0% found this document useful (0 votes)
21 views41 pages

Basisc of Electrolysers

Uploaded by

Aravind Shankar
Copyright
© © All Rights Reserved
We take content rights seriously. If you suspect this is your content, claim it here.
Available Formats
Download as PDF, TXT or read online on Scribd
0% found this document useful (0 votes)
21 views41 pages

Basisc of Electrolysers

Uploaded by

Aravind Shankar
Copyright
© © All Rights Reserved
We take content rights seriously. If you suspect this is your content, claim it here.
Available Formats
Download as PDF, TXT or read online on Scribd
You are on page 1/ 41

Effects of various parameters of different porous transport layers in

proton exchange membrane water electrolysis

Zhenye Kang*, Shaun M. Alia, James L. Young, Guido Bender

Chemistry and Nanoscience Center


National Renewable Energy Laboratory (NREL)
Golden, CO 80401, USA

* Corresponding author. Tel: +1-303-275-4199; Email address: [email protected]


1

Pursuant to the DOE Public Access Plan, this document represents the authors' peer-reviewed, accepted manuscript.
The published version of the article is available from the relevant publisher.
Highlights:

• A general electrolyzer model was used to deconvolute overpotential contributions

• Electrolyzer anode PTLs with different properties were experimentally tested to validate

the model

• The effects of the mass transport in anode PTLs were characterized

• Effects of different hydrophobic additives were quantitively investigated

Pursuant to the DOE Public Access Plan, this document represents the authors' peer-reviewed, accepted manuscript.
The published version of the article is available from the relevant publisher.
Graphical Abstract

Pursuant to the DOE Public Access Plan, this document represents the authors' peer-reviewed, accepted manuscript.
The published version of the article is available from the relevant publisher.
Abstract

Porous transport layers (PTLs) play an important role in proton exchange membrane water

electrolysis (PEMWE) cells. The PTL facilitates water and gas transport, as well as thermal and

electrical conduction, and is required to sustain good contact with adjacent components. It is

expected that using PTLs with variations in material properties such as structure, composition,

thickness and wettability results in performance changes of the PEMWE. In this study, a general

mathematical PEMWE model is developed that separates and analyzes the contributions of ohmic,

activation, diffusion and Nernst potentials. For model validation, three inherently different anode

PTL structures (carbon paper, sintered titanium particles, and titanium felt) are operated over a

range of conditions. Additionally, the effects of PTL wettability were used to verify the model

using Polytetrafluoroethylene (PTFE) treated Toray papers with PTFE loading ranging from 0%-

20%. The modeling results of both PTFE treated and untreated materials show good agreement

with the experimental data. Mass transport or diffusion loss is the primary reason for performance

differences between PTFE treated and untreated PTLs. Sintered titanium PTLs with thicknesses

above 1 mm suffer from up to 33% increased ohmic losses without indicating any obvious changes

in activation and diffusion losses when compared to untreated PTLs. The losses of the cell increase

when using PTFE treated Toray paper. Individual contributions are quantified and assigned to

increased ohmic, activation, and diffusion losses. In conclusion, the proposed model offers insights

into the overpotential contributions of a PEMWE. It is a useful tool for predicting performance of

various PTL materials and can be applied for PTL development and optimization efforts.

Key Words: porous transport layer; overpotential; loss separation; mass transport; proton

exchange membrane water electrolysis.

Pursuant to the DOE Public Access Plan, this document represents the authors' peer-reviewed, accepted manuscript.
The published version of the article is available from the relevant publisher.
1. Introduction

Hydrogen has been considered as an ideal energy carrier to store renewable energy. Proton

exchange membrane water electrolysis (PEMWE) as a means for hydrogen production offers high

product purity, fast load response times, small footprints, high-efficiencies, and low maintenance

efforts. It is regarded as a promising technology especially when coupled with renewable energy

sources [1-7]. Although the overwhelming majority of hydrogen today is produced from steam

methane reforming, electrolysis is expected to see highly increased deployment with the increased

use of renewables and electrolysis cost reduction efforts [8, 9]. At reduced feedstock costs,

electrolyzer capital cost becomes dominant, and understanding and mitigating component losses

of for example porous transport layers (PTLs), catalysts, membranes, or separators is critical to

advance the technology while reducing cost [10, 11]. PTLs, sometimes alternatively referred to as

liquid/gas diffusion layers (LGDLs), are one of the critical components in PEMWE cells [12-15].

In the cell assembly, PTLs are located between the flow field and the catalyst layer. They provide

electrical conduction and water/gas transport and should have appropriate electrical/thermal

conductivity, mechanical strength, water/gas transport properties, and corrosion resistance, while

maintaining interfacial contact with their adjacent components [16-19].

Various PTL types have been utilized and studied in the reported electrolyzer research. These

include titanium-based felts [20-23], meshes [24, 25], foams [21, 26], sintered particles [20, 27-

29], and specifically engineered thin materials [30-36]. For short term operation conventional

carbon paper can also be employed as anode PTLs [15, 37-40]. This however, requires to closely

monitor the operating conditions and time, because the carbon material is unstable and corrodes

quickly at anode potentials larger than 1.23 V vs Standard Hydrogen Electrode (SHE). For

example, Ito et al. investigated various Ti felts with different porosities and mean pore diameters

Pursuant to the DOE Public Access Plan, this document represents the authors' peer-reviewed, accepted manuscript.
The published version of the article is available from the relevant publisher.
for utilization in PEMWE. They found that cell performance improved with decreasing mean pore

diameter down to 10 μm, and that porosities greater than 50% did not significantly influence

performance [41]. Liu et al. introduced an iridium coating on Ti felt PTLs, which improved

performance by reducing the PTL’s interfacial contact resistance [23]. Grigoriev et al. investigated

sintered titanium particle PTLs and found that the mean pore size of the particles and the thickness

of the titanium plates have a significant effect on PEMWE performance. They observed an

optimum sphere particle size between 50 to 75 μm and an optimum pore size between 12 and

13 μm and found that varying porosity between 0.35 and 0.40 did not have a significant influence

on electrolysis efficiency [42]. Lettenmeier et al. developed PTLs with pore gradients by thermally

spraying titanium particles with different sizes. These PTLs achieved a performance which was

comparable to state-of-the-art sintered plates and increased with respect to mesh PTLs. The results

were attributed to reduced mass transport losses at high current density [12]. Hwang et al. studied

the effects of PTFE loaded GDLs on the performance of PEM unitized regenerative fuel cells

(URFCs). They found that PTFE loading can improve the fuel cell performance at dry and fully

wet condition, while it deteriorates the performance at intermediate relative humidity [43, 44]. In

another study, they also showed the effects of PTFE loadings on PEM URFC performance in

electrolysis mode and limited effects were found for operation below 1.0 A cm-2 [45]. Sadeghifar

et al. investigated the effects of PTFE loading on thermal conductivity of the carbon GDLs, and

they found that increasing PTFE loading reduces GDL thermal conductivity and increases

interfacial contact resistance, with neither effect being beneficial [46].

Experimental research can generally benefit from the application of modeling which can increase

understanding and is furthermore a promising method to optimize PEMEC designs and operating

conditions while saving time and cost [47]. Most electrolyzer cell models developed in the past

Pursuant to the DOE Public Access Plan, this document represents the authors' peer-reviewed, accepted manuscript.
The published version of the article is available from the relevant publisher.
decade focused on analyzing voltage losses [48-50], optimizing components [32, 47, 51],

structures [52, 53], investigating mass transport [54, 55] and studying effects of operating

conditions [50, 56, 57]. A significant amount of modeling work has been published that focuses

on the effects within PTLs. Recently, for example, Zlobinski et al. used neutron-imaging

techniques and took a closer look at water and gas distribution profiles at various current densities

within anode PTLs. The found that the two-phase flow in PTLs is purely capillary driven for a

wide range of operating conditions. Pressure and current density showed no significant impact on

the water and gas distribution inside the porous media [58]. Schuler et al. investigated the bulk,

surface and transport properties of Ti felt PTL materials, and correlated the Ti felt structures to the

PEMEC performance. They modeled thermal processes and transport loss in anode PTLs and

deduced design guidelines for better performing PTLs [59, 60]. Lee et al. developed a stochastic

modelling technique for simulating sintered titanium powder-based PTLs, and they investigated

the Ti packing density and the filling radius in this study. They also used this model to study the

impact of the PTL’s powder diameter and the impact of porosity on reactant transport and PTL-

catalyst layer (CL) interfacial contact [61, 62]. Most of the work in the literature focuses on a

single PTL material each, and the work cannot be directly applied to other structural PTL materials.

In this work, we present a mathematical model that allows the separation of overpotentials in cells

operated with various PTL materials of different properties. Various common PTL types were

employed for its development including PTLs that were exposed to hydrophobic treatments. The

latter served validation purposes as it was expected that such treatments diminish the transport

properties of the PTL and reduce cell performance. Similar materials are used in fuel cells for

performance improvement [63-65]. In contrast to PEMWE cells, fuel cells operate in a gas rich

environment and wettability of gas diffusion media (GDM) has a large effect. Reported studies

Pursuant to the DOE Public Access Plan, this document represents the authors' peer-reviewed, accepted manuscript.
The published version of the article is available from the relevant publisher.
present, for example, the effects on water distribution [66], PTFE loading impacts on GDM

properties [46] and flow properties [67], and cell performance [43, 44].

In this work, a general PEMWE model is presented that enables performance predictions of

different anode PTLs. The model is developed using performance data from cells operated with

various PTL materials and subsequently validated by predicting and measuring the performance

effects of PTLs with hydrophobic treatments. PTLs with different structures, including Toray

paper (5% PTFE treated and untreated), sintered titanium particles, and Ti felt, are experimentally

investigated under different temperatures and pressures in a 5 cm2 PEMWE single cell.

