You are on page 1of 9

ECS Transactions, 75 (14) 1165-1173 (2016)

10.1149/07514.1165ecst ©The Electrochemical Society

Gas Crossover Mitigation in PEM Water Electrolysis: Hydrogen Cross-over


Benchmark Study of 3M’s Ir-NSTF Based Electrolysis Catalyst-Coated Membranes

Dmitri Bessarabova, Andries J. Krugera, Sean M Luopab, Jiyoung Parkb, Attila A.


Molnarb, and Krzysztof A. Lewinskib
a
HySA Infrastructure Center, Faculty of Engineering, North-West University, Private
Bag X6001, Potchefstroom 2520, South Africa
b
3M Company Center, Bldg. 201-01-E-21, St. Paul, MN 55144-1000, USA

In this work we present some preliminary data, such as


permeability (gas crossover) of oxygen and hydrogen as a function
of current density and other operational variables, aimed at
establishing baselines for unmitigated hydrogen crossover of 3M’s
electrolyzer MEAs based on 3M’s NSTF low-PGM loading
catalyst and several types of perfluoro sulfonic acid based PEM
membranes, both of which are widely commercially available
(such as Nafion™ membranes), and membranes made by 3M.
Experimental challenges associated with ex situ and in situ gas
crossover measurements will be also discussed.

Introduction

New trends in the development of polymer electrolyte membrane (PEM) water


electrolysis systems have opened up new technology gaps and requirements that have not
yet been addressed with respect to PEM water electrolysis. For example, as hydrogen is
now considered as one of the best solutions for large-scale energy storage (1), new
megawatt PEM electrolyzers are required. These trends are evident in new, recently
established, large-scale PEM water electrolysis installations, especially for power-to-gas
projects in Europe, and plans are underway to treat large quantities contaminated water at
the Fukushima Nuclear Power Plant.
PEM water electrolysis for hydrogen production has been in use for several years
now, without substantial improvements having been made. However, with the new focus
on hydrogen as the energy carrier, there is increasing interest in low cost and high
efficiency hydrogen production. There are two main ways in which the cost of hydrogen
production via PEM water electrolysis can be reduced: reduce the capital expenses
(CAPEX) and/or to reduce the operating expenses. For example, according to Proton
Onsite (2), analysis has already shown that increasing the number of cells in a single
stack reduces the per kW capital cost of the stack. It was also shown (2) that by
increasing the current density of each cell stack, it is possible to reduce the capital costs.
3M has recently demonstrated a very effective way to address reducing high CAPEX,
by increasing the range of current densities at which electrolyzers can operate, from a
maximum of about state-of-the-art 2.0 A cm-2 to as much as 20 A cm-2, by employing
3M’s new proprietary nanostructured thin film (NSTF) catalysts and more conductive 3M
perfluorosulfonic acid (PFSA) based electrolytes in the electrolyzer’s membrane
electrode assembly (3).

1165
Downloaded on 2016-10-25 to IP 203.64.11.45 address. Redistribution subject to ECS terms of use (see ecsdl.org/site/terms_use) unless CC License in place (see abstract).
ECS Transactions, 75 (14) 1165-1173 (2016)

What the cited work does not addresses, however, is an inherent limitation associated
with 3M’s Ir-NSTF based anode catalysts, namely their near complete inertness to
hydrogen oxidation reaction (HOR). An additional trade-off between performance and
gas crossover also exists when thinner PEM membranes are used. Thinner PEM
membranes result in high hydrogen crossover, introducing a shortcoming in the otherwise
complete list of favorable requirements for catalyst coated membranes (CCMs) that has
to be met for a membrane-electrode assembly (MEA) to be successfully employed in a
PEM water electrolyzer (4, 5). This was clearly demonstrated in the early 3M catalyst
trials where fuel-cell-derived Pt-alloy based NSTF catalysts were used for water
electrolysis. The reported durability of such electrolyzer catalysts was excellent, but
performance left room for improvement (6). The performance aspect of 3M electrolyzer
catalysts has since been addressed by switching to electrolyzer specific (and Ir based)
catalyst compositions. Such an approach has left a utility gap by eliminating possible
hydrogen crossover mitigating components from the catalyst. To address this problem,
alternative mitigation strategies have to be devised.
In summary, an area of PEM water electrolysis development of continuing interest is
the improvement of the stack performance with a simultaneous reduction in membrane
gas crossover. Gas (hydrogen and oxygen) crossover occurs in PEM water electrolysis
and is a complex function of operational conditions, such as current density, temperature,
differential pressure, etc., as well as the nature of the membranes. There is a trade-off
between achieving better performance, and an increase in gas crossover, using thinner
membranes r, and compromised durability and safety.

