Progress in Energy and Combustion Science 37 (2011) 221e291
Contents lists available at ScienceDirect
Progress in Energy and Combustion Science
journal homepage: www.elsevier.com/locate/pecs
Review
Water transport in polymer electrolyte membrane fuel cells
Kui Jiao a, Xianguo Li a, b, *
a
b
20/20 Laboratory for Fuel Cells and Green Energy RD&D, University of Waterloo, Waterloo, Ontario, Canada
State Key Laboratory of Engines, Tianjin University, Tianjin, China
a r t i c l e i n f o
a b s t r a c t
Article history:
Received 24 February 2010
Accepted 7 June 2010
Available online 14 July 2010
Polymer electrolyte membrane fuel cell (PEMFC) has been recognized as a promising zero-emission
power source for portable, mobile and stationary applications. To simultaneously ensure high membrane
proton conductivity and sufficient reactant delivery to reaction sites, water management has become one
of the most important issues for PEMFC commercialization, and proper water management requires good
understanding of water transport in different components of PEMFC. In this paper, previous researches
related to water transport in PEMFC are comprehensively reviewed. The state and transport mechanism
of water in different components are elaborated in detail. Based on the literature review, it is found that
experimental techniques have been developed to predict distributions of water, gas species, temperature
and other parameters in PEMFC. However, difficulties still remain for simultaneous measurements of
multiple parameters, and the cell and system design modifications required by measurements need to be
minimized. Previous modeling work on water transport in PEMFC involves developing rule-based and
first-principle-based models, and first-principle-based models involve multi-scale methods from
atomistic to full cell levels. Different models have been adopted for different purposes and they all
together can provide a comprehensive view of water transport in PEMFC. With the development of
computational power, application of lower length scale methods to higher length scales for more
accurate and comprehensive results is feasible in the future. Researches related to cold start (startup
from subzero temperatures) and high temperature PEMFC (HT-PEMFC) (operating at the temperatures
higher than 100 C) are also reviewed. Ice formation that hinders reactant delivery and damages cell
materials is the major issue for PEMFC cold start, and enhancing water absorption by membrane electrolyte and external heating have been identified as the most effective ways to reduce ice formation and
accelerate temperature increment. HT-PEMFC that can operate without liquid water formation and
membrane hydration greatly simplifies water management strategy, and promising performance of HTPEMFC has been demonstrated.
Ó 2010 Elsevier Ltd. All rights reserved.
Keywords:
Polymer electrolyte membrane fuel cell
(PEMFC)
Water management
Water transport
Cold start
High temperature PEMFC (HT-PEMFC)
Contents
1.
2.
Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 224
1.1.
Fundamental principles . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 224
1.2.
Origin and importance of water management . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 226
1.3.
Strategy and impact of water management . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 227
1.4.
Scope and objective . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 228
State of water . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 229
2.1.
In membrane . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 230
2.2.
In gas diffusion layer . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 231
2.3.
In catalyst layer . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 232
2.4.
In flow channel . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 233
2.5.
Summary . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 233
* Corresponding author. 20/20 Laboratory for Fuel Cells and Green Energy RD&D, University of Waterloo, Waterloo, Ontario, Canada. Tel.: þ1 519 888 4567x36843; fax: þ1
519 888 6197.
E-mail address:
[email protected] (X. Li).
0360-1285/$ e see front matter Ó 2010 Elsevier Ltd. All rights reserved.
doi:10.1016/j.pecs.2010.06.002
222
3.
4.
5.
6.
7.
8.
9.
10.
11.
12.
K. Jiao, X. Li / Progress in Energy and Combustion Science 37 (2011) 221e291
Mechanism of water transport . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 233
3.1.
In membrane . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 233
3.1.1.
Proton transport . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 234
3.1.2.
Diffusion of water . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 236
3.1.3.
Electro-osmotic drag effect . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 237
3.1.4.
Hydraulic permeation . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 238
3.1.5.
Reactant transport . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 238
3.1.6.
Membrane expansion . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 238
3.2.
In gas diffusion layer . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 239
3.2.1.
Diffusion and convection . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 239
3.2.2.
Surface tension and wall adhesion effects in porous media: capillary effect . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 240
3.2.3.
Condensation and evaporation of water . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 242
3.3.
In catalyst layer . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 242
3.4.
In flow channel . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 244
3.5.
Summary . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 244
Experimental observation . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 244
4.1.
Current distribution measurement . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 244
4.2.
High frequency resistance distribution measurement . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 245
4.3.
Gas species concentration measurement . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 245
4.4.
Temperature distribution measurement . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 246
4.5.
Water visualization . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 246
4.6.
Summary . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 249
Overview of numerical models . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 249
5.1.
Level of scale . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 249
5.2.
History of model development . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 251
5.3.
Summary . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 252
Multi-dimensional multi-component multiphase model with full cell geometry . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 252
6.1.
Computational domain . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 253
6.2.
Two-fluid model . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 253
6.3.
Mixture model . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 255
6.4.
Boundary conditions and numerical implementation . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 257
6.5.
Two-fluid model vs mixture model . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 258
6.6.
