number of heating hours of 22-139h during the heating season (depending on the combination of weather and
Selection and peer-review under responsibility of the scientific committee of the 2nd International Conference on Energy and
renovation
Power, scenarios considered). On the other hand, function intercept increased for 7.8-12.7% per decade (depending on the
ICEP2018.
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
* Corresponding responsibility of the Scientific Committee of The 15th International Symposium on District Heating and
author.
Cooling.
E-mail address: zhenjun@uow.edu.au (Z Ma)
* Corresponding author.
E-mail address: zhenjun@uow.edu.au (Z Ma)
Keywords:©Heat
1876-6102 2018demand; Forecast;
The Authors. Climatebychange
Published Elsevier Ltd.
This is an open access article under the CC BY-NC-ND
1876-6102 © 2018 The Authors. Published by Elsevier Ltd. license (https://creativecommons.org/licenses/by-nc-nd/4.0/)
Selection and peer-review under responsibility of the scientific committee of the 2nd International Conference on Energy and Power, ICEP2018.
This is an open access article under the CC BY-NC-ND license (https://creativecommons.org/licenses/by-nc-nd/4.0/)
Selection and peer-review under responsibility of the scientific committee of the 2nd International Conference on Energy and Power, ICEP2018.
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 © 2019 The Authors. Published by Elsevier Ltd.
This is an open access article under the CC BY-NC-ND license (https://creativecommons.org/licenses/by-nc-nd/4.0/)
Selection and peer-review under responsibility of the scientific committee of the 2nd International Conference on Energy and Power, ICEP2018.
10.1016/j.egypro.2019.02.198
500 Ahmed K. Albdoor et al. / Energy Procedia 160 (2019) 499–506
2 Albdoor et al. / Energy Procedia 00 (2018) 000–000

Keywords: enthalpy exchangers; porous membranes; moisture diffusivity

1. Introduction

Air-to-air energy recovery ventilators (ERVs) are widely used to reduce the energy required for conditioning the
air in building applications [1]. This type of technology transfers valuable sensible and latent energy in the exhaust
air stream to the supply air stream. Using ERVs as part of building heating, ventilation, and air-conditioning
(HVAC) systems can save up to 65 % of energy consumed in processing the fresh air, and improve indoor air quality
[2].
Membrane enthalpy exchangers (MEEs) use semi-permeable membranes that separate exhaust and supply air
streams and allow for both heat and moisture transfer between the two air streams. Compared to other energy
recovery devices (e.g. enthalpy wheels and runaround membrane enthalpy exchangers), air-to-air MEEs offer
advantages such as low cross-contamination rate, simplicity, high effectiveness and less embodied energy [3].
Membrane properties can significantly influence the design and performance of MEEs. Moisture diffusivity is
one of the most important properties of the membranes [4]. Niu and Zhang [5] and Min and Su [4] found that
membrane moisture permeability can significantly affect both the latent effectiveness and total effectiveness (i.e.
enthalpy effectiveness). On the other hand, the moisture diffusion resistance is not constant and it varies with
changes in temperature and humidity [6]. This can result in significant performance variations of the MEE under
various operating conditions. To evaluate these variations, a set of experiments are often required in order to test the
membrane moisture diffusion under a wide range of operating conditions.
Typically, membranes are classified into dense and porous types according to the mean pore size. The pore size of
‘dense’ membranes is commonly of the order of 0.1 nm, while that of ‘porous’ membranes is of the order of 0.1 μm
[7]. Although both types of membranes have been used in the MEE fabrication, they have completely different
moisture transfer mechanisms.
In the majority of the previous studies, the moisture diffusivity of the membrane was considered either a constant
value [8, 9] obtained from a single experimental test or a variable value [10]. However, only a few studies
investigated the influence of the operating conditions on the permeability of the porous membranes, and their impact
on the performance of the MEE.
In this study, the influence of the membrane pore size on the moisture diffusion resistances of five different
porous membranes was measured using a wet cup test method under different test conditions. The moisture
diffusivity of the tested membranes was then calculated. The performance of a crossflow MEE using these
membranes was then evaluated using a theoretical model to calculate its latent effectiveness.