Gas Permeability (Gas Crossover)

Research and development in the area of hydrogen fuel cells and electrolyzers comprising
PEMs has led to an understanding of the important effect of gas permeation, which takes
place in fuel cells and electrolysis stacks, and how it affects their overall performance.
The effects of hydrogen and oxygen permeation on cell polarization, as well as the effects
of the crossover on membrane degradation, are well known (7–11). One of the first
reports of gas permeation in PEM water electrolysis was published as earlier as 1977 (11).
It was suggested that the mass transport of gases during PEM electrolysis occurs via a
diffusion mechanism. It was also suggested that the water content in the membrane phase
is a very important factor (it is the water phase that dissolves gases most of all). The
permeability was defined as the amount of gas per unit pressure difference which passes
through a unit area of a membrane of unit thickness (11). Several mass-transport
processes were identified and measured. These included: water hydrodynamic permeation,
driven by a pressure difference, and water electro-osmotic permeation, which is
proportional to the current density (11).

Definition of Permeability and Measurement Units

The measurable steady-state flux density (J) of a gas permeating through a nonporous
(dense) membrane under Fick’s law approximations can be presented in terms of a
membrane permeability coefficient (P):

= ∆p
J P
[1]
A 

1166
Downloaded on 2016-10-25 to IP 203.64.11.45 address. Redistribution subject to ECS terms of use (see ecsdl.org/site/terms_use) unless CC License in place (see abstract).
ECS Transactions, 75 (14) 1165-1173 (2016)

where P is the permeability coefficient,  is the membrane thickness, A is the membrane


area, and ∆p is the differential partial pressure (7).
For many practical applications, it is assumed that there is a linear concentration
profile of a gas in a membrane (12). As a widely accepted practical approximation, it is
assumed that the solubility of gases (S) in membranes under operating conditions follows
Henry’s law:

C=S× p [2]

where C is the gas concentration in a membrane (cm3(gas) cm–3(polymer)) and p is the


gas partial pressure. In this case, S is the solubility coefficient and it has the following
units: cm3(gas) cm–3(polymer) cm–1 Hg.
Combining equations [1] and [2] results in equation [3] for the flux gas density:

∆p
= ∆p = DS
J P
[3]
A  

where D is the apparent diffusion coefficient (units: cm2 sec–1), assumed to be a constant,
(7).
It is important to note here that the use of only one apparent diffusion coefficient for
water-saturated PFSA membranes should be treated as an assumption. Due to the
presence of water in the channels of PFSA membranes, multiple diffusion coefficients
should be considered (7). In an idealized situation, the gas permeability coefficient can be
defined from equation [3] as a product of diffusion and solubility coefficients:

P = D×S [4]

The permeability coefficient for gases in polymeric membranes can be expressed in a


variety of system units (13). In this work we report data in the “Barrer” permeability units
cm3(STP) cm sec-1 cm-2 cmHg-1 (13). Under steady-state diffusion conditions, P could
also be expressed in terms of steady-state crossover current density ( i∞ ):

nFP∆p
i∞ = [5]

the permeability coefficient, and ∆p is the gas partial pressure difference.


where n is assumed to be 2 for hydrogen and 4 for oxygen, F is the Faraday constant, P is

A comprehensive analysis of gas permeation in PFSA membranes has been described


(7). From analysis of the literature, it is considered that reliable data can be produced for
dry PFSA membranes only.
In the presence of water vapor or liquid water, PFSA membranes tend to absorb water,
which results in an increase in gas permeation and experimental results typically depend
on the experimental setup (e.g., electrochemical or concentration-driven, etc.), procedure
for membrane humidification, even duration of experiments, etc. Hydrogen permeation
as a function of relative humidity and temperature is described in literature (7). See
Figure 1.