Summary . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 260
Modeling water transport in membrane . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 260
7.1.
Macroscopic approach . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 260
7.1.1.
Diffusive model . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 260
7.1.2.
Chemical potential model . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 260
7.1.3.
Hydraulic model . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 261
7.1.4.
Combinational model . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 261
7.2.
Bottom-up approach and physical models . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 262
7.2.1.
Modeling ionomer self-organization . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 262
7.2.2.
Modeling proton transport . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 262
7.2.3.
Physical models with simplified membrane micro-structure . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 262
7.3.
Summary . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 263
Modeling water transport in gas diffusion layer . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 264
8.1.
Homogeneous approach . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 264
8.2.
Structure generation . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 264
8.3.
Volume-of-fluid model . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 265
8.4.
Lattice Boltzmann model . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 266
8.5.
Rule-based model . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 267
8.5.1.
Full morphology model . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 267
Pore-network model . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 268
8.5.2.
8.6.
Summary . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 268
Modeling water transport in catalyst layer . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 268
9.1.
Agglomerate model . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 269
9.2.
Molecular dynamics and office-lattice pseudo particle simulations . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 269
9.3.
Summary . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 270
Modeling water transport in flow channel . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .271
10.1.
Volume-of-fluid model . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 271
10.2.
Other aspects . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 272
10.3.
Summary . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 272
Cold start . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .273
11.1.
Experimental work . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 274
11.2.
Numerical model . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 274
11.3.
Cold start characteristics . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 275
11.4.
Summary . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 279
High temperature polymer electrolyte membrane fuel cell . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 280
12.1.
Experiential work . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 280
12.2.
Numerical model . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 281
12.3.
Performance . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 282
K. Jiao, X. Li / Progress in Energy and Combustion Science 37 (2011) 221e291
12.4.
Summary . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
Summary and outlook . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
Acknowledgments . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
References . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
13.
Nomenclature
a
A
Bo
c
C
Ca
Cp
d
D
DL
e
E
EW
f
F
g
h
H
I
j
J
j0
k
K
Kn
l
L
_
m
M
n
!
n
nd
!
nw
p
Q_
r
R
Re
RH
s
S
SH
t
T
T0
!
tw
!
u
U
V
We
x
water activity, specific area (m 1)
cell geometric area (m2)
bond number
molar concentration (kmol m 3)
electric capacity (F m 2)
capillary number
specific heat (J kg 1 K 1)
diameter (m)
mass diffusivity (m2 s 1)
doping level
velocity (m s 1)
electrical potential (V), energy (J mol 1)
equivalent weight of ionomer (kg kmol 1)
distribution function, pre-exponential factor
Faraday’s constant (96,487 C mol 1)
Gibbs free energy (J kg 1 or J kmol 1), gravity (m s 2)
enthalpy and latent heat (J kg 1 and J kmol 1),
surrounding heat transfer coefficient (W m 2 K 1)
Henry’s constant (Pa m3 kmol 1)
current density (A cm 2)
reaction rate (A m 3)
mass or molar flux (kg m 2 s 1 or kmol m 2 s 1)
volumetric exchange current density (A m 3)
thermal conductivity (W m 1 K 1)
permeability (m2)
Knudsen number
mean free path (m)
characteristic length (m)
mass flow rate (kg s 1)
molecular weight (kg kmol 1)
number of moles of electron for per mole of hydrogen
or oxygen
unit vector normal to interface
electro-osmotic drag coefficient (H2O per Hþ)
unit vector normal to wall
pressure (Pa)
heat transfer rate (W)
pore radius (m)
universal gas constant (8.314 J mol 1 K 1), radius (m)
Reynolds number
relative humidity
entropy (J kg 1 K 1 or J kmol 1 K 1), volume fraction
source term, mass or heat transfer term, entropy
(J kg 1 K 1 or J kmol 1 K 1)
analogous to the Sherwood number
time (s)
temperature (K)
volume averaged cell temperature (K)
unit vector tangential to wall
velocity (m s 1)
velocity (m s 1)
electrical potential (V)
Weber number
position or coordinate (m)
X
Y
z
. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 283
. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 283
. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 284
. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 284
mole fraction
mass fraction
charge number or valence
Abbreviation
AFC
alkaline fuel cell
BP
bipolar plate
BGK
Bhatnagar–Gross–Krook
CCD
charge-coupled device
CFD
computational fluid dynamics
CG
coarse grained
CGMD coarse grained molecular dynamics
CL
catalyst layer
CO
carbon monoxide
DSC
differential scanning calorimetry
DMFC
direct methanol fuel cell
DPD
dissipative particle dynamics
EIS
electrochemical impedance spectroscopy
EOD
electro-osmotic drag
FM
full morphology
GDL
gas diffusion layer
GC
gas chromatograph
HFR
high frequency resistance
HOR
hydrogen oxidation reaction
HT-PEMFC high temperature polymer electrolyte membrane
fuel cell
IR
infrared
LB
lattice Boltzmann
LG
lattice gas
MC
Monte Carlo
MCFC
molten carbonate fuel cell
MD
molecular dynamics
MEA
membrane electrode assembly
MPL
micro porous layer
NMR
nuclear magnetic resonance
ORR
oxygen reduction reaction
PAFC
phosphoric acid fuel cell
PN
pore network
PBI
polybenzimidazole
PEM
polymer electrolyte membrane
PEMFC polymer electrolyte membrane fuel cell
PFSA
perfluorosulfonic acid
PTFE
polytetrafluoroethylene
SOFC
solid oxide fuel cell
VOF
volume-of-fluid
Greek letters
transfer coefficient, water transport coefficient
(kmol2 J 1 m 1 s 1)
b
convective transport coefficient
g
water phase change rate (s 1)
G
uptake coefficient
d
thickness or average distance (m), time interval (s)
P
chemical potential (J kmol 1)
3
porosity
z
water transfer rate (s 1)
a
223
224
K. Jiao, X. Li / Progress in Energy and Combustion Science 37 (2011) 221e291
h
q
w
i
k
l
m
n
x
r
s
2
s
Y
f
u
efficiency, overpotential (V)
contact angle ( )
mobility (m2 kmol J 1 s 1)
interfacial drag coefficient
electrical conductivity (S m 1), surface curvature (m
water content in ionomer
dynamic viscosity (kg m 1 s 1)
kinematic viscosity (m2 s 1)
stoichiometry ratio
density (kg m 3)
surface tension (N m 1)
condensation and evaporation rate coefficients
tortuosity, relaxation time
relative mobility
electrical potential (V)
volume fraction of ionomer in catalyst layer
1
)
Subscripts and superscripts
0
intrinsic value
a
anode
act
activation
air
air
B
bulk or binary
BP
bipolar plate
c
cathode, capillary
cap
capillary
ce
condensation/evaporation
cell
cell characteristic
CL
catalyst layer
conc
concentration
cond
condensation
cs
charged site
desb
desublimation
diff
diffusive
eff
effecitive
ele
electronic
EOD
electro-osmotic drag
equil
equilibrium
evap
evaporation
f
frozen
F
Fickian
fl
fluid phase
FPD
freezing point depression
fmw
frozen membrane water
g
gas
1. Introduction
1.1. Fundamental principles
A fuel cell is an energy conversion device that converts the
chemical energy stored in fuels and oxidants into electricity
through electrochemical reactions. With different kinds of electrolyte, fuel cells can be classified into different types. The most
common types of fuel cells are the polymer electrolyte membrane
(also called the proton exchange membrane) fuel cell (PEMFC),
direct methanol fuel cell (DMFC) (the same as PEMFC but uses
methanol instead of hydrogen as the fuel), alkaline fuel cell (AFC),
phosphoric acid fuel cell (PAFC), molten carbonate fuel cell (MCFC)
and solid oxide fuel cell (SOFC). Different types of fuel cells are
suitable for different applications. Some types of fuel cells are most
suitable for stationary power generation, such as PAFC, MCFC and
SOFC, and some other types of fuel cells are mostly used for
GDL
H2
H2 O
hyd
I
i-g
i, j
ice
in
ion
K
l-i
lattice
lq
m
mem
N
n-f
n-i
n-v
nf
nmw
O2
ohm
open
out
pore
ref
r
sat
sl
sld
surf
surr
t
T
th
u
v-i
v-l
vp
vp/lq
wall
gas diffusion layer
hydrogen
water
hydraulic
intro
gas phase in species i
the ith and jth components
ice
inlet
ionic
Knudsen
liquid water to ice (vice versa)
lattice
liquid water
mass (for source term)
membrane or ionomer
normal condition
non-frozen membrane water to frozen membrane
water (vice versa)
non-frozen membrane water to ice
non-frozen membrane water to vapour (vice versa)
non-frozen
non-frozen membrane water
oxygen
ohmic
open circuit
outlet
pore
reference state
reversible
saturation
solid
solid
surface or interface
surroundings
time
energy (for source term)
thermodynamic
momentum (for source term)
vapour to ice
vapour to water liquid (vice versa)
water vapour
interface between vapour and liquid
surrounding wall of the cell
vehicular or portable applications, such as PEMFC, DMFC and AFC.
Benefiting from the advantages such as low operating temperature,
high power density and zero/low emission, PEMFC has increasingly
become the most promising candidate as the power source for
automotive and backup power applications. PEMFC is also the most
popular one under research and development compared with other
types of fuel cell. However, further improvements in terms of
performance, durability and cost are necessary before commercial
application.
Fig. 1 shows the schematics of a single PEMFC and a PEMFC stack
with three single cells. Typically a single PEMFC (Fig. 1a) consists of
an anode and a cathode, and a polymer electrolyte membrane
(PEM) in between. At the anode, hydrogen flows into the flow
channel through the gas diffusion layer (GDL) to the catalyst layer
(CL). In the anode CL, hydrogen splits into protons (hydrogen ions)
and electrons. The protons pass through the membrane and travel
to the cathode CL, however, the electrons cannot pass through the
K. Jiao, X. Li / Progress in Energy and Combustion Science 37 (2011) 221e291
225
Fig. 1. Schematics of (a) a single PEMFC and (b) a PEMFC stack with three single cells.