Polarization curves up to 3.0 A cm-2 are conducted in order to observe mass transport effects. The

performance differences between each PTL is analyzed utilizing Vi curves and electrochemical

impedance spectroscopy (EIS). The results are used to determine model parameters. Then, the

developed model is employed to predict performance of a Toray paper with 20% PTFE loading

and subsequently validated through experiments. Results of all PTLs are discussed with regards to

the effects of different PTL parameters, diffusion losses, and overall PEMWE performance.

2. Model development

In order to investigate and predict the effects of different PTL parameters on both performance

and mass transport, a general PEMWE model was established using a MATLAB/Simulink

platform. The total cell voltage VPEMEC is divided into four parts: (i) the Nernst voltage 𝑉𝑁𝑒𝑟𝑛𝑠𝑡 ,

which is calculated using the Nernst equation, (ii) the activation overpotential 𝜂𝑎𝑐𝑡 , (iii) the ohmic

overpotential 𝜂𝑂ℎ𝑚 , and (iv) the diffusion overpotential 𝜂𝑑𝑖𝑓𝑓 , as shown in equation (1) [16].

𝑉𝑃𝐸𝑀𝐸𝐶 = 𝑉𝑁𝑒𝑟𝑛𝑠𝑡 + 𝜂𝑎𝑐𝑡 + 𝜂𝑂ℎ𝑚 + 𝜂𝑑𝑖𝑓𝑓 (1)

Pursuant to the DOE Public Access Plan, this document represents the authors' peer-reviewed, accepted manuscript.
The published version of the article is available from the relevant publisher.
Each of these contributions to the total cell voltage is individually calculated in the following four

sections.

2.1 Nernst voltage

The Nernst voltage is the theoretical minimum voltage for PEM electrolyzer cells when neglecting

other overpotentials. It can be calculated from the Nernst equation or Gibbs free energy as given

by equation (2) [16],

0.5 0.5
𝑅𝑇 𝛼𝐻2 𝛼𝑂 𝑅𝑇 𝛼𝐻2 𝛼𝑂
𝑉𝑁𝑒𝑟𝑛𝑠𝑡 = 𝑉0 + ln ( 2
) = 1.229 − (9 × 10−4 ) × (𝑇 − 298.0) + ln( 2
) (2)
2𝐹 𝛼𝐻2 𝑂 2𝐹 𝛼𝐻2 𝑂

𝑉0 is the reversible voltage, R is the ideal gas constant with a value of 8.314 J K-1 mol-1, F is the

Faraday constant with a value of 96485 C mol-1, T is the operating temperature in Kelvin (K), and

𝛼𝐻2 and 𝛼𝑂2 are the activity of ideal hydrogen and oxygen gas, respectively, which can be

calculated from the hydrogen and oxygen partial pressures, respectively [16].

2.2 Activation overpotential

The activation overpotential of the cell describes the loss due to the electrochemical reaction. It

represents the overpotential to drive the electron transfer and electrochemical reaction kinetics as

described by the Butler-Volmer model. It is calculated using equation (3) [68].

𝑅𝑇𝑎𝑛 𝑖 𝑅𝑇𝑐𝑎𝑡 𝑖
𝜂𝑎𝑐𝑡 = 𝜂𝑎𝑐𝑡,𝑎𝑛 + 𝜂𝑎𝑐𝑡,𝑐𝑎𝑡 = 𝛼 arcsinh (2𝑠𝑖 )+𝛼 arcsinh⁡(2𝑖 ) (3)
𝑎𝑛 𝐹 0,𝑎𝑛 𝑐𝑎𝑡 𝐹 0,𝑐𝑎𝑡

In this equation, 𝑇𝑎𝑛 and 𝑇𝑐𝑎𝑡 are the anode and cathode temperatures, respectively, and s is the

liquid saturation at the interface between CL and PTL, which is defined as the ratio of liquid water

volume to total pore volume in PTLs [55]. For the anode and cathode charge transfer coefficients,

values of 𝛼𝑎𝑛 =2.0 and 𝛼𝑐𝑎𝑡 = 1.5 were used, respectively. Finally, 𝑖0,𝑎𝑛 and 𝑖0,𝑐𝑎𝑡 are the

exchange current densities of anode and cathode, respectively. Their values are greatly affected by

Pursuant to the DOE Public Access Plan, this document represents the authors' peer-reviewed, accepted manuscript.
The published version of the article is available from the relevant publisher.
(i) the activity of the catalyst materials and (ii) the operating temperature [69], and can be

calculated as shown in equations (4) and (5):

𝑖0,𝑎𝑛 = 𝑖𝑎𝑛0 ∙ 𝛾𝑎𝑛 40 ∙ (1 − 𝜏)3 (4)

𝑖0,𝑐𝑎𝑡 = 𝑖𝑐𝑎𝑡0 ∙ 𝛾𝑐𝑎𝑡 8 ∙ (1 − 𝜏)3 (5)

𝑖𝑎𝑛0 and 𝑖𝑐𝑎𝑡0 are the initial reference exchange current densities at 80°C. These values are catalyst

specific and were fitted for our samples to be 1.5·10-8 A cm-2 and 5.0·10-2 A cm-2 for the anode

(IrO2) and cathode (Pt/C), respectively. All used values in this study are shown in Table 3. 𝛾 is a

coefficient related to temperature, which can be calculated using Equation (6) and (7). The additive

coverage ratio 𝜏 is a parameter that represents how much surface area is covered by the additive,

and it is given by equation (8).

𝑎𝑛 𝑇
𝛾𝑎𝑛 = 353.15 (6)

𝑇𝑐𝑎𝑡
𝛾𝑐𝑎𝑡 = (7)
353.15

𝜌𝑏𝑎𝑠𝑒∙𝜒𝑎𝑑𝑑𝑖𝑡𝑖𝑣𝑒 0.8
𝜏 = [𝜌 ] (8)
𝑎𝑑𝑑𝑖𝑡𝑖𝑣𝑒 ∙(1−𝜒𝑎𝑑𝑑𝑖𝑡𝑖𝑣𝑒 )

2.3 Ohmic overpotential

The ohmic overpotential describes all losses that are related to charge transport of electrons and

ions in the cell. In this model, the ohmic overpotential is calculated using a simplified electrical

circuit of the PEMWE cell, which contains the resistances from each component (bipolar plates,

PTLs, CLs, PEM) and interfacial contact resistances between each component. The individual

component resistances are based on literature results [70, 71], and their values can be found in

Table 3. In this study, it is assumed that the interfacial contact resistance between anode PTL and

10

Pursuant to the DOE Public Access Plan, this document represents the authors' peer-reviewed, accepted manuscript.
The published version of the article is available from the relevant publisher.
CL is the main contributor to the interfacial contact resistance, and it can be calculated using

equation (9) [47].

𝑅𝑖𝑛𝑡𝑒𝑟𝑓𝑎𝑐𝑖𝑎𝑙 = 𝐾 ∙ 𝜀/(1 − 𝜏)3 (9)

K is the interfacial contact resistance coefficient, which is a fitted value and equal to 12.5 mΩ cm2,

80 mΩ cm2 and 45 mΩ cm2 for Toray paper, sintered Ti PTLs and Ti felt, respectively. 𝜀 is the

calculated porosity of PTLs and 𝜏 is the coverage ratio of the nonconductive additives in the PTL.

The PTL porosity is expected to change when a wettability treatment is added. It can be estimated

based on the initial density of the PTLs without any treatment 𝜌𝑃𝑇𝐿,0 , the density of the base

materials for the PTLs, 𝜌𝑏𝑎𝑠𝑒 (2.0 g cm-3 for the carbon fiber in the Toray paper, 4.506 g cm-3 for

Ti bulk material in the sintered Ti PTLs and Ti felt), the additive material density 𝜌𝑎𝑑𝑑𝑖𝑡𝑖𝑣𝑒 (2.16

g cm-3 for PTFE) and the additive loadings 𝜒𝑎𝑑𝑑𝑖𝑡𝑖𝑣𝑒 from the treatment with weight percentage,

which is represented by equation (10) [72].

𝜌𝑃𝑇𝐿,0 𝜌𝑃𝑇𝐿,0
𝜀 =1− −𝜌 ∙ 𝜒𝑎𝑑𝑑𝑖𝑡𝑖𝑣𝑒 (10)
𝜌𝑏𝑎𝑠𝑒 𝑎𝑑𝑑𝑖𝑡𝑖𝑣𝑒

Based on the property of the PTLs and its additive treatment, the equilibrium contact angle (static

sessile-drop contact angle) of the PTLs can also be calculated based on a modified form of the

Cassie equation, that includes a term for considering porosity [72, 73] as shown in equation (11),

cos 𝜃𝑒𝑞 = (1 − 𝜏) ∙ cos 𝜃𝑏𝑎𝑠𝑒 + 𝜏 ∙ cos 𝜃𝑎𝑑𝑑𝑖𝑡𝑖𝑣𝑒 − 𝜀 (11)

where 𝜃𝑏𝑎𝑠𝑒 and 𝜃𝑎𝑑𝑑𝑖𝑡𝑖𝑣𝑒 is the contact angle of the solid base material and additive material,

respectively, which equals to 80° for the carbon material, 45° for Ti material, 108° for PTFE, and

100° for FEP [73].

2.4 Diffusion overpotential

11

Pursuant to the DOE Public Access Plan, this document represents the authors' peer-reviewed, accepted manuscript.
The published version of the article is available from the relevant publisher.
Finally, the diffusion overpotential represents the effects of mass transport processes in the cell. It

is calculated using the following equation [16].