1167
Downloaded on 2016-10-25 to IP 203.64.11.45 address. Redistribution subject to ECS terms of use (see ecsdl.org/site/terms_use) unless CC License in place (see abstract).
ECS Transactions, 75 (14) 1165-1173 (2016)

Similar relationships for nitrogen permeation as a function of relative humidity and


temperature have also been reported (14). It was concluded that gas permeability in PFSA
membranes can be divided into transport through the polymer and the water phases of the
ionomer (14). Later, a few models for gas permeation were presented that consider
different diffusional paths for the gas diffusion (8, 15, for example).

120
NRE 211 membrane
950C
100
H2 Permeability, Barrer

80

60
450C

40

20

0
0 20 40 60 80 100
RH, %

Figure 1. Hydrogen permeation in NRE 211 as a function of relative humidity and


temperature (7).

Experimental Measurements

Typical polarization curves for a 25 cm2 active membrane area cell are shown in Figure 2.

Figure 2. Typical polarization curves.

1168
Downloaded on 2016-10-25 to IP 203.64.11.45 address. Redistribution subject to ECS terms of use (see ecsdl.org/site/terms_use) unless CC License in place (see abstract).
ECS Transactions, 75 (14) 1165-1173 (2016)

EIS measurements were also conducted (Figure 3). EIS is a noninvasive method used
to investigate certain properties of a PEM electrolyzer cell. When validating gas
crossover, the ohmic resistance will be of interest. The ohmic resistance includes the
membrane resistance, which relates to membrane changes such as membrane thinning.

Figure 3. Electrochemical impedance spectroscopy (0.5 A cm-2, 1 A cm-2, 2 A cm-2).

Gas permeation in PFSA membranes can be measured in situ and ex situ. Details of
ex-situ measurements are given in (7). The real challenge is to measure gas crossover
under the conditions of a working PEM electrolysis cell. The key challenge is a dynamic
and effective transient behavior of water sorption in such a membrane system under the
condition of water electrolysis. In this paper we used gas chromatography as a detection
method.
A schematic diagram of the experimental setup to run electrolysis experiments and in
situ measurements of gas crossover is shown in Figure 4. It included the following
components: SRI 8610C GC equipped with a FID for detecting H2 in O2, using argon as
carrier gas; Bruker Scion 456 equipped with a TCD for detecting O2 in H , using high
purity H2 as carrier gas and a baseline gas. Furthermore, gas concentrations of both H2 in
O2 (anode) and O2 in H2 (cathode) were measured by in-line gas analyzers (supplied with
the Greenlight test station: Greenlight Innovation G13-290 PEM electrolyzer test station

825 EW (100 µm); anode loading: 0.5 mg cm-2, Ir; cathode loading: 0.5 mg cm-2, Pt;
with Emerald software). The following experimental details also applied. Membrane: 3M

temperature range: from ambient to 80 °C.


Additional measurements were done at 3M. Detector: Agilent 490 Micro GC, in-house
built test station based on Biologic VSP/Z-01 potentiostat with BSTR-100 booster, and
Ismatec pump for DI water flow control (Reglo CPF Digital Part # ISM321 w/Pump
Head RH1CKC). The cell was 50 cm2 from Fuel Cell Technologies, with Giner inc.
provided 50 cm3 anode flow field. Water flow rate to the cell was 75 cm3 min-1.
Permeation data (in the format of permeance) were calculated using the following
equation [7]:

Permeance =
V *Y
A∆p
[6]

1169
Downloaded on 2016-10-25 to IP 203.64.11.45 address. Redistribution subject to ECS terms of use (see ecsdl.org/site/terms_use) unless CC License in place (see abstract).
ECS Transactions, 75 (14) 1165-1173 (2016)

where V is the volume flow rate of the cathode or anode gas, Y is the volumetric fraction
of the gas obtained after GC analysis, A is the active membrane area and ∆p is the partial
pressure difference. Membrane thickness was used to calculate the permeability.