Reversible voltage
1.2
Open circuit voltage is less than reversible voltage
due to fuel crossover and internal currents
Cell V oltag e, V
1
Voltage initially falls fast mainly
due to activation loss
Voltage falls more linearly and slowly
mainly due to ohmic loss
0.8
0.6
0.4
membrane, but travel through an external circuit to the cathode,
thus generating electricity. At the same time, on the cathode side,
air or oxygen flows into the flow channel through the GDL to the CL.
In the cathode CL, oxygen reacts with the protons and electrons
from the anode, producing water and heat. Due to the water
concentration and pressure differences between the anode and
cathode, and the proton transport across the membrane, water can
travel through the membrane in both directions. On the anode side,
the reaction that hydrogen splits into protons and electrons is
a hydrogen oxidation reaction (HOR):
H2 /2Hþ þ 2e
On the cathode side, the reaction that oxygen, protons and
electrons form water is an oxygen reduction reaction (ORR):
Voltage falls faster at high currents
mainly due to concentration loss
0.2
0
0
0.2
0.4
0.6
(1)
0.8
1
-2
Current Density, A cm
Fig. 2. Sample polarization curve of a single PEMFC.
1.2
1
O þ 2Hþ þ 2e /H2 O
2 2
(2)
The overall reaction is simply hydrogen reacting with oxygen
producing water, electrical energy and heat:
226
K. Jiao, X. Li / Progress in Energy and Combustion Science 37 (2011) 221e291
1
H2 þ O2 /H2 O þ Electrical Energy þ Heat
2
(3)
Single cells are often connected in series to form a PEMFC stack
to produce higher voltages. As shown in Fig. 1b, with three single
cells connected in series, the supplied gases have to be distributed
into the single cells through the inlet manifolds, and the exhaust
gases have to be removed through the outlet manifolds.
Fig. 2 presents a sample currentevoltage graph of a single PEMFC,
which is also called the polarization curve. This figure shows the
voltage outputs at different current outputs. The current has been
normalized by the area of the fuel cell in terms of a standard unit of
current density (A cm 2), because PEMFCs with different sizes
obviously produce different currents, and a normalized current unit
makes fuel cell performance comparable. Theoretically, a PEMFC
would supply any amount of current under the condition of sufficient fuel supply, while maintaining a constant voltage determined
by thermodynamics. In practice, however, the actual voltage output
of a fuel cell is less than the ideal thermodynamically predicted value
(reversible voltage). Furthermore, the more current that is drawn
from a real fuel cell, the lower the voltage output of the cell, limiting
the total power that can be delivered.
The reversible (theoretically maximum possible) cell potential,
Er (V), can be calculated as
Er ¼
Dg
(4)
nF
where Dg (J mol 1) is the change of Gibbs free energy of the overall
reaction per mole of hydrogen, n (¼ 2) the number of moles of
electrons transferred per mole of hydrogen, and F
(¼96,487 C mol 1) the Faraday’s constant. It should be noticed that
the values of Dg are different for different reaction products (e.g.
liquid water and vapour), and the product water is typically in
liquid form at the PEMFC operating condition [1]. Since the change
of Gibbs free energy (Dg, J mol 1) represents the maximum useful
work available from the reaction, the thermal (theoretically
maximum possible) efficiency, hth, can be defined as
hth ¼
Dg
¼ 1
Dh
T Ds
Dh
1
(5)
where Dh (J mol ) is the change of enthalpy of the overall reaction
per mole of hydrogen (the total energy change of the reaction),
T the temperature (K), and Ds (J mol 1 K 1) the change of entropy of
the overall reaction per mole of hydrogen. The total amount of
energy transformed into heat at the reversible cell voltage is TDs,
which is also called the reversible heat. By using Equation (5), the
thermal efficiency can be calculated to be 83% at 25 C with product
water in liquid state. Practical fuel cells can have efficiency close to
this theoretical maximum efficiency, while practical heat engines
are very difficult to achieve even about 50% of the theoretical
maximum (or Carnot) efficiency [1].