𝑅𝑇𝑎𝑛 𝐶 𝑅𝑇𝑐𝑎𝑡 𝐶
𝜂𝑑𝑖𝑓𝑓 = ln (𝐶 𝑂2 ) + ln (𝐶 𝐻2 ) + 𝜁𝑖 2 (𝜃𝑒𝑞 − 𝜃0 ) (12)
2𝐹 𝑂2 ,0 2𝐹 𝐻2 ,0

𝐶𝑂2 and 𝐶𝐻2 are the oxygen and hydrogen concentrations at the interface of CL and PEM,

respectively. 𝐶𝑂2 ,0 and 𝐶𝐻2,0 indicate the reference values. 𝜁 is a coefficient with a fitted value of

0.47 for all the cases, and 𝜃0 is the equilibrium contact angle of the untreated PTLs.

3. Experimental details

The cells used in this work had 5 cm2 active areas and were assembled using identical catalyst

coated membranes (CCMs) and cathode PTL materials while varying the anode PTLs. All PTLs

were cut with a cutting die to an exact 5 cm2 geometry. The material properties of the sample cells

are given in Table 1. Four different PTLs, including (i) 5% PTFE treated Toray paper (TGP-H-

120), (ii) 0% PTFE (untreated) Toray paper 120 from Fuel Cell Earth LLC, (iii) sintered Ti particle

plates (porous titanium sheet PT22-S) from Edgetech Industries LLC, and (iv) Ti felt (2GDL10-

0,35) from Bekaert, were used as anode PTLs and the results employed to fit the model parameters.

A fifth PTL, 20% PTFE treated Toray paper also from Fuel Cell Earth LLC was subsequently used

to verify the model’s ability to predict cell performance. Although carbon paper cannot be used as

anode PTLs in electrolyzer systems that are operated over extended time periods, it was used in

this study for short term operation to validate the model with an additional data point.

Figure 1 shows scanning electron microscope (SEM) images of the PTLs. It is apparent when

comparing the untreated Toray paper shown in Figure 1 (A) with the 5% PTFE treated Toray paper

12

Pursuant to the DOE Public Access Plan, this document represents the authors' peer-reviewed, accepted manuscript.
The published version of the article is available from the relevant publisher.
in Figure 1 (B) that a binder material is used in the untreated Toray paper and that the porosity is

additionally reduced by the PTFE treatment. The binder materials specifically accumulate at the

carbon-fiber cross-sections. The fluorine content of 0% PTFE (non PTFE treated), and 5% and 20%

PTFE treated Toray paper was mapped with EDS and the results are shown in Figure 1 (E). The

fluorine signal from the 0% PTFE treated Toray paper shows a similar intensity than a carbon tape

background (not shown in the figure), which allows for the conclusion that the binder material in

untreated Toray paper does not contain any fluorine. The 5% and 20% PTFE treated Toray papers

exhibit proportionately increasing fluorine intensities and the fluorine appears to be more

uniformly distributed at the 20% PTFE loading. During cell assembly the Toray papers are

compressed by approximately 13% which warrants a good electrical contact while maintaining

high porosity.

SEM images of the Ti based sintered particles and felt PTL materials used are shown in Figure 1 (C)

and 1 (D), respectively. Both have a distinct morphology and compared to Toray paper

significantly lower porosities (see Table 1). They were likely transformed into a single Ti based

porous layer by using a high pressure and temperature process. When assembled in the PEMWE

cell, it is assumed that the Ti PTLs maintain their thickness and cannot be compressed due to their

rigid, metal-based structure. Generally, the Ti felt that was used had a higher porosity than the

sintered Ti material, and a structure and thickness similar to the Toray papers.

Table 1. Details of cell setups with different PTLs


Model Development Model Prediction & Validation
5% PTFE
Treated Untreated Sintered Ti 20% PTFE Treated Toray
Anode PTLs Ti Felt
Toray Paper Toray Paper Particles Paper TGP-H-120
TGP-H-120

13

Pursuant to the DOE Public Access Plan, this document represents the authors' peer-reviewed, accepted manuscript.
The published version of the article is available from the relevant publisher.
~351 μm ~350 μm ~1080 μm ~350 μm
Anode PTL ~351 μm thickness
thickness thickness thickness thickness
Properties 78% porosity
78% porosity 78% porosity 30-40% porosity 59% porosity
Anode Gasket
~305 μm ~305 μm ~1067 μm ~330 ~305 μm
Thickness
Anode PTL
~13.10% ~12.90% ~0% TBD** ~13.10%
Compression
Anode: 2.26 Anode: 2.37 Anode: 2.26 mg Anode: 2.36
mg Ir/cm2 mg Ir/cm2 Ir/cm2 mg Ir/cm2 Anode: 2.32 mg Ir/cm2
JULICH IEK-3
Cathode: 0.74 Cathode: 0.79 Cathode: 0.76 Cathode: 0.78 Cathode: 0.74 mg Pt/cm2
CCM
mg Pt/cm2 mg Pt/cm2 mg Pt/cm2 mg Pt/cm2 N117
N117 N117 N117 N117
5% PTFE Treated Toray Paper
Cathode PTL 5% PTFE Treated Toray Paper
~351 μm thickness with 78%
and properties ~351 μm thickness with 78% porosity
porosity
Cathode Gasket
~305 μm ~305 μm
Thickness
Cathode PTL
~13% ~13% ~17% TBD** ~13%
Compression
**Since the Ti felt is somewhat compressible and its relation between elasticity/plasticity behavior
and ranges are unknown at this point in time, the PTL compression is also unknown.

Figure 1. SEM images of (A) untreated Toray paper; (B) 5% PTFE treated Toray paper,

(C) Sinter Ti PTLs, (D) Ti felt, and (E) EDX mapping of Fluorine for Toray papers with

different PTFE loadings

14

Pursuant to the DOE Public Access Plan, this document represents the authors' peer-reviewed, accepted manuscript.
The published version of the article is available from the relevant publisher.
CCMs with IrO2 anodes (2.3 mgIr cm-2), Pt/C cathodes (0.8 mgPt cm-2), and Nafion 117

membranes were provided by Forschungszentrum Jülich (FZJ), Germany. The detailed fabrication

procedure can be found in the literature [23], and the specific sample details are given in Table 1.

Toray paper TGP-H-120 with 5% PTFE treatment and measured 351-μm thickness and nominal

78% porosity was employed as cathode PTLs for all tests. Graphite bipolar plates (BPs) with single

serpentine flow channels and an active area of 5 cm2 were used at both anode and cathode. DI

water, preheated to the operating temperature at the inlets of the cell, was supplied in co-flow

configuration at a rate of 50 ml/min to the anode and cathode cell inlet. Additionally, the cell

hardware was temperature controlled to the operating temperature using the anode BP as control

point by two heating pads at anode and cathode endplates. Therefore, we can safely assume that

the internal cell temperature was well controlled at the set value although various anode PTLs with

different thermal conductivities were used. The electrolyzer cell was assembled with compression

delivered by eight bolts each tightened to a torque of 4.52 Nm. The anode and cathode outlets were

each individually backpressure controlled. Both polarization curves and EIS were conducted using

a Gamry potentiostat (Reference 3000) with a 30 A booster system. The frequency range of the

EIS measurements were 100 kHz to 100 mHz using a 5% perturbation of the DC current.

An established test protocol used in previous benchmarking efforts was used for the experiments

[15]. For startup, DI water was supplied to both anode and cathode at 80°C for about 1 hour to

bring the cell to operating temperature. Cell conditioning consisted of three steps: (i) a

galvanostatic hold at 0.2 A cm-2 for 0.5 hour, (ii) a galvanostatic hold at 1.0 A cm-2 for 0.5 hour,

and (iii) a potentiostatic hold at 1.7 V for 8 hours. After the conditioning, polarization curves were

conducted using galvanic control and stepping from 0.01 A cm-2 to 3.0 A cm-2 in 19 equal 5-

minute long steps with a 2.5-V limit for both up-scan and down-scan. Step sizes were 25 mA cm-

15

Pursuant to the DOE Public Access Plan, this document represents the authors' peer-reviewed, accepted manuscript.
The published version of the article is available from the relevant publisher.
2 between 25 mA cm-2 - 200 mA cm-2, 250 mA cm-2 between 250 mA cm-2 – 2 A cm-2, and finally

500 mA cm-2 between 2 - 3 A cm-2. Cell performance was obtained by averaging the last 1-minute

data of each step from both up-scan and down-scan. EIS was conducted within the up-scan at the

end of each current step up to a current density of 1.6 A cm-2, i.e. the maximum current limit of

the instrument. Cell temperature was first set to 80°C, then 60°C applying a range of backpressures

up to 1.72 bar (25 psi) including local ambient pressure (0.88 bar) and simulated sea level ambient

pressure (1.013 bar) at each temperature.

4. Results and discussions

In this section, we discuss the experimental results first. Next, the developed model is validated by

using the experimental data and then used to visualize the voltage losses. At last, we use the model

to predict the performances with different PTLs under various conditions.

4.1 Experimental investigation of different PTLs

Performances of four different anode PTLs are presented and discussed, and Nyquist EIS plot

fittings are utilized to analyze the effects of different anode PTLs. Figure 2 (A) and (B) compare

the performances achieved with four different anode PTLs at both 60 and 80°C, respectively. The

performance progression with current density of the four cells varies. While the cells with the

untreated Toray paper (blue) and the Ti felt (green) samples show a linear progression without

seeing transport effects, the sintered particle material (red) shows a slight decrease in ohmic loss

with increasing current densities and the 5% PTFE (black) sample an increase. For example, at

60°C the performance of the 5% PTFE (black) is better than that of the sintered sample (red).