Nomenclature
J Mass flux kg/m2.s Acronyms
G Weight change kg Le Lewis number
T Time during which weight change s NTUm Number of mass transfer units
occurred
Ac Test cup area m2 Nu Nusselt number
A Total transfer area m2 Sh Sherwood number
Dm Moisture diffusivity m2/s
Rm Moisture diffusion resistance m2.s/kg
rm Volumetric moisture diffusion resistance s/m Greek letters
K Convective mass transfer coefficient m/s εl Latent effectiveness
𝑚𝑚𝑚𝑚̇𝑚𝑚𝑚𝑚𝑚𝑚𝑚𝑚𝑚𝑚𝑚𝑚 Minimum mass flow rate kg/s Ρ density kg/m3
Dv Diffusivity of water vapour in air m2/s subscripts
Dh Hydraulic diameter m A Air
W Absolute humidity kg/kg E Exhaust
Um Total mass transfer coefficient m/s S Supply
Ahmed K. Albdoor et al. / Energy Procedia 160 (2019) 499–506 501
Albdoor et al. / Energy Procedia 00 (2018) 000–000 3

2. Experiments

2.1. Materials and test matrix

Five porous membranes with different pore sizes were chosen for the test. The membrane samples were supplied
by Tianshan Precision Filter Material Co., Ltd. (Hangzhou, China), and their properties are presented in Table 1.
The tests were conducted under different humidity and temperature conditions as listed in Table 2.

Table 1 Properties of the five membranes tested

Porous membrane Description Pore size (μm) Thickness (μm)


PVDF45 Polyvinylidene difluoride 0.45 100
Nylon45 Nylon 0.45 100
PVDF22 Polyvinylidene difluoride 0.22 100
Nylon10 Nylon 0.1 100
PES10 Polyethersulfone 0.1 100

Table 2 Test matrix

Test cases 1 2 3 4 5 6 7 8 9
Temperature (ºC) 27.5 30.0 32.5 27.5 30.0 32.5 27.5 30.0 32.5
Relative humidity (%) 80 80 80 50 50 50 30 30 30

2.2. Moisture diffusion resistance measurement

The moisture diffusivity of the membranes used in enthalpy exchangers has a crucial role in influencing the latent
effectiveness and total effectiveness. The water vapour flux transferred through the membrane is inversely
proportional to the moisture diffusion resistance (MDR). There are various devices and methods available to
measure the moisture permeability of the membranes [11]. In this study, the wet cup method specified in the
standard of ASTM E96 [12] was used as an experimental approach as this method offers a low cost, simplicity and
high accuracy [13]. As shown in Fig. 1, the test apparatus consisted of a 90 mm diameter PVC cup and distilled
water, and the test sample was mounted on the rim of the cup by covering with a gasket and flange to hold its
position. The cup assembly was placed on an electronic weighing scale with a resolution of 0.01 g, and connected to
the data logger and tested in an environmental chamber, which was capable of maintaining the temperature and
relative humidity at specified set points with uncertainties of ± 0.1 oC and 1.0 %, respectively. A controlled vapour
pressure gradient was created for driving the moisture transfer from the high humidity region inside the cup of 100
% to the environmental chamber where the humidity and temperature were changed according to the test conditions.
Once the test reached the steady state, the constant water vapour transmission rate (J) was then calculated according
to Eq. (1).

𝐺𝐺𝐺𝐺
𝐽𝐽𝐽𝐽 = 𝑡𝑡𝑡𝑡 𝐴𝐴𝐴𝐴 (1)
𝑐𝑐𝑐𝑐

The corrections due to the still air and surface resistance were performed according to those specified in the
standard [12]. The water vapour transmission rate through the membrane can also be expressed as Eq. (2) [14]. The
volumetric moisture diffusion resistance can then be calculated by Eq. (3).