It is difficult to maintain the cell voltage at a high level under
current load. As shown in Fig. 2, the voltage output of a PEMFC in
operation is less than the reversible voltage due to the irreversible
losses. The total loss increases with the increment of the current
density. There are four major types of fuel cell losses [2]: the loss
due to fuel crossover and internal currents, activation loss, ohmic
loss, and mass transport or concentration loss. The fuel crossover
and internal currents are the waste of fuel passing and electron
conduction through the electrolyte, respectively. The electrolyte
should only transport protons, however very small amount of fuel
diffusion and electron flow will always be possible. It does have
a marked influence on the open circuit voltage (the cell voltage at
zero current), which explains why the open circuit voltage is always
smaller than the reversible voltage. However, this type of loss
reduces considerably when a meaningful amount of the current is
drawn from the cell. The activation loss is caused by the slowness of
the reactions taking place on the surface of the electrodes. A
proportion of the voltage generated is lost in driving the electrochemical reaction that transfers the electrons and protons to or
from the electrode. In Fig. 2, it is represented by the initial sharp
drop of the cell voltage. The ohmic loss is the straightforward
resistance to the transport of electrons and protons through the
materials of the electrodes, membranes and the various interconnections. This voltage drop is essentially proportional to current
density, represented by the almost linear fall in the middle of the
performance curve in Fig. 2. The concentration or mass transport
loss results from the change in concentration of the reactants at the
surface of the electrodes as they are consumed along the flow
channel from the inlet to the outlet. Concentration affects voltage
via the change of differential pressure of reactant. That is why this
type of irreversibility is called concentration loss. On the other
hand, since the reduction in concentration is the result of a failure
to transport sufficient reactant to the electrode surface or catalyst
sites, this type of irreversibility is also called mass transport loss. In
Fig. 2, such loss can be observed at high current density range as
a nonlinear rapid drop, because sufficient reactant supply is the
controlling factor to obtain large amount of current. However, even
with sufficient reactant supply, the water flooding in the cell caused
by improper water removal may also result in concentration or
mass transport loss due the blockage of the reaction sites.
The operating voltages of a PEMFC at different current densities
can be calculated by using the reversible voltage subtracting all the
voltage losses due to the irreversibilities, and because the open
circuit voltage is equal to the reversible voltage subtracting the
voltage loss due to the fuel crossover and internal currents, therefore a relationship between the operating cell voltage and open
circuit voltage can be obtained
Ecell ¼ Eopen
hohm
hact
hconc
(6)
where Ecell and Eopen are the operating and open circuit voltages,
respectively; and hact, hohm and hconc represent the voltage losses
due to the activation loss, ohmic loss and concentration or mass
transport loss, respectively. As mentioned before, the voltage loss
due to fuel crossover and internal currents reduces considerably
when a meaningful amount of current is drawn from the cell
(Eopen z Er). The operating voltage is usually lower than 0.8 V when
drawing a useful current. Therefore, as mentioned earlier in this
section, many single cells have to be connected in series to form
a PEMFC stack to produce a higher voltage (as demonstrated in
Fig. 1b).
1.2. Origin and importance of water management
The water management issue has been going with PEMFCs since
the initial development by General Electric Company in 1950s
[3e5] and the first practical application for U.S. Gemini space
missions from 1962 to 1966 [6]. Until nowadays, tremendous
studies in this area are still continuing for achieving better
performance.
The major cause of the water management issue is the proton
conductor e the membrane. The mostly widely used type of
membrane as the proton conductor is the perfluorosulfonic acid
(PFSA) polymer membrane. The most famous one in the PFSA
family in the last thirty years is the Nafion membrane invented by
E.I. DuPont de Nemours due to its relatively high durability and
proton conductivity. Other PFSA polymer membranes like Aciplex,
Dow and Flemion have also been widely tested and used. Even
though the chemical structures are different for the different PFSA
polymer membranes, the membrane morphologies and the base
K. Jiao, X. Li / Progress in Energy and Combustion Science 37 (2011) 221e291
mechanisms of proton transport are similar. The PFSA polymer
membranes have acceptable proton conductivity only when they
are well hydrated. The electrical resistance increases when the
membrane is drying out, resulting in the increment of the ohmic
loss and heating. The increased local heating further accelerates
the local drying by evaporating more water, resulting in vicious
self-accelerated destruction of cell performance. The PFSA polymer
membranes are also unstable at high temperatures [1], therefore
the local heating also limits the lifetime of the membranes, hence
the cell lifetime as well. Since PFSA polymer membranes feature
the similar characteristics (water adsorption/desorption, proton
conduction etc.), only Nafion membrane is discussed in the major
part of this paper. Other types of membranes (such as hydrocarbon
membranes) have also attracted attentions, with the main objective of operating at higher temperatures and less relative humidity.
The development of such membranes is currently under active
research, and one of the most promising membranes, acid doped
polybenzimidazole (PBI) membrane that can tolerate higher
operating temperatures and has acceptable proton conductivity
without humidification, is discussed in Section 12.
As mentioned above, membrane hydration is of paramount
importance for PEMFC. In order to hydrate the membrane,
humidified reactant gases are often supplied. However, due to the
water production from the electrochemical reactions, as well as the
low operating temperature of PEMFC (typically around 80 C) that
leads to the almost unavoidable water condensation, liquid water
may be present and flood the electrode pore regions, giving rise to
the so-called water flooding phenomena, which severely reduce
the rate of reactant supply to the reaction sites and degrade the cell
performance considerably. The water flooding is most severe in
cathode CL due to the fact that water is produced in this layer and
the electro-osmotic drag (EOD) also causes water migrating form
anode CL to cathode CL. In fact, the concentration or mass transport
loss described previously is mostly likely caused by the water
flooding of cathode electrode. On the other hand, the EOD effect
also causes drying out of the portions of membrane close to
the anode side, resulting unevenly distributed water across the
membrane and large ohmic loss close to the anode side. On the
stack level, the supplied water (vapour or liquid) may be unevenly
distributed to each cell, causing flooding in some cells and drying
out in others. Therefore, proper membrane hydration without
causing electrode flooding by water, commonly referred to as water
management, remains one of the major technical challenges of
PEMFC, and maintaining the dynamic balance of water in the cell
during operation is essential for better water management and
stable performance.