Performances diverge at current densities above 250 mA cm-2 and converge again at about

2.8 A cm-2. While the performance of the cell with the sintered material (red) seems to improve,

16

Pursuant to the DOE Public Access Plan, this document represents the authors' peer-reviewed, accepted manuscript.
The published version of the article is available from the relevant publisher.
that of the 5% PTFE (black) sample deteriorates. At 80°C this effect is less pronounced. The two

samples have a similar performance at current densities up to 2 A cm-2. At higher current densities

the performance of the 5% Toray paper (black) decreases and diverges at about 2.25 A cm-2 from

the performance of the sintered Ti particles (red), which seems to progress linearly at this

temperature. At both temperatures, the Ti felt PTL (green) achieves a better performance than the

sintered Ti PTL (red), and a lower performance than the untreated Toray paper (blue), which shows

the highest performance from all samples.

The HFR-free polarization curves and HFR values allow insights into the processes that may cause

these performance trends. The HFR values were extracted from the EIS data and used to compute

the HFR-free polarization curves up to 1.6 A cm-2 in Figure 2 by using equation (13).

𝑉𝐻𝐹𝑅−𝑓𝑟𝑒𝑒 = 𝑉𝑃𝐸𝑀𝐸𝐶 − 𝑖 ∙ 𝐻𝐹𝑅 (13)

At both temperatures, the PTLs, with exception of the 5% PTFE treated Toray paper (black), have

very similar HFR-free performances indicating that their performance differences are mainly due

to ohmic losses. The HFR-free performance of the 5% PTFE (black) sample however is lower

(higher cell voltage) at both temperatures. This effect is pronounced with current density. The

sintered Ti particle PTL (red) has the highest HFR and its value decreases with current density, for

example at 60°C from about 246 m cm2 at 200 mA cm-2 to about 230 m cm2 at 1.6 A cm-2. The

HFR value of the Ti felt (green) sample also decreases though much less significantly (60°C:

190 m cm2 at 200 mA cm-2 to 186 m cm2 at 1.6 A cm-2). The HFR for untreated Toray paper

(blue) is 173 m cm2 at 60°C and independent of current density, as expected, and that of the 5%

PTFE treated Toray paper (black) increases (60°C: 182 m cm2 at 200 mA cm-2 to 192 m cm2

at 1.6 A cm-2).

17

Pursuant to the DOE Public Access Plan, this document represents the authors' peer-reviewed, accepted manuscript.
The published version of the article is available from the relevant publisher.
The origin of the HFR decrease with increasing currents observed for both Ti PTLs (green, red) is

subject to a detailed study which will be published in the near future. We currently speculate that

the phenomenon is associated with a local temperature increase which may improve the proton

conductivity of the ionomer in the electrode. The increase of HFR and decrease in HFR-free

performance observed with the 5% treated PTL (black), however, is likely related to the local water

content and the water/oxygen transport [74]. At high current densities, both the water consumption

by the reaction and by the electroosmotic drag are increased while water transport through the 5%

treated PTL may be hindered significantly more than in the other PTLs by the opposing increased

gas flux from the produced gas. This reduction in water flux results in this particular case in

decreasing proton conductivity of CL and membrane and increased diffusion losses.

Figure S1 (see supplementary) shows additional polarization curves of the four PTLs conducted

at three different backpressures and both 60 and 80°C. Increasing the backpressure from 0.88 bar

up to 1.72 bar resulted in all cases in a slight decrease of the performance. This trend is most

pronounced for the sintered Ti PTL at both temperatures (Figure S1(C)) and the 5% PTFE treated

Toray paper at 60°C (Figure S1(A)). In the latter case the effect additionally increases with current

density, which again indicates increasing diffusion losses.

18

Pursuant to the DOE Public Access Plan, this document represents the authors' peer-reviewed, accepted manuscript.
The published version of the article is available from the relevant publisher.
Figure 2. Polarization curves and HFR-free voltage of different PTLs at (A) 60oC and (B)

80oC

Figure 3 shows results from the EIS experiments as well as respective equivalent electrical circuit

(EEC) analysis. Additional spectra are shown for all the tested PTLs in Figures S2 and S3 (see

supplementary) for two temperatures, two current densities and three pressures each.

19

Pursuant to the DOE Public Access Plan, this document represents the authors' peer-reviewed, accepted manuscript.
The published version of the article is available from the relevant publisher.
Figure 3 (A) shows an example of the trends that were observed in this study using the 5% PTFE

treated Toray paper as anode PTLs. The trends include a reduction of the high frequency loop,

development of a low frequency loop at high current densities, and an increase in HFR. Figure 3 (B)

demonstrates the applicability of a previously introduced modified Randles model [30], depicted

in the figure inset, onto the current data set. It consists of an inductor L, a resistor Rohm, and a

constant phase element (CPE) in parallel to a second resistor R2 and a Warburg element Wd. The

EIS Nyquist plots for all the PTLs were fitted for both 60°C and 80°C. The fits show good

agreement to the experimental results as shown in the example in Figure 3 (B) for 5% treated Toray

paper (Figure S4, see supplementary materials). The fit and resulting differences in ohmic,

activation, and diffusion resistances are plotted in Figure 3 (C) (top, middle, bottom, respectively)

and Figure S5 (see supplementary materials).

20

Pursuant to the DOE Public Access Plan, this document represents the authors' peer-reviewed, accepted manuscript.
The published version of the article is available from the relevant publisher.
Figure 3. Nyquist EIS plots for 5% PTFE treated Toray paper (A) experimental data at 80°C

and 14.9 psi with the arrow representing the frequency decreasing direction, (B) EIS fitting

at 1.6 A cm-2, (C) fitted ohmic resistance, activation resistance and diffusion resistance of

different PTLs at 80°C.

Table 2 lists the fitted parameter values for 1.6 A cm-2 and both 60 and 80°C. L represents the

inductance of the system, ROhm the ohmic resistance of the cell, R2 the activation resistance and Rd

the diffusion resistance. At 1.6 A cm-2, the smallest HFR, with 172.5 and 150 mΩ cm2 at 60 and

80°C, respectively, belongs to the cell with the untreated Toray paper. The cell using 5% PTFE

treated Toray paper has a slightly larger HFR, while that using 20% PTFE treated Toray paper has

an HFR that is increased by 68% and 31% at 60 and 80°C operation, respectively (see Table S1 in

supplementary materials). The data indicate that the increased PTFE loading of the Toray paper

increases the ohmic resistance of the cell, specifically at high current densities (Figure 3(C), Ohmic

resistance). The cell built with the sintered Ti PTLs also has a high HFR (229.5 and 190 mΩ cm2

at 60°C and 80°C, respectively). In this case, the relatively large 1 mm thickness of the material

and surface oxidation of the Ti material may have contributed to the elevated ohmic resistance. In

contrast, the Ti felt material features the second lowest HFR, i.e. 186 and 159 mΩ cm2 at 60°C and

80°C, respectively. With a similar thickness than the untreated Toray paper, this increase in

resistance may be attributed to the surface oxidation of Ti material alone. As in Figure 2, trends

exist of the HFR with current density: PTFE treated carbon paper materials increase in resistance

with current density depending on their level of PTFE treatment, while Ti materials decrease with

current density, possibly, as discussed, due to local temperature changes.

21

Pursuant to the DOE Public Access Plan, this document represents the authors' peer-reviewed, accepted manuscript.
The published version of the article is available from the relevant publisher.
The distance between the two x-axis intercepts of each of the EIS Nyquist plots represents the

activation loss [75, 76], which describes the performance of the catalyst/catalyst layer. It is

governed by catalyst and electrode properties including catalyst material, loading and in-plane

electrode resistance. The activation loss can be expected to be identical for identical CCMs.

However, the data in Figure 3(C) (middle) indicate that the activation resistances of the cells were

slightly impacted by the PTLs. Generally, the activation resistance decreased with increasing

current density, which was expected and is due to the overpotential available at the catalyst to drive

the electrochemical reaction. The intensity of the trend however varied, depending on the PTL

material. The lowest activation resistance was observed with untreated Toray paper, followed by

sintered PTL, 5% PTFE treated Toray paper, Ti Felt, 20% PTFE treated Toray paper. Two different

things may contribute to this trend. On the one hand, the increase of activation resistance with

PTFE treatment of the Toray paper implies that the PTFE treatment is reducing the number of

accessible reaction sites in the catalyst layer by the addition of electrically insulating

fluoropolymer onto the carbon fibers of the PTL. On the other hand, the larger mean pore size of

the Ti felt may reduce access to reaction sites when compared to the sintered PTL. [29, 74, 77] To

access the entire catalyst layer within a pore, electrical current has to flow in the in-plane direction

through the catalyst layer, i.e. in direction of the largest electrical resistance. It is conceivable that

large pores are not entirely accessible due to this ohmic loss and thus catalyst material in the center

of the pore becomes inaccessible [30, 35, 39].

Finally, the bottom plot of Figure 3 (C) shows the fitted values of the diffusion resistance. The

results indicate together with the numerical values from Table 2 that the diffusion loss is negligible

for any material other than PTFE treated Toray paper. For the 5% treated Toray paper however,

diffusion losses were observed that increase with increasing current densities and higher PTFE

22

Pursuant to the DOE Public Access Plan, this document represents the authors' peer-reviewed, accepted manuscript.
The published version of the article is available from the relevant publisher.
loading. In PEMWE cells, the working environment of anode PTLs is essentially inverted

compared to fuel cells. The PTFE treated Toray paper is a diffusion media typically used in fuel

cells. The material is treated with PTFE to increase the hydrophobicity and improve the cell’s

water management. In PEMWE cells water needs to be transported from the flow field to the CL

surface through the PTLs and the generated oxygen gas needs to be removed, typically in form of

bubbles, in the reversed direction. It is desirable for anode PTLs in PEMWE to have high water

permeability and hydrophilicity to allow water transport readily into and through the porous

structures. The presented data indicates that any PTFE treatment changes the wettability of the

PTL, resulting in a more hydrophobic PTL, less favorable water transport, higher diffusion losses

and thus lower performance. In summary, it can be concluded that the PTL may impact ohmic and

activation losses and that PTFE treatments additionally affect the diffusion loss of PEMWE cells.