Figure 4. Schematic diagram of the experimental setup at HySA.

Effect of the Current Density

Under the conditions of electrical polarization during PEM electrolysis,


electroosmotic phenomena occur, and charge carriers (protons in the case of PEM water
electrolysis ) diffuse across the membrane together with water molecules. It is reasonable
then to assume that additional “electroosmotic” water provides additional capacity to
dissolve gases, thus resulting in an increase in permeation as a function of current density.
There is much published evidence (we cite here only the first ones) that gas permeation
during PEM electrolysis depends on the current density. Interestingly, however, there is
no consensus on the nature of the phenomena contributing to an increase in gas cross-
over. For example, one of the explanations to account for an increase in hydrogen
permeation with an increase in current density is the buildup of hydrogen oversaturation
at the cathode side. This oversaturation can reach values in the order of 20 bar at 1 A cm–
2
, as was discussed in the early 1980s by the team of researchers from Brown Boveri
Research Center, Switzerland.

shown on Figure 5. Data were produced using 3M 825 EW (100 µm) membranes with an
A typical relationship between hydrogen cross-over as a function of current density is

Ir anode and a Pt cathode. Due to the sensitivity, data are reported in dimensionless
normalized format. Data showed a strong relationship between current density and H2
crossover. The calculations were done by using equation [6], water partial pressure
difference was calculated based on the temperature of experiments. Change in the
membrane thickness due to the water uptake was not accounted for as membranes were
assumed to be compressed and sandwiched between electrodes. Electrolysis was
conducted for 2 hrs for each gas sample. We are aware that optimal duration of the
electrolysis experiments for gas sampling is a not known at that stage and is subject of a

1170
Downloaded on 2016-10-25 to IP 203.64.11.45 address. Redistribution subject to ECS terms of use (see ecsdl.org/site/terms_use) unless CC License in place (see abstract).
ECS Transactions, 75 (14) 1165-1173 (2016)

additional research. The currents were not run in a reverse order for this set of data.
However, the observed trend is evident.

Normalised Hydrogen Permeability


0.9
0.8 50 deg C
0.7
0.6
0.5
0.4
0.3
0.2
0.1
0
0 1 2 3 4 5 6
c/d, Amps*cm-2

Figure 5. Preliminary hydrogen crossover test results of 3M NSTF-based high-


performance CCM at HySA. Crossover data is presented in dimensionless normalized
format.

Figure 6 shows the percentage of oxygen in hydrogen at 50° C for a 150 cm2 short
stack.

Figure 6. Percentage of oxygen in hydrogen at 50 °C for 150 cm2 short stack.

1171
Downloaded on 2016-10-25 to IP 203.64.11.45 address. Redistribution subject to ECS terms of use (see ecsdl.org/site/terms_use) unless CC License in place (see abstract).
ECS Transactions, 75 (14) 1165-1173 (2016)

We also conducted ex-situ permeation measurements with the same membranes to


confirm our method. As described in (7), a single permeation cell was used for
experiments. In fact, we used a single electrolysis cell. The 3M CCM was dried over 48
hrs by passing nitrogen (sweep gas). Hydrogen was fed to the anode and nitrogen was fed
to the cathode side as a sweep gas at 25 °C. Flow rates of both hydrogen and nitrogen
were measured. Gas chromatography was used for gas detection in the sweep gas as
described in (7). The flow rate of hydrogen was approximately 10 times more than
nitrogen sweep gas. The permeability values were reported in Barres and were measured
in the range of just a few Barrers. This is in line with the classic and well-trusted data for
dry PFSA membranes reported in (16).