1.3. Strategy and impact of water management
The simplest way to avoid the water flooding is to operate
PEMFC on non-humidified inlet gases with acceptable performance
degradation. Without humidifying the inlet gases, the humidification subsystem can be cast off, and the water and heat removal as
well as the mass transport of reactant gases are all improved due to
the fact that water is mostly likely in the vapour form. Utilizing the
product water for membrane hydration and balancing the water in
the cell are the key factors for operating PEMFC on non-humidified
gases. However, a meaningful relative humidity of the gas stream in
the cathode for membrane hydration can only be obtained if the
cell temperature is lower than 60 C, and operating the cell at
elevated temperatures (e.g. 80 C) results in severe reduction of the
relative humidity of the gas stream and hence reduces the
membrane hydration level significantly [2,7]. The idea of balancing
the water in the cell is to utilize the water produced in the cathode
CL to hydrate the membrane close to the anode side because water
227
can travel through the membrane driven by concentration or
pressure gradients. Since the relative humidity of the gas stream in
the cathode is often the highest at flow channel exit due to the
accumulation of product water, the counter flow arrangement by
placing the anode and cathode flows in opposite directions is
widely used to enhance the water transport from cathode to anode
[2,7]. However, the EOD causing water migrating from anode CL to
cathode CL becomes fast at high current densities and leads to
severe dehydration of the portions of membrane close to the anode
side, and hence further humidification is needed rather than just
unitizing the water produced in cathode. For the operating pressure
range from 1 to 3 atm, it has been shown that the cell performance
when operating on non-humidified inlet gases is about 20e40%
lower than operating on fully humidified (the relative humidify is
100%) inlet gases for the operating temperatures lower than 60 C,
and the performance degradation is expected to be much more
severe at elevated operating temperatures [7]. Since both the
electrochemical reaction kinetics and membrane conductivity can
be improved at higher temperatures, operating the cell at the
highest possible temperature while maintaining the membrane
durability is critical to achieve better performance, and therefore
the normal operating temperature for PEMFC is usually controlled
at about 80 C. At this level of operating temperature, especially
when a meaningful current is drawn from the cell, external
humidification becomes essential. Rather than utilizing the product
water, alternative ways to hydrate the membrane while operating
on non-humidified inlet gases were also demonstrated, such as
supplying water directly to the membrane through wicks [8] and
imbedding catalyst particles in the membrane to produce water
from the fuel and oxidants moved into the membrane [9,10].
However, none of these methods have been widely accepted
because more complicated cell/system design and operating
strategy are needed, which may also cause other problems.
As mentioned above, it is very difficult to achieve the “best”
performance when operating PEMFC without external humidification, therefore, external humidification of PEMFC by humidifying
the reactant gases inside or outside the stack becomes essential. In
this scenario, the water flooding mentioned earlier may become
a severe problem, and water management is clearly composed of
two interrelated issues: complete hydration of the membrane
electrolyte and product water removal e a dynamic balance of
water needs to be achieved to satisfy the two conditions.
External humidification of PEMFC can be achieved by employing
water columns, reactant gas recirculation, and direct water injection into either the anode or/and cathode compartment or into the
reactant gas streams outside the stack. For the water column
technique, process water is stored in a container and reactant gas is
introduced at the bottom of the container. The saturated reactant
gas stream leaves the container at or near the top of the container.
The amount of water that can be picked up by the gas stream and
brought into the stack therefore depends on the humidification
temperature. Low temperatures result in low water partial pressures, while high temperatures result in low reactant partial pressures. This technique works well for low reactant flow rate, such as
pure hydrogen stream. For air at high current density operations,
high air flow rate tends to carry small droplets with it and bring
them into the cathode electrode, causing water flooding; while air
stream itself may be unsaturated due to the short residence time,
resulting in membrane drying out locally. The reactant gas-recirculation design utilizes the water taken by the exit gas streams to
humidify both the anode and cathode inlet gases. Satisfactory stack
operation may be achieved without external process water.
However, this method requires an external piece of equipment,
often a compressor, to recirculate the gas streams, and the parasitic
power consumption may not be neglected for high recirculation
228
K. Jiao, X. Li / Progress in Energy and Combustion Science 37 (2011) 221e291
rates [1]. In the direct water injection design, water can be injected
in vapour (steam) or liquid form and injected directly into the
anode and cathode compartment, or to the fuel and oxidant
streams before their entry to the stack. Direct injection of steam
and liquid outside the stack are easy to accomplish, while into the
stack compartment directly requires complex designs, and the
components needed to achieve this tend to be expensive to
construct and difficult to incorporate into a practical stack. For
direct liquid water injection, more water can be introduced into the
reaction compartment to maintain the membrane hydration than
what is possible by just saturating the inlet gases by injecting
steam, the evaporation of liquid water also absorb waste heat
which facilitates the cell cooling. However, distributing the injected
liquid water evenly to each singe cell of a stack is difficult [11,12],
resulting in both the flooded and dry cells in a stack. All the external
humidification methods described here require extra equipment
for the system design and/or more complex stack design, which
increase both the cost and parasitic power consumption, and hence
need to be evaluated before implementing to practical applications.