Table 2. Nyquist EIS plot fitting parameters with different PTLs at 1.6 A cm-2
Untreated Toray 5% PTFE Treated
Sintered Ti PTLs Ti Felt
paper Toray paper
60°C 80°C 60°C 80°C 60°C 80°C 60°C 80°C
L [H] 11.01e-9 9.562e-9 16.87E-9 17.62e-9 11.59e-9 12.06e-9 8.845e-9 10.62e-9
ROhm [Ohm] 0.0345 0.03 0.0382 0.0343 0.0459 0.038 0.0372 0.0318
Q [F.s^(a-1)] 4.137 5.553 2.934 3.241 2.34 4.388 0.5098 0.5839
a 0.6171 0.6293 0.703 0.7093 0.6623 0.6065 0.7512 0.7647
R2 [Ohm] 0.004352 0.004086 0.004919 0.004928 0.004595 0.004467 0.005662 0.004914
Rd [Ohm] 15.3e-12 3.66e-15 0.002842 0.002696 2.354e-6 2.343e-9 1.376e-6 25.17e-9
td [s] 0.334 0.02495 3.508 2.796 1.773e-3 4.486e-3 5.409e-6 3.171e-6

Figure 4 shows the electrical resistance of different PTLs measured with a four-wire DC technique

in a PEMWE cell assembled with two PTLs and their gaskets but without membrane and catalyst

layers. The total ohmic resistance of this measurement is shown in Figure 4 for the PTL materials

used in this work. Compared to the untreated material, the 5% PTFE treatment slightly increases

the electrical resistance of the Toray paper from about 14.5 to 16 m cm2. This trend is

23

Pursuant to the DOE Public Access Plan, this document represents the authors' peer-reviewed, accepted manuscript.
The published version of the article is available from the relevant publisher.
significantly magnified for the 20% PTFE treatment which has a resistance of approximately

38 m cm2 and matches in value to the sintered material. Ti felt with 350 µm individual PTL

thickness, being much thinner than the 1 mm thick sintered Ti PTL, achieves lower electrical

resistance but its resistance is still higher than untreated and 5% PTFE treated Toray paper. It is

known that Ti material develops oxides layer that passivate the surface of the material with regards

to its electrical interface resistance [23]. Since the pristine unused Ti materials used in this work

show increased contact resistances it is likely that TiOx passivation layers contributed to the

increased ohmic losses previously discussed. In general, this validation of combined material and

contact resistance is in good agreement with the results extracted from EIS. It confirms that the

selection of materials with regards to their material and interface properties is essential for

optimizing cell performance.

Figure 4. Electrical resistance of different PTLs

4.2 Model validation

24

Pursuant to the DOE Public Access Plan, this document represents the authors' peer-reviewed, accepted manuscript.
The published version of the article is available from the relevant publisher.
In this section, the experimental data from the previous section is used to obtain all the fitted

parameters in the model and validate the model. Table 3 shows the basic geometrical and

physicochemical parameters used in the presented model. Figure 5 (A) and (B) show the

comparison between the experimental data and modeling results for the four different PTLs at

operating temperatures of 60 and 80°C, respectively. The presented model agrees in all cases with

the experimental data over the entire current density range. Note that since identical CCMs were

used for the experiments, the exchange current density and charge transfer coefficient were

assumed to be constant, while the interfacial contact resistance coefficients of Toray paper,

sintered Ti PTLs and Ti felt were adjusted for each case due to their differing structure and contact

properties.

Table 3. Basic parameters for the comprehensive PEMWE modeling.


Description, symbol Value, unit
MEA active area 5.0 cm2
PEM Nafion 117 with ~175 µm thickness
Basic operating pressure, P Anode: 14.9 psi, cathode: 14.9 psi
Operating temperature, T 60°C and 80°C
Toray paper 350 𝜇𝑚 thickness
Sintered Ti particles 1100 𝜇𝑚 thickness
Toray paper initial porosity 0.78
Sintered Ti PTL porosity 0.4
Carbon fiber density 2.0 g cm-3 [72]
Ti material density 4.506 g cm-3
Carbon material contact angle 80° [72]
Ti material contact angle 45° [78]
Wettability treatment additive density PTFE: 2.16 g cm-3; FEP: 2.15 g cm-3 [72]
Additive contact angle PTFE: 108°; FEP: 100° [72]
CL thickness 15 𝜇𝑚
Toray paper electrical resistivity 8.0×10-2 Ω cm [79]

25

Pursuant to the DOE Public Access Plan, this document represents the authors' peer-reviewed, accepted manuscript.
The published version of the article is available from the relevant publisher.
Graphite electrical resistivity 1.5×10-3 Ω cm [80]
Anode CL electrical resistivity (Through plane) 5×10-2 Ω cm [79]
Cathode CL electrical resistivity (Through plane) 1.4×10-3 Ω cm [79]
Liquid water dynamic viscosity, 𝜇𝐻2 𝑂 3.55×10-4 N s m-2 [79]
Liquid water density, 𝜌𝐻2𝑂 1000 kg m-3
Charge transfer coefficient, αan and αcat Anode: 2.0, Cathode: 1.5
Exchange current density at 80oC (A cm-2) (Fitted) Anode: 1.5× 10−8 , Cathode: 5.0× 10−5
Membrane humidification degree 22
Interfacial contact resistance coefficient, K (Fitted) Toray paper: 12.5 mΩ cm2
Sintered Ti PTLs: 80 mΩ cm2
Ti felt: 45 mΩ cm2

26

Pursuant to the DOE Public Access Plan, this document represents the authors' peer-reviewed, accepted manuscript.
The published version of the article is available from the relevant publisher.
Figure 5. Model validation at different temperatures (A) 60oC (B) 80oC.

4.3 Overpotential visualization

27

Pursuant to the DOE Public Access Plan, this document represents the authors' peer-reviewed, accepted manuscript.
The published version of the article is available from the relevant publisher.
The model lends itself to diagnosing and analyzing the loss contributions of individual cell

components, including PTLs and electrodes. Therefore, in this section, the model is utilized to

visualize the overpotential losses. By taking advantage of the model, cell voltage at higher current

densities could be simulated and deconvoluted by using the model without any risks of damaging

PEMEC components due to corrosion or any limitations (maximum current/voltage) from the

instrument. Figure 6 shows the breakdown of the cell overpotentials for four different PTLs at

80°C and a current density range up to 5 A cm-2. For constant temperature, the Nernst voltage

(grey section in Figure 6) is constant across all experimental configurations. The activation

overpotential (act, pink) is relatively independent from current density and the largest loss below

approximately 2.0 A cm-2. For the presented data, act has similar values among the PTLs, which

has the same trend compared to EIS data fitting in Figure 3(C). This is expected since identical

CCMs were used in this work. The Ohmic overpotential (Ohm, blue) increases linearly with current

density. For the presented cases it surpasses the activation loss when the current density increases

above approximately 2.0 A cm-2. As discussed in previous sections, this loss is governed by

material properties and can also be influenced by interface related processes. Compared to the

other loss contributions, the diffusion overpotential (diff, orange) of the untreated Toray paper and

both Ti PTLs are negligible throughout the entire current density range. The diffusion

overpotential of the 5% PTFE treated Toray paper however starts to become apparent at about

1.5 A cm-2. It increases nonlinearly with current density and becomes more dominant than the

activation overpotential at high current densities. In Figure 6 (A), the diffusion loss is attributed to

the hydrophobic PTFE treatment that changes the wettability of the Toray paper. With 5% PTFE

treatment, the equilibrium contact angle on Toray paper is increased, indicating more hydrophobic

behavior and leading to higher diffusion loss. The results show that the wettability of the PTLs is

28

Pursuant to the DOE Public Access Plan, this document represents the authors' peer-reviewed, accepted manuscript.
The published version of the article is available from the relevant publisher.
a critical parameter. Unfortunately, the hydrophilic treatment on anode PTLs can, as indicated by

the negligible diffusion losses of the untreated materials, only result in very limited performance

improvements. It is much more effective to focus on reducing activation and ohmic overpotentials

instead.

Figure 6. Contributions of each overpotential to polarization curves with different PTLs at

80oC (A) 5% PTFE treated Toray paper; (B) Untreated Toray paper; (C) Sintered Ti

particles; (D) Ti felt.

29

Pursuant to the DOE Public Access Plan, this document represents the authors' peer-reviewed, accepted manuscript.
The published version of the article is available from the relevant publisher.
4.4 Modeling prediction

In this section, the model is used to investigate the performance of the 20% PTFE treated Toray

paper at first, and then to simulate the effects of PTFE loading in different PTLs at various

temperatures. The model is further validated by predicting the performance of 20% PTFE treated

Toray paper via adjusting the PTFE loading parameter. As shown in Figure 7, the data is

subsequently compared to experimental data at 60 and 80°C. The diffusion overpotential of the

20% PTFE treated Toray paper becomes apparent above 0.7 A cm-2. It increases nonlinearly with

current density and becomes more dominant than both activation and Ohmic overpotentials at high

current densities. The results indicate good agreement between the predictive and the experimental

data sets, thus demonstrating the ability of the model to investigate critical PTL parameters.