𝑤𝑤𝑤𝑤𝑠𝑠𝑠𝑠 −𝑤𝑤𝑤𝑤𝑒𝑒𝑒𝑒
𝐽𝐽𝐽𝐽 = (2)
𝑅𝑅𝑅𝑅𝑚𝑚𝑚𝑚
502
4 Ahmed K.
Albdoor Albdoor
et al. et al.
/ Energy / Energy
Procedia 00Procedia 160 (2019) 499–506
(2018) 000–000

𝑟𝑟𝑟𝑟𝑚𝑚𝑚𝑚 = 𝑅𝑅𝑅𝑅𝑚𝑚𝑚𝑚 𝜌𝜌𝜌𝜌𝑎𝑎𝑎𝑎 (3)

Fig. 1 Wet cup test apparatus.

3. Latent effectiveness model of the MEE

The latent effectiveness of the MEE was theoretically calculated to evaluate the performance of the MEE.
Moisture diffusivity of the porous membranes was required to calculate the latent effectiveness. The correlation
between the moisture diffusivity and the operating conditions could be obtained from the experimental results.
For cross flow with unmixed streams, the latent effectiveness can be calculated with ε-NTU method as shown in
Eq. (4) [14], in which the number of the mass transfer unit is defined by Eq. (5).
0.78 �−1
𝑒𝑒𝑒𝑒𝑒𝑒𝑒𝑒𝑒𝑒𝑒𝑒�−𝑁𝑁𝑁𝑁𝑁𝑁𝑁𝑁𝑁𝑁𝑁𝑁𝑚𝑚𝑚𝑚
𝜀𝜀𝜀𝜀𝑙𝑙𝑙𝑙 = 1 − 𝑒𝑒𝑒𝑒𝑒𝑒𝑒𝑒𝑒𝑒𝑒𝑒 � −0.22 � (4)
𝑁𝑁𝑁𝑁𝑁𝑁𝑁𝑁𝑁𝑁𝑁𝑁𝑚𝑚𝑚𝑚

𝜌𝜌𝜌𝜌𝑎𝑎𝑎𝑎 𝑁𝑁𝑁𝑁𝑚𝑚𝑚𝑚𝐴𝐴𝐴𝐴
𝑁𝑁𝑁𝑁𝑁𝑁𝑁𝑁𝑁𝑁𝑁𝑁𝑚𝑚𝑚𝑚 = (5)
𝑚𝑚𝑚𝑚̇𝑚𝑚𝑚𝑚𝑚𝑚𝑚𝑚𝑚𝑚𝑚𝑚

The overall mass transfer coefficient is the only unknown parameter needed to obtain the latent effectiveness,
which can be obtained from Eq. (6) [14]. The membrane moisture transfer resistance, which is the middle term in
the parentheses of Eq. (6), is dependent on the type of the membrane used. This term cannot be neglected as it
accounts for more than half of the overall mass transfer coefficient [14].

1 𝛿𝛿𝛿𝛿 1 −1 1 𝑟𝑟𝑟𝑟𝑚𝑚𝑚𝑚 1 −1
𝑁𝑁𝑁𝑁𝑚𝑚𝑚𝑚 𝐴𝐴𝐴𝐴 = � + + � =� + + � (6)
𝑘𝑘𝑘𝑘𝑒𝑒𝑒𝑒 𝐴𝐴𝐴𝐴 𝐷𝐷𝐷𝐷𝑚𝑚𝑚𝑚 𝐴𝐴𝐴𝐴 𝑘𝑘𝑘𝑘𝑠𝑠𝑠𝑠 𝐴𝐴𝐴𝐴 𝑘𝑘𝑘𝑘𝑒𝑒𝑒𝑒 𝐴𝐴𝐴𝐴 𝐴𝐴𝐴𝐴 𝑘𝑘𝑘𝑘𝑠𝑠𝑠𝑠 𝐴𝐴𝐴𝐴

The exhaust air and supply air convective mass transfer coefficients were assumed to be identical, and can be
calculated by Eq. (7), where the Sherwood number is determined using Eq. (8). The values of the Lewis number
(Le) and Nusselt number (Nu) were taken as 0.85 and 7.54, respectively [10, 15].