Even though the membrane may be properly hydrated at the
normal operating temperature (around 80 C) with the external
humidification methods described above, the excessive water
(usually in liquid form) needs to be removed or balanced to avoid
the water flooding problem and to better hydrate the membrane.
The water removal and balancing are usually accomplished by
optimizing the stack/system design and operating condition.
Although various water management strategies have been
proposed, water is still typically removed by pumping air into
cathode flow fields. Therefore, the flow field design is critical for
water management. To the geometrical configurations of the flow
fields, a variety of different designs are known and the conventional
designs typically comprise pin-type [13,14], straight-parallel
[15,16], serpentine [17,18], integrated [19,20], interdigitated [21,22],
and porous metal/carbon foam flow fields [23,24]. Every single flow
field design has its own advantages and disadvantages, and modified flow fields by incorporating the advantages of the different
designs are often used. One example is that the straight-parallel
flow field design features low pressure drop from the channel inlet
to outlet but weak water removal ability, while the serpentine flow
field design features high pressure drop but effective water
removal, and therefore a so-called serpentine-parallel flow field by
using numbers of serpentine flow channels connecting in parallel
has been demonstrated for the compromised design [25,26].
Different designs of flow field are also suitable for different water
management systems, for example the interdigitated flow field has
been demonstrated to be the most suitable design when used with
direct liquid water injection [27]. More details about the design
considerations for flow field in PEMFC can be found in [28,29]. The
GDL and CL properties such as thickness, porosity, permeability,
wettability, catalyst loading (for CL) and ionomer fraction (for CL)
also have strong impacts on water removal [30e40]. One common
way is to add polytetrafluoroethylene (PTFE) to these layers to
increase the hydrophobic level to expel water, and inserting
a hydrophobic micro porous layer (MPL) between GDL and CL in
cathode was also found to be an effective way to push water from
cathode CL into membrane [31e34,41e45]. Even though it was
found that using MPL could improve the cell performance, the
actual function of this layer is still under debate. One explanation is
that using MPL between GDL and CL may reduce the electrical and
thermal contact resistances. Another explanation is as mentioned
above, more water can be retained in CL and membrane to improve
the proton conductivity. Therefore, MPL may act as electrical/
thermal transfer smoother and mass transfer resistance, and
further investigation is needed to find its impacts on actual transport phenomena. Although the Nafion membrane has been
recognized as the “standard” membrane for PEMFC, different
thicknesses of the Nafion membrane are available and therefore
need to be considered for water removal and balance. Optimizing
the operating condition by controlling the operating temperature
and pressure is also an effective way to remove and balance the
water. Generally, increasing the operating temperature and
decreasing the operating pressure may all reduce the amount of
liquid water formation, and vice versa [1,2]. However, the promotions for both the electrochemical kinetics and reactant transport
by increasing the operating temperature and pressure, the durability of the stack components, as well as the power consumptions
for cell cooling and pressurization all need to be considered. The
impervious cathode and anode flow distribution plates can also be
replaced by porous hydrophilic plates or something similar
[13,14,46e48]. With appropriate matching of the pore sizes among
the anode electrode, anode porous flow distribution plate, cathode
electrode and cathode porous flow distribution plate, a judicially
controlled set of pressure differences between the oxidant stream
and the adjacent cooling water stream and between the cooling
water and the fuel stream transfers the product water in the
cathode to the cooling water stream, then to the anode by the
pressure differentials between the various streams involved, and
from the anode to the cathode via EOD. The water flooding in the
cathode can be avoided while the portions of membrane close to
the anode side can be further hydrated. Precision monitoring and
control of the pressure differentials add the cost of the system and
operation and maintenance. It might be also possible to run the
anode and cathode at different pressures to accelerate the water
permeation from the cathode to anode through the membrane.
However, much larger pressure differential is required by
comparing to pressurizing the water to the cooling channel, and
therefore the mechanical strength of the MEA needs to be carefully
evaluated. Many other methods to remove and/or balance the
water are also available, such as using a liquid-permeable electricity-conductive layer for storage and transport of water [49],
using integrated EOD pumping to remove water [50], and
sequentially exhausting each cell of a stack to avoid single cell
flooding [51].
Apparently, many water management methods are available and
proper implementation is needed. In reality, a number of the water
management methods described here can be combined to their best
effects (e.g. [13,14,27,47e55]). Understanding of the water transport
in PEMFC is therefore critical to achieve the optimal water
management strategy, and hence the improved cell performance.
Some reviews on water management of PEMFC were published
recently [56e61]. Bazylak [56] reviewed the visualization techniques
for water distribution in PEMFC. The reviews conducted by Li et al.