Conclusions

Gas permeation in PFSA membranes can be measured ex situ as a function of


temperature, relative humidity, and differential pressure, and data can be reported as
permeability (or permeance). Under the conditions of electrical polarization during PEM
electrolysis, measurements (especially on the stack level) are complicated by
electroosmotic water transport and dual-phase gas–water flow, occurring at both the
anode and cathode (especially at the anode side). This presents a challenge in establishing
the gas partial pressure difference required for the accurate calculation of permeability.
Furthermore, it is likely that recombination of hydrogen takes place, which would result
in only apparent (not true) gas crossover values being measured. Further research into
method development and data interpretation is clearly requited. It appears that currently
(with current developments), data can be produced for comparison purposes only; further
developments are required to report data in standard permeation units.

Acknowledgments

This work was supported by the DST HySA Infrastructure KP5 program. We thank
Jan-Hendrik van der Merwe for the electrochemical measurements.

References

1. D. Bessarabov, in PEM Electrolysis for Hydrogen Production: Principles and


Applications, D. Bessarabov, H. Wang, H. Li and N. Zhao, Editors, p. vii, CRC
Press, New York (2015).
2. M. Schiller and E. Anderson, Gas for Energy, 1, 44-47, https://www.gas-for-
energy.com/ (2014).
3. K. A. Lewinski and S. M. Luopa, (Abstract 1948, invited), Spring ECS Meeting,
Chicago, IL, May 24-28 (2015).
4. K. A. Lewinski, D. van der Vliet, S. M. Luopa, (Abstract 1457), Fall ECS
Meeting, Phoenix, AZ, Oct 11-15 (2015).
5. K. A. Lewinski, D. van der Vliet, S. M. Luopa, ECS Trans., 69(17), 893 (2015).
6. M. K. Debe, S. M. Hendricks, G. D. Vernstrom, M. Meyers, M. Brostrom, M.
Stephens, Q. Chan, J. Willey, M. Hamden, C. K. Mittelsteadt, C. B. Capuano, K.
E. Ayers, E. B. Anderson, J. Electrochem. Soc., 159(6), K165 (2012).

1172
Downloaded on 2016-10-25 to IP 203.64.11.45 address. Redistribution subject to ECS terms of use (see ecsdl.org/site/terms_use) unless CC License in place (see abstract).
ECS Transactions, 75 (14) 1165-1173 (2016)

7. D. Bessarabov, in PEM Fuel Cell Diagnostic Tools, H. Wang, Xiao-Zi Yuan, Hui
Li, Editors, Chapter 21, p.443, CRC Press, New York (2011).
8. M. Schalenbach, M.A. Hoeh, J.T. Gostick, W. Lueke, D. Stolten, J. Phys. Chem.
C, 119, 25156 (2015).
9. J. Zhang, Y. Tang, Y., C. Song, H. Wang, J. Power Sources, 163, 532 (2006).
10. C. K. Mittelsteadt and H. Liu, In Handbook of fuel cells, W. Vielstich, H.A.
Gasteiger, A. Lamm, Editors, John Wiley and Sons, Chapter 23, vol. 5., p. 345
(2009).
11. A. B. Laconti, A. R. Fragala, J. R. Boyack, In: Symposium on Electrode Materials
and Processes for Energy Conversion and Storage, Philadelphia, Pa., May 9-12,
Proceedings (A79-11776 02-25), Princeton, N.J., Electrochemical Society, p. 354
(1977).
12. R. Ash, R. M. Barrer, D. G. Palmer, Br. J. Appl. Phys. 16, 873 (1965).
13. S.A. Stern, J. Polymer Sci., A-2: Polymer Physics, 6(11), 1933 (1968)
14. C. Mittelsteadt, M. Umbrell, (Abstract #770), 207th ECS Meeting, Quebec City,
Canada, (2005).
15. R. Datta, D.J. Martino, Y. Dong and P. Choi, in PEM Electrolysis for Hydrogen
Production: Principles and Applications, D. Bessarabov, H. Wang, H. Li, and N.
Zhao, Editors, Chapter 12, p. 243, CRC Press, New York (2015).
16. J.S. Chiou and D.R. Paul, I&EC Research, 27, 2161 (1988).

1173
Downloaded on 2016-10-25 to IP 203.64.11.45 address. Redistribution subject to ECS terms of use (see ecsdl.org/site/terms_use) unless CC License in place (see abstract).

You might also like