𝑆𝑆𝑆𝑆ℎ 𝐷𝐷𝐷𝐷𝑣𝑣𝑣𝑣
𝑘𝑘𝑘𝑘 = (7)
𝐷𝐷𝐷𝐷ℎ

𝑆𝑆𝑆𝑆ℎ = 𝑁𝑁𝑁𝑁𝑁𝑁𝑁𝑁𝐿𝐿𝐿𝐿𝑒𝑒𝑒𝑒 −1⁄3 (8)


Ahmed K. Albdoor et al. / Energy Procedia 160 (2019) 499–506 503
Albdoor et al. / Energy Procedia 00 (2018) 000–000 5

4. Results and discussion

4.1. Effects of membrane pore size on moisture diffusion resistance

Fig. 2a) provides a comparison of the moisture diffusion resistances of the five membranes under the various
relative humidity conditions specified in Table 2 and at a constant temperature of 30 oC (i.e. test cases 2, 5 and 8).
The PVDF45 membrane showed the lowest resistance while Nylon10 offered the highest resistance among the five
porous membranes tested. The membrane resistances were insensitive to the relative humidity difference across the
membrane. The result agreed with the conclusion reported by Ge et al. [6].
Fig. 2b) shows the effect of the temperature on the moisture diffusion resistance of the five membranes. The
investigations were performed at different temperatures and the same relative humidity of 50 % in the environmental
chamber (i.e. test cases 4-6). The MDR of the membranes slightly increased as the temperature increased.

a 65.00
b 70.00
Moisture diffusion resistance (m 2 .s/kg)

Moisture diffusion resistance (m 2 .s/kg)


60.00 65.00

60.00
55.00
55.00
50.00
50.00
45.00
45.00
40.00
40.00
35.00 PVDF45 Nylon45 PVDF22 Nylon10 PES10 PVDF45 Nylon45 PVDF22 Nylon10 PES10
35.00

30.00 30.00
20 30 40 50 60 70 80 90 26 27 28 29 30 31 32 33
Relative humidity (%) Temperature (oC)

Fig. 2 Effects of the test conditions on the moisture diffusion resistance a) Humidity influence; and b) Temperature influence.

Fig. 3 presents the influence of the pore size on the MDR of the PVDF and Nylon membranes at a constant
temperature of 32.5 oC. It is worth noting that the PES membrane used had a single pore size and was therefore not
presented in this figure. For the PVDF membrane, the MDR increased from 44.8 to 55.9 m2.s/kg at the relative
humidity of 80 % and from 48.7 to 63.1 m2.s/kg at the relative humidity of 30 % when the pore size changed from
0.45 µm to 0.22 µm. The MDR of the Nylon membrane increased from 56.5 to 62.4 at 80 % relative humidity and
from 60 to 68.5 at 30 % relative humidity when the pore size was 0.45 µm and 0.1 µm, respectively.

80
Moisture diffusion resistance (m 2 .s/kg)

70 RH=80% RH=30%

60

50

40

30

20

10

0
PVDF45 PVDF22 Nylon45 Nylon10

Fig. 3 The effect of the pore size on moisture diffusion resistance at 32.5 oC.
504 Ahmed K. Albdoor et al. / Energy Procedia 160 (2019) 499–506
6 Albdoor et al. / Energy Procedia 00 (2018) 000–000

4.2. Validation of the exchanger latent effectiveness model

Fig. 4 presents a comparison of the results from the theoretical model and the experimental results reported in
[16] for a particular enthalpy exchanger. The latent effectiveness was calculated for three MEE cores, which have
different channel heights and different numbers of the channels as presented in [16] but have the same outer
dimensions. It can be seen that an acceptable agreement between the two sets of values can be observed. The
maximum deviation between the theoretical and experimental data was 6.5 %.

0.70
0.65
0.60
Latent effectiveness (%)

0.55
0.50
0.45
0.40
0.35
0.30
Exp. Core A Exp. Core B Exp. Core.C
0.25 Calc. Core A Calc.Core B Calc. Core C
0.20
100 120 140 160 180 200 220
Volume flaw rate (m 3 /h)

Fig. 4 Model validation results using the experimental data reported in [16].