30

Pursuant to the DOE Public Access Plan, this document represents the authors' peer-reviewed, accepted manuscript.
The published version of the article is available from the relevant publisher.
Figure 7. (A) Modeling prediction and comparison to experimental results of 20% PTFE treated

Toray paper and (B) overpotential visualization at 80°C

31

Pursuant to the DOE Public Access Plan, this document represents the authors' peer-reviewed, accepted manuscript.
The published version of the article is available from the relevant publisher.
Figure 8 shows predictive modeling results for PEMWE cells operated with PTFE loadings

ranging from 0% to 30%. The data illustrate the influence of PTFE loading on overall PEMWE

performance (Figure 8 (A)) and the respective diffusion overpotential loss of the cell

(Figure 8 (B)). The untreated (0% PTFE) Toray paper is predicted to have the best performance

and the smallest diffusion loss. By increasing the PTFE content to 5%, the diffusion loss will

significantly increase from < 10 mV to > 100 mV at 2.0 A cm-2. This trend dramatically continues

when further increasing the PTFE loading, which results in very high cell voltages, poor

performance and thus an impractical cell configuration.

Figure 8. Effects of PTFE loading in Toray paper on PEMWE performance (A) Polarization

curves with different PTFE loadings; (B) Diffusion losses comparison

The hydrophobic treatment of carbon paper will result in much higher cell voltage due to the

increased diffusion losses, while the effects on Ti based PTLs are still unknown. The model was

utilized to predict the performance and quantify the effects of additive loadings and operating

32

Pursuant to the DOE Public Access Plan, this document represents the authors' peer-reviewed, accepted manuscript.
The published version of the article is available from the relevant publisher.
temperatures onto these PTLs. Previous research showed that a water contact angle difference of

only 8° on the pure fluoropolymer surface could cause tremendous influences on porous material

equilibrium contact angle [72]. Therefore, different additives (PTFE and FEP) are also studied.

Figure 9 (A) shows the voltage of PEMWE at 2.0 A cm-2 using PTFE or FEP treated Toray paper,

sintered Ti PTLs and Ti felt. Toray paper achieves lower cell voltages than Ti PTLs for similar

fluoropolymer loading. The differences between Ti based PTLs and carbon PTLs can be explained

by the structural differences and equilibrium contact angle changes. With 0% hydrophobic

treatment, Ti based PTLs have higher cell voltage due to the larger ohmic resistances as discussed

before, and they also obtain a hydrophilic nature. With the increase of hydrophobic additive

loading, the equilibrium contact angle of Ti based PTLs increases much faster than that of Toray

paper, which causes a more rapidly increasing cell voltage. Toray paper has a more hydrophobic

nature even without any treatment, which is due to the large porosity of Toray paper (~78%) and

the large contact angle of single carbon fiber (~80°). The equilibrium contact angle (shown in

Figure 9 (B)) will increase from ~128° to ~141° or 135° with 30% weight percentage PTFE or

FEP treatment, respectively. The equilibrium contact angle will be larger with PTFE than FEP

under the same loading due to the 8° difference between PTFE and FEP. The sintered Ti PTLs

exhibit a hydrophilic nature when the PTFE or FEP loading is smaller than 5%, which is due to

the small contact angle of Ti materials (~45°). Ti felt shows a larger contact angle than sintered Ti

PTLs due to its larger porosity. The sintered Ti particles and Ti felt are also more sensitive to

hydrophobic treatment than carbon paper and the PTFE treated sample has a higher equilibrium

contact angle than that treated with FEP. Figure 9 (C), (D) and Figure S7 (see supplementary

materials) show the quantified temperature impacts. With the increase of operating temperature,

the cell performance will gradually improve. However, when fluoropolymer is added to the PTLs,

33

Pursuant to the DOE Public Access Plan, this document represents the authors' peer-reviewed, accepted manuscript.
The published version of the article is available from the relevant publisher.
especially for the sintered Ti PTLs and Ti felt, the cell voltage will increase significantly. When

the cell is operated at higher temperature, the Nernst potential will drop based on Equation (2).

The proton conductivity of the Nafion membrane will improve, which will lead to reduced ohmic

resistance. Meanwhile, the diffusion process, catalytic activity and kinetics could also be improved

at higher temperatures.

34

Pursuant to the DOE Public Access Plan, this document represents the authors' peer-reviewed, accepted manuscript.
The published version of the article is available from the relevant publisher.
Figure 9. Effects of hydrophobic agent loadings on (A) total cell voltage, (B) PTL equilibrium

contact angle, and effects of temperature on cell voltage of (C) Toray paper, (D) Sintered Ti

PTLs at 2.0 A cm-2 and 80°C with different additive loadings

4. Conclusion

In this study, a general mathematic PEMWE model is established which uses terms that attribute

loss contributions to specific PTL parameters including PTL material, structure, and wettability.

The performance characteristics of MEAs with three inherently different anode PTLs, i.e. carbon

paper, sintered Ti particle, and Ti felt, have been investigated and successfully used for model

development and validation. The results indicate that PEMWE performance is closely related to

the anode PTL properties, which can significantly impact ohmic, activation and diffusion losses.

This was verified by the model’s ability to separate the overpotentials within the cell. Shifting the

PTL wettability to a more hydrophobic nature by adding hydrophobic additives/agents increased

not only the diffusion loss significantly, as expected, but also the ohmic loss and activation loss.

The additional losses of the system were specifically detrimental to PEMWE cell performance

when operating at high current densities. The introduced model was capable to quantitively predict

the performance of different PTLs and highlight the detrimental effects of hydrophobic PTL

treatments. It was successfully verified as a predictive tool for PTL performance evaluation and is

readily available to support PTL selection and optimization of developmental PTL and MEA

constructs.

Acknowledgement

35

Pursuant to the DOE Public Access Plan, this document represents the authors' peer-reviewed, accepted manuscript.
The published version of the article is available from the relevant publisher.
This work was authored by the National Renewable Energy Laboratory, operated by Alliance for

Sustainable Energy, LLC, for the U.S. Department of Energy (DOE) under Contract No. DE-

AC36-08GO28308. Funding provided by U.S. Department of Energy Office of Energy Efficiency

and Renewable Energy (EERE) Hydrogen and Fuel Cell Technologies Office (HFTO). The U.S.

Government retains and the publisher, by accepting the article for publication, acknowledges that

the U.S. Government retains a nonexclusive, paid-up, irrevocable, worldwide license to publish or

reproduce the published form of this work, or allow others to do so, for U.S. Government purposes.

The views expressed in the article do not necessarily represent the views of the DOE or the U.S.

Government. We would also like to thank Dr. Marcelo Carmo and the Forschungszentrum Jülich

for providing CCMs and discussions that supported this study.

References

[1] J.A. Turner, A realizable renewable energy future, Science, 285 (1999) 687-689.
[2] X. Zhou, X. Yang, M.N. Hedhili, H. Li, S. Min, J. Ming, K.-W. Huang, W. Zhang, L.-J. Li, Symmetrical
synergy of hybrid Co9S8-MoSx electrocatalysts for hydrogen evolution reaction, Nano Energy, 32 (2017)
470-478.
[3] M.P. Browne, Z. Sofer, M. Pumera, Layered and two dimensional metal oxides for electrochemical
energy conversion, Energy & Environmental Science, 12 (2019) 41-58.
[4] S. Siracusano, N. Hodnik, P. Jovanovic, F. Ruiz-Zepeda, M. Šala, V. Baglio, A.S. Aricò, New insights into
the stability of a high performance nanostructured catalyst for sustainable water electrolysis, Nano
Energy, 40 (2017) 618-632.
[5] L. Wang, V.A. Saveleva, S. Zafeiratos, E.R. Savinova, P. Lettenmeier, P. Gazdzicki, A.S. Gago, K.A.
Friedrich, Highly active anode electrocatalysts derived from electrochemical leaching of Ru from metallic
Ir0. 7Ru0. 3 for proton exchange membrane electrolyzers, Nano energy, 34 (2017) 385-391.
[6] B. Pivovar, N. Rustagi, S. Satyapal, Hydrogen at Scale (H2@ Scale): Key to a Clean, Economic, and
Sustainable Energy System, The Electrochemical Society Interface, 27 (2018) 47-52.
[7] M. Wang, J.H. Park, S. Kabir, K.C. Neyerlin, N.N. Kariuki, H. Lv, V.R. Stamenkovic, D.J. Myers, M. Ulsh,
S.A. Mauger, Impact of Catalyst Ink Dispersing Methodology on Fuel Cell Performance Using in-Situ X-ray
Scattering, ACS Applied Energy Materials, 2 (2019) 6417-6427.
[8] S.M. Alia, S. Shulda, C. Ngo, S. Pylypenko, B.S. Pivovar, Iridium-based nanowires as highly active,
oxygen evolution reaction electrocatalysts, ACS Catalysis, 8 (2018) 2111-2120.
[9] G. Yang, S. Yu, Z. Kang, Y. Li, G. Bender, B.S. Pivovar, J.B. Green Jr, D.A. Cullen, F.Y. Zhang, Building
Electron/Proton Nanohighways for Full Utilization of Water Splitting Catalysts, Advanced Energy
Materials, (2020) 1903871.