4.3. Effects of operating conditions on latent effectiveness of the MEE

The latent effectiveness under different operating conditions was calculated to evaluate the performance of the
MEE. The dimensions of a crossflow MEE as listed in Table 3 were used in this analysis. The calculation was
carried out based on an air indoor temperature and relative humidity of 25 ºC and 47 %, respectively.

Table 3 Specifications of the crossflow MEE [15].

MEE specifications Value


Membrane length 0.25 m
Membrane width 0.25 m
Channel height 2 mm
Number of channels of each stream 180
Mass flow rate 0.1 kg/s

The moisture diffusivity through the membranes was calculated based on the experimental data for all
membranes tested and the results are presented in Fig. 5a). The diffusivity of each membrane was slightly affected
by the test conditions. The PVDF45 membrane yielded the maximum average diffusivity (i.e. 1.9 x10-6 m2/s), while
the Nylon10 membrane offered the lowest average diffusivity (i.e. 1.41 x10-6 m2/s).
In order to extend the results of the experimental tests to a wider range of the operating conditions, the
correlations between the moisture diffusivity and the driving force (i.e. humidity difference) were obtained from the
PVDF45 and PVDF22 membrane test cases. Fig. 5b) presents the test results and the correlations for the two
membranes. The moisture diffusivity of the membranes decreased almost linearly with increasing humidity
difference.
Ahmed K. Albdoor et al. / Energy Procedia 160 (2019) 499–506 505
Albdoor et al. / Energy Procedia 00 (2018) 000–000 7

a 2.10E-06
b 2.50E-06

1.90E-06 Dm = 2.05x10 -6 - 1.08x10 -5 (ws - we)


Moisture diffusivity (m 2 /s)

Moisture diffusivity (m 2 /s)


2.00E-06

1.70E-06
1.50E-06
1.50E-06 Dm = 1.82x10 -6 - 1.85x10 -5 (ws - we)
1.00E-06
1.30E-06

5.00E-07 PVDF22 PVDF45


1.10E-06 PVDF45 Nylon45 PVDF22 Nylon10 PES10
Linear (PVDF22) Linear (PVDF45)

9.00E-07 0.00E+00
0 2 4 6 8 10 0 0.005 0.01 0.015 0.02 0.025
Test No. Humidity difference (ws -we ) (kg/kg)

Fig. 5 a) Moisture diffusivity of the tested membranes, and b) The correlations of the moisture diffusivity and the moisture difference of the
PVDF45 and PVDF22 membranes.

Fig. 6 shows the variation of the latent effectiveness with changes of outdoor relative humidity at two different
temperatures for the PVDF45 and PVDF22 membranes. The effectiveness was calculated using the moisture
diffusivity obtained from the correlations generated in Fig. 5b). The constant average effectiveness was also plotted.
As expected the performance of the PVDF45 membrane was better than the PVDF22 membrane. The outdoor
temperature almost did not influence the latent effectiveness, and the effectiveness of the two outdoor temperatures
overlapped for each membrane. The outdoor relative humidity had a small impact on the latent effectiveness. The
latent effectiveness decreased slightly from 67 % to 66.4 % for the PVDF45 membrane, and from 65.8 % to 64.5 %
for the PVDF22 membrane when the outdoor humidity increased from 45 % to 90 %.

0.680

0.670
Latent effectiveness (%)

0.660

0.650

0.640

PVDF45 Variable T30 PVDF45 constant


0.630
PVDF22 variable T 30 PVDF22 constant
PVDF45 variable T35 PVDF22 variable T35
0.620
40 50 60 70 80 90 100
Outdoor relative humidity (%)

Fig. 6 Latent effectiveness under different operating conditions.