36

Pursuant to the DOE Public Access Plan, this document represents the authors' peer-reviewed, accepted manuscript.
The published version of the article is available from the relevant publisher.
[10] B. Pivovar, Catalysts for fuel cell transportation and hydrogen related uses, Nature Catalysis, 2
(2019) 562.
[11] F. Godínez-Salomón, L. Albiter, S.M. Alia, B.S. Pivovar, L.E. Camacho-Forero, P.B. Balbuena, R.
Mendoza-Cruz, M.J. Arellano-Jimenez, C.P. Rhodes, Self-supported hydrous iridium–nickel oxide two-
dimensional nanoframes for high activity oxygen evolution electrocatalysts, ACS Catalysis, 8 (2018)
10498-10520.
[12] P. Lettenmeier, S. Kolb, N. Sata, A. Fallisch, L. Zielke, S. Thiele, A. Gago, K. Friedrich, Comprehensive
investigation of novel pore-graded gas diffusion layers for high-performance and cost-effective proton
exchange membrane electrolyzers, Energy & Environmental Science, 10 (2017) 2521-2533.
[13] S. Park, J.-W. Lee, B.N. Popov, A review of gas diffusion layer in PEM fuel cells: Materials and
designs, International Journal of Hydrogen Energy, 37 (2012) 5850-5865.
[14] G. Chen, G. Zhang, L. Guo, H. Liu, Systematic study on the functions and mechanisms of micro
porous layer on water transport in proton exchange membrane fuel cells, International Journal of
Hydrogen Energy, 41 (2016) 5063-5073.
[15] G. Bender, M. Carmo, T. Smolinka, A. Gago, N. Danilovic, M. Mueller, F. Ganci, A. Fallisch, P.
Lettenmeier, K. Friedrich, Initial approaches in benchmarking and round robin testing for proton
exchange membrane water electrolyzers, International Journal of Hydrogen Energy, 44 (2019) 9174-
9187.
[16] M. Carmo, D.L. Fritz, J. Mergel, D. Stolten, A comprehensive review on PEM water electrolysis,
International journal of hydrogen energy, 38 (2013) 4901-4934.
[17] X. Zhang, Y. Yang, L. Guo, H. Liu, Effects of carbon corrosion on mass transfer losses in proton
exchange membrane fuel cells, International Journal of Hydrogen Energy, 42 (2017) 4699-4705.
[18] Z. Kang, Development of Novel Thin Membrane Electrode Assemblies (MEAs) for High-Efficiency
Energy Storage, (2018).
[19] Z. Kang, S.M. Alia, J.L. Young, G. Bender, Investigation of Porous Transport Layer Parameters for
Proton Exchange Membrane Water Electrolysis, Meeting Abstracts, The Electrochemical Society, 2019,
pp. 1747-1747.
[20] H. Ito, T. Maeda, A. Nakano, A. Kato, T. Yoshida, Influence of pore structural properties of current
collectors on the performance of proton exchange membrane electrolyzer, Electrochimica Acta, 100
(2013) 242-248.
[21] F. Arbabi, A. Kalantarian, R. Abouatallah, R. Wang, J. Wallace, A. Bazylak, Feasibility study of using
microfluidic platforms for visualizing bubble flows in electrolyzer gas diffusion layers, Journal of Power
Sources, 258 (2014) 142-149.
[22] G. Yang, J. Mo, Z. Kang, F.A. List, J.B. Green, S.S. Babu, F.-Y. Zhang, Additive manufactured bipolar
plate for high-efficiency hydrogen production in proton exchange membrane electrolyzer cells,
International Journal of Hydrogen Energy, 42 (2017) 14734-14740.
[23] C. Liu, M. Carmo, G. Bender, A. Everwand, T. Lickert, J.L. Young, T. Smolinka, D. Stolten, W. Lehnert,
Performance enhancement of PEM electrolyzers through iridium-coated titanium porous transport
layers, Electrochemistry Communications, 97 (2018) 96-99.
[24] S.M. Steen III, J. Mo, Z. Kang, G. Yang, F.-Y. Zhang, Investigation of titanium liquid/gas diffusion
layers in proton exchange membrane electrolyzer cells, International journal of green energy, 14 (2017)
162-170.
[25] S. Siracusano, A. Di Blasi, V. Baglio, G. Brunaccini, N. Briguglio, A. Stassi, R. Ornelas, E. Trifoni, V.
Antonucci, A. Aricò, Optimization of components and assembling in a PEM electrolyzer stack,
International Journal of Hydrogen Energy, 36 (2011) 3333-3339.
[26] M.S. Hossain, B. Shabani, Metal foams application to enhance cooling of open cathode polymer
electrolyte membrane fuel cells, Journal of Power Sources, 295 (2015) 275-291.

37

Pursuant to the DOE Public Access Plan, this document represents the authors' peer-reviewed, accepted manuscript.
The published version of the article is available from the relevant publisher.
[27] P. Lettenmeier, S. Kolb, F. Burggraf, A. Gago, K. Friedrich, Towards developing a backing layer for
proton exchange membrane electrolyzers, Journal of Power Sources, 311 (2016) 153-158.
[28] S. Grigoriev, K. Dzhus, D. Bessarabov, P. Millet, Failure of PEM water electrolysis cells: Case study
involving anode dissolution and membrane thinning, International Journal of Hydrogen Energy, 39
(2014) 20440-20446.
[29] J. Lopata, Z. Kang, J. Young, G. Bender, J. Weidner, S. Shimpalee, Effects of the Transport/Catalyst
Layer Interface and Catalyst Loading on Mass and Charge Transport Phenomena in Polymer Electrolyte
Membrane Water Electrolysis Devices, Journal of The Electrochemical Society, 167 (2020) 064507.
[30] Z. Kang, J. Mo, G. Yang, S.T. Retterer, D.A. Cullen, T.J. Toops, J.B. Green Jr, M.M. Mench, F.-Y. Zhang,
Investigation of thin/well-tunable liquid/gas diffusion layers exhibiting superior multifunctional
performance in low-temperature electrolytic water splitting, Energy & Environmental Science, 10 (2017)
166-175.
[31] Z. Kang, J. Mo, G. Yang, Y. Li, D.A. Talley, S.T. Retterer, D.A. Cullen, T.J. Toops, M.P. Brady, G. Bender,
Thin film surface modifications of thin/tunable liquid/gas diffusion layers for high-efficiency proton
exchange membrane electrolyzer cells, Applied Energy, 206 (2017) 983-990.
[32] G. Yang, S. Yu, Z. Kang, Y. Dohrmann, G. Bender, B.S. Pivovar, J.B. Green Jr, S.T. Retterer, D.A. Cullen,
F.-Y. Zhang, A novel PEMEC with 3D printed non-conductive bipolar plate for low-cost hydrogen
production from water electrolysis, Energy Conversion and Management, 182 (2019) 108-116.
[33] Y. Li, Z. Kang, X. Deng, G. Yang, S. Yu, J. Mo, D.A. Talley, G.K. Jennings, F.-Y. Zhang, Wettability
effects of thin titanium liquid/gas diffusion layers in proton exchange membrane electrolyzer cells,
Electrochimica Acta, 298 (2019) 704-708.
[34] Y. Li, Z. Kang, J. Mo, G. Yang, S. Yu, D.A. Talley, B. Han, F.-Y. Zhang, In-situ investigation of bubble
dynamics and two-phase flow in proton exchange membrane electrolyzer cells, International Journal of
Hydrogen Energy, 43 (2018) 11223-11233.
[35] J. Mo, Z. Kang, G. Yang, Y. Li, S.T. Retterer, D.A. Cullen, T.J. Toops, G. Bender, B.S. Pivovar, J.B. Green
Jr, In situ investigation on ultrafast oxygen evolution reactions of water splitting in proton exchange
membrane electrolyzer cells, Journal of Materials Chemistry A, 5 (2017) 18469-18475.
[36] Z. Kang, S. Yu, G. Yang, Y. Li, G. Bender, B.S. Pivovar, J.B. Green Jr, F.-Y. Zhang, Performance
improvement of proton exchange membrane electrolyzer cells by introducing in-plane transport
enhancement layers, Electrochimica Acta, 316 (2019) 43-51.
[37] F. Yi, S. Chen, Effect of activated carbon fiber anode structure and electrolysis conditions on
electrochemical degradation of dye wastewater, Journal of hazardous materials, 157 (2008) 79-87.
[38] C.A. Bonino, J.J. Concepcion, J.A. Trainham, T.J. Meyer, J. Newman, Water electrolysis with a
homogeneous catalyst in an electrochemical cell, Journal of The Electrochemical Society, 160 (2013)
F1143-F1150.
[39] Z. Kang, G. Yang, J. Mo, Y. Li, S. Yu, D.A. Cullen, S.T. Retterer, T.J. Toops, G. Bender, B.S. Pivovar,
Novel thin/tunable gas diffusion electrodes with ultra-low catalyst loading for hydrogen evolution
reactions in proton exchange membrane electrolyzer cells, Nano Energy, 47 (2018) 434-441.
[40] J.L. Young, F. Ganci, S. Madachy, S. Fischer, M. Carmo, G. Bender, PEM Electrolyzer Characterization
and Limitations When Using Carbon-Based Hardware and Material Sets, Meeting Abstracts, The
Electrochemical Society, 2019, pp. 1541-1541.
[41] H. Ito, T. Maeda, A. Nakano, C.M. Hwang, M. Ishida, A. Kato, T. Yoshida, Experimental study on
porous current collectors of PEM electrolyzers, International journal of hydrogen energy, 37 (2012)
7418-7428.
[42] S. Grigoriev, P. Millet, S. Volobuev, V. Fateev, Optimization of porous current collectors for PEM
water electrolysers, International journal of hydrogen energy, 34 (2009) 4968-4973.