5. Conclusion

Moisture diffusion resistance is a very important property of the membranes used in membrane enthalpy
exchangers. The moisture diffusion resistances of five different porous membranes were measured using a wet cup
method. The influences of the pore size on the diffusive resistance at different test conditions were evaluated. A
theoretical model of the latent effectiveness of a crossflow membrane enthalpy exchanger was also presented.
The results showed that the test conditions and the pore size slightly affected the moisture diffusion resistance.
The PVDF45 membrane showed the lowest diffusive resistance, while the Nylon10 offered the highest diffusive
resistance. The latent effectiveness was almost not affected by the outdoor temperature while it was slightly affected
by the outdoor humidity. The effectiveness decreased from 67 % to 66.4 % for the PVDF45 membrane, and from
65.8 % to 64.5 % for the PVDF22 membrane when the relative humidity increased from 45 % to 90 %.
506 Ahmed K. Albdoor et al. / Energy Procedia 160 (2019) 499–506
8 Albdoor et al. / Energy Procedia 00 (2018) 000–000

Acknowledgements

The first author would like to acknowledge the Higher Committee for Education Development of Iraq for the
provision of his PhD scholarship and support.

References

[1] M. Justo Alonso, P. Liu, H.M. Mathisen, G. Ge, C. Simonson. "Review of heat/energy recovery exchangers for use in ZEBs in cold climate
countries." Building and Environment 84 (2015): 228-37.
[2] L.Z. Zhang. "Progress on heat and moisture recovery with membranes: From fundamentals to engineering applications." Energy Conversion
and Management 63 (2012): 173-95.
[3] A. Engarnevis, R. Huizing, S. Green, S. Rogak. "Heat and mass transfer modeling in enthalpy exchangers using asymmetric composite
membranes." Journal of Membrane Science 556 (2018): 248-62.
[4] J. Min, M. Su. "Performance analysis of a membrane-based enthalpy exchanger: Effects of the membrane properties on the exchanger
performance." Journal of Membrane Science 348 (2010): 376-82.
[5] J.L. Niu, L.Z. Zhang. "Membrane-based Enthalpy Exchanger: material considerations and clarification of moisture resistance." Journal of
Membrane Science 189 (2001): 179-91.
[6] G. Ge, G.I. Mahmood, D.G. Moghaddam, C.J. Simonson, R.W. Besant, S. Hanson, et al. "Material properties and measurements for semi-
permeable membranes used in energy exchangers." Journal of Membrane Science 453 (2014): 328-36.
[7] J. Woods. "Membrane processes for heating, ventilation, and air conditioning." Renewable and Sustainable Energy Reviews 33 (2014): 290-
304.
[8] S. Koester, M. Falkenberg, M. Logemann, M. Wessling. "Modeling heat and mass transfer in cross-counterflow enthalpy exchangers."
Journal of Membrane Science 525 (2017): 68-76.
[9] L.Z. Zhang. "Heat and mass transfer in plate-fin enthalpy exchangers with different plate and fin materials." International Journal of Heat
and Mass Transfer 52 (2009): 2704-13.
[10] J. Min, M. Su. "Performance analysis of a membrane-based energy recovery ventilator: Effects of outdoor air state." Applied Thermal
Engineering 31 (2011): 4036-43.
[11] A.M. Elizabeth, K. Myoungsook, S. Huensup. "A comparison of standard methods for measuring water vapour permeability of fabrics."
Measurement Science and Technology 14 (2003): 1402.
[12] ASTM E96/E96 M-05. Standard test methods for water vapor transsmission of materials, ASTM International, West Conshohocken (2005).
[13] P. Slanina, Š. Šilarová. "Moisture transport through perforated vapour retarders." Building and Environment 44 (2009): 1617-26.
[14] P. Liu, M. Justo Alonso, H.M. Mathisen, C. Simonson. "Performance of a quasi-counter-flow air-to-air membrane energy exchanger in cold
climates." Energy and Buildings 119 (2016): 129-42.
[15] J. Min, J. Duan. "Comparison of various methods for evaluating the membrane-type total heat exchanger performance." International
Journal of Heat and Mass Transfer 100 (2016): 758-66.
[16] L.Z. Zhang. "Performance Deteriorations from Flow Maldistribution in Air-to-Air Heat Exchangers: A Parallel-Plates Membrane Core
Case." Numerical Heat Transfer, Part A: Applications 56 (2009): 746-63.

You might also like