38

Pursuant to the DOE Public Access Plan, this document represents the authors' peer-reviewed, accepted manuscript.
The published version of the article is available from the relevant publisher.
[43] C.-M. Hwang, M. Ishida, H. Ito, T. Maeda, A. Nakano, A. Kato, T. Yoshida, Effect of PTFE contents in
the gas diffusion layers of polymer electrolyte-based unitized reversible fuel cells, Journal of
International Council on Electrical Engineering, 2 (2012) 171-177.
[44] C. Hwang, H. Ito, T. Maeda, A. Nakano, A. Kato, T. Yoshida, Flow field design for a polymer
electrolyte unitized reversible fuel cell, ECS Transactions, 50 (2013) 787-794.
[45] C.M. Hwang, M. Ishida, H. Ito, T. Maeda, A. Nakano, Y. Hasegawa, N. Yokoi, A. Kato, T. Yoshida,
Influence of properties of gas diffusion layers on the performance of polymer electrolyte-based unitized
reversible fuel cells, International journal of hydrogen energy, 36 (2011) 1740-1753.
[46] H. Sadeghifar, N. Djilali, M. Bahrami, Effect of Polytetrafluoroethylene (PTFE) and micro porous
layer (MPL) on thermal conductivity of fuel cell gas diffusion layers: Modeling and experiments, Journal
of Power Sources, 248 (2014) 632-641.
[47] Z. Kang, J. Mo, G. Yang, Y. Li, D.A. Talley, B. Han, F.-Y. Zhang, Performance modeling and current
mapping of proton exchange membrane electrolyzer cells with novel thin/tunable liquid/gas diffusion
layers, Electrochimica Acta, 255 (2017) 405-416.
[48] B. Han, S.M. Steen III, J. Mo, F.-Y. Zhang, Electrochemical performance modeling of a proton
exchange membrane electrolyzer cell for hydrogen energy, International Journal of Hydrogen Energy, 40
(2015) 7006-7016.
[49] A. Awasthi, K. Scott, S. Basu, Dynamic modeling and simulation of a proton exchange membrane
electrolyzer for hydrogen production, International journal of hydrogen energy, 36 (2011) 14779-14786.
[50] T. Yigit, O.F. Selamet, Mathematical modeling and dynamic Simulink simulation of high-pressure
PEM electrolyzer system, International Journal of Hydrogen Energy, 41 (2016) 13901-13914.
[51] A. Nouri-Khorasani, E.T. Ojong, T. Smolinka, D.P. Wilkinson, Model of oxygen bubbles and
performance impact in the porous transport layer of PEM water electrolysis cells, International journal
of hydrogen energy, 42 (2017) 28665-28680.
[52] S.H. Frensch, A.C. Olesen, S.S. Araya, S.K. Kær, Model-supported characterization of a PEM water
electrolysis cell for the effect of compression, Electrochimica Acta, 263 (2018) 228-236.
[53] S. Toghyani, E. Afshari, E. Baniasadi, Three-dimensional computational fluid dynamics modeling of
proton exchange membrane electrolyzer with new flow field pattern, Journal of Thermal Analysis and
Calorimetry, 135 (2019) 1911-1919.
[54] B. Han, J. Mo, Z. Kang, G. Yang, W. Barnhill, F.-Y. Zhang, Modeling of two-phase transport in proton
exchange membrane electrolyzer cells for hydrogen energy, International Journal of Hydrogen Energy,
42 (2017) 4478-4489.
[55] B. Han, J. Mo, Z. Kang, F.-Y. Zhang, Effects of membrane electrode assembly properties on two-
phase transport and performance in proton exchange membrane electrolyzer cells, Electrochimica Acta,
188 (2016) 317-326.
[56] M. Espinosa-López, C. Darras, P. Poggi, R. Glises, P. Baucour, A. Rakotondrainibe, S. Besse, P. Serre-
Combe, Modelling and experimental validation of a 46 kW PEM high pressure water electrolyzer,
Renewable energy, 119 (2018) 160-173.
[57] S. Toghyani, E. Afshari, E. Baniasadi, S. Atyabi, G. Naterer, Thermal and electrochemical
performance assessment of a high temperature PEM electrolyzer, Energy, 152 (2018) 237-246.
[58] M. Zlobinski, T. Schuler, F.N. Büchi, T.J. Schmidt, P. Boillat, Transient and Steady State Two-Phase
Flow in Anodic Porous Transport Layer of Proton Exchange Membrane Water Electrolyzer, Journal of The
Electrochemical Society, 167 (2020) 084509.
[59] T. Schuler, R. De Bruycker, T.J. Schmidt, F.N. Büchi, Polymer Electrolyte Water Electrolysis:
Correlating porous transport layer structural properties and performance: Part I. Tomographic analysis
of morphology and topology, Journal of The Electrochemical Society, 166 (2019) F270-F281.

39

Pursuant to the DOE Public Access Plan, this document represents the authors' peer-reviewed, accepted manuscript.
The published version of the article is available from the relevant publisher.
[60] T. Schuler, T.J. Schmidt, F.N.J.J.o.T.E.S. Büchi, Polymer Electrolyte Water Electrolysis: Correlating
Performance and Porous Transport Layer Structure: Part II. Electrochemical Performance Analysis, 166
(2019) F555-F565.
[61] J. Lee, A. Bazylak, Stochastic Modelling For Controlling the Structure of Sintered Titanium Powder-
Based Porous Transport Layers for Polymer Electrolyte Membrane Electrolyzers, Journal of The
Electrochemical Society, 166 (2019) F1000-F1006.
[62] J. Lee, A. Bazylak, Optimizing Porous Transport Layer Design Parameters via Stochastic Pore
Network Modelling: Reactant Transport and Interfacial Contact Considerations, Journal of The
Electrochemical Society, 167 (2020) 013541.
[63] Y. Chen, T. Tian, Z. Wan, F. Wu, J. Tan, M. Pan, Influence of PTFE on water transport in gas diffusion
layer of polymer electrolyte membrane fuel cell, INTERNATIONAL JOURNAL OF ELECTROCHEMICAL
SCIENCE, 13 (2018) 3827-3842.
[64] R. Omrani, B. Shabani, Review of gas diffusion layer for proton exchange membrane-based
technologies with a focus on unitised regenerative fuel cells, International Journal of Hydrogen Energy,
(2019).
[65] S. Wang, X. Li, Z. Wan, Y. Chen, J. Tan, M. Pan, Effect of hydrophobic additive on oxygen transport in
catalyst layer of proton exchange membrane fuel cells, Journal of Power Sources, 379 (2018) 338-343.
[66] C. Tötzke, I. Manke, C. Hartnig, R. Kuhn, H. Riesemeier, J. Banhart, Investigation of carbon fiber gas
diffusion layers by means of synchrotron X-ray tomography, ECS Transactions, 41 (2011) 379-386.
[67] V.N. Burganos, E.D. Skouras, A.N. Kalarakis, An integrated simulator of structure and anisotropic
flow in gas diffusion layers with hydrophobic additives, Journal of Power Sources, 365 (2017) 179-189.
[68] A.A. Rahim, A.S. Tijani, S.K. Kamarudin, S. Hanapi, An overview of polymer electrolyte membrane
electrolyzer for hydrogen production: Modeling and mass transport, Journal of Power Sources, 309
(2016) 56-65.
[69] H. Su, V. Linkov, B.J. Bladergroen, Membrane electrode assemblies with low noble metal loadings
for hydrogen production from solid polymer electrolyte water electrolysis, International Journal of
Hydrogen Energy, 38 (2013) 9601-9608.
[70] Z. Abdin, C. Webb, E.M. Gray, Modelling and simulation of a proton exchange membrane (PEM)
electrolyser cell, International Journal of Hydrogen Energy, 40 (2015) 13243-13257.
[71] B. Han, S.M. Steen, J. Mo, F.-Y. Zhang, Electrochemical performance modeling of a proton exchange
membrane electrolyzer cell for hydrogen energy, International Journal of Hydrogen Energy, 40 (2015)
7006-7016.
[72] D.L. Wood, Fundamental material degradation studies during long-term operation of hydrogen/air
PEMFCs, 2007.
[73] T. Cubaud, M. Fermigier, Advancing contact lines on chemically patterned surfaces, Journal of
colloid and interface science, 269 (2004) 171-177.
[74] M. Bernt, A. Siebel, H.A. Gasteiger, Analysis of Voltage Losses in PEM Water Electrolyzers with Low
Platinum Group Metal Loadings, Journal of The Electrochemical Society, 165 (2018) F305-F314.
[75] S. Sun, Y. Xiao, D. Liang, Z. Shao, H. Yu, M. Hou, B. Yi, Behaviors of a proton exchange membrane
electrolyzer under water starvation, Rsc Advances, 5 (2015) 14506-14513.
[76] N. Dale, M. Mann, H. Salehfar, A. Dhirde, T. Han, Ac impedance study of a proton exchange
membrane fuel cell stack under various loading conditions, Journal of Fuel Cell Science and Technology,
7 (2010) 031010.
[77] U. Babic, E. Nilsson, A. Pătru, T.J. Schmidt, L. Gubler, Proton Transport in Catalyst Layers of a
Polymer Electrolyte Water Electrolyzer: Effect of the Anode Catalyst Loading, Journal of The
Electrochemical Society, 166 (2019) F214-F220.

40

Pursuant to the DOE Public Access Plan, this document represents the authors' peer-reviewed, accepted manuscript.
The published version of the article is available from the relevant publisher.
[78] Z. Kang, G. Yang, J. Mo, S. Yu, D.A. Cullen, S.T. Retterer, T.J. Toops, M.P. Brady, G. Bender, B.S.
Pivovar, Developing titanium micro/nano porous layers on planar thin/tunable LGDLs for high-efficiency
hydrogen production, International Journal of Hydrogen Energy, 43 (2018) 14618-14628.
[79] J. Mo, Z. Kang, G. Yang, S.T. Retterer, D.A. Cullen, T.J. Toops, J.B. Green, F.-Y. Zhang, Thin liquid/gas
diffusion layers for high-efficiency hydrogen production from water splitting, Applied energy, 177 (2016)
817-822.
[80] I. Poco Graphite, http://poco.com/MaterialsandServices/Graphite/IndustrialGrades/AXF5Q.aspx,
2017.

41

Pursuant to the DOE Public Access Plan, this document represents the authors' peer-reviewed, accepted manuscript.
The published version of the article is available from the relevant publisher.

You might also like