[57] and Ji and Wei [61] focused on the diagnosis and mitigation of
water flooding in PEMFC. Studies on water transport in MEA were
reviewed by Dai et al. [58]. Yousfi-Steiner et al. [59] reviewed the
voltage degradation issues related to water management, and
Schmittinger and Vahidi [60] reviewed the durability issues related
to water management. Each of the reviews in [56e61] focused on
one or few specific topics of water transport in PEMFC, and the
details of state of water and mechanism of water transport in PEMFC
were only partially presented in these reviews due to the narrow
scopes. Therefore, it is timely to comprehensively summarize the
progress on the understanding of water transport in PEMFC.
1.4. Scope and objective
The main objective of the present paper is to summarize the
current status of understanding on water transport in PEMFC. The
various water transport processes in PEMFC are therefore elaborated in different sections by reviewing the experimental and
229
K. Jiao, X. Li / Progress in Energy and Combustion Science 37 (2011) 221e291
CL
Gas mixture
Gas mixture
Carbon support
Liquid water
Liquid water
Catalyst
Flow channel
Polymer electrolyte
Gas mixture
SO3−
Flexible perfluoracarbon
(gas permeable)
Solid material
Liquid water
GDL
Liquid water,
H+ and H3O+
Hydrophobic backbone
Membrane
a
CL
Gas mixture
Gas mixture
Carbon support
Ice
Catalyst
Flow channel
Polymer electrolyte
Gas mixture
−
SO3
Flexible perfluoracarbon
(gas permeable)
Solid material
Ice
Ice
GDL
Liquid water,
H+ and H3O+
Hydrophobic backbone
Membrane
b
Fig. 3. Schematics of a single PEMFC with the structure of each cell component illustrated (a: normal operating condition; b: cold start).
Normal Operating Condition:
In Membrane
Non-frozen
Membrane Water
In CL
Non-frozen
Membrane Water
In GDL and
Flow Channel
Vapour
Liquid
numerical studies. This article is organized as follows. The state and
transport mechanism of water are explained in Sections 2 and 3,
respectively; the water transport related experimental observations are reviewed in Section 4; the water transport related
numerical models are reviewed in Sections 5e10 including both the
first-principle-based and rule-based models, and both the topdown and bottom-up models are included for the first-principlebased models; the two special cases for starting from subzero
temperatures (cold start) and operating above the boiling point of
water (high temperature PEMFC (HT-PEMFC)) are introduced in
Sections 11 and 12, respectively; and finally the summary and
outlook is given in Section 13.
Vapour
Liquid
Cold Start:
In Membrane
Non-frozen
Membrane Water
Frozen
Membrane Water
In CL
Non-frozen
Membrane Water
Liquid
In GDL and
Flow Channel
Vapour
Liquid
Vapour
Ice
Ice
Fig. 4. Schematics of water phase change in different components of PEMFC for both
normal operating condition and cold start.
2. State of water
Due to the differences in materials and in local operating
conditions among the different components of PEMFC, water can
be present in different states with different phase change
processes. Many efforts in both the experiment and theory have
been paid to understand the state of water in PEMFC. Generally, the
230
K. Jiao, X. Li / Progress in Energy and Combustion Science 37 (2011) 221e291
state and phase change of water are different in membrane, CL, GDL
and flow channel. Fig. 3 shows the schematics of a single PEMFC
with the structure of each cell component illustrated for both the
normal operating condition and cold start, and the schematics of
water phase change are given in Fig. 4. As shown in Figs. 3a and 4
for normal operating conditions (cell temperature generally
ranges from 60 to 80 C), water exists in forms of vapour and liquid
in the flow channel and pore regions of GDL and CL (liquid water is
formed if the local water vapour pressure is higher than the local
water saturation pressure), the ionomer (polymer electrolyte) in
membrane and CL also absorbs certain quantities of water in liquid
state or bound to Hþ (e.g. H3Oþ). Due to the fact that PEMFC is often
considered to be used for automotive applications, and it is almost
unavoidable for vehicles driving below the freezing point of water
in winter, PEMFC must be able to successfully start up from subzero
temperatures, which is referred to as “cold start”. During a cold
start process for PEMFC, the initial cell temperature is usually equal
to the surrounding temperature (usually under 0 C in winter), and
water will mostly likely freeze in this kind of scenario. As shown in
Figs. 3b and 4 for PEMFC cold start, the formation of liquid water
can be almost neglected since it freezes to ice (note that ice and
liquid water may still co-exist when the local cell temperature
increases or decreases to the freezing point of water, resulting in ice
melting or liquid freezing). Therefore, water usually exists in forms
of vapour and ice in the pore regions of GDL and CL (ice is formed if
the local water vapour pressure is higher than the local water
saturation pressure) for PEMFC cold start. Since the ice formed can
easily stick on the solid materials of CL and GDL and difficult to
move, the ice formation in flow channel might be neglected. Figs.
3b and 4 also show that water in the ionomer of membrane and CL
may also freeze at subzero temperatures. Apparently, water in
PEMFC can create very complex scenarios, and the different states
of water need to be classified systematically for better understanding of water transport in PEMFC. In this section, the states of
water are classified and described in different components of
PEMFC to provide a general view of water to guide the discussions
in the following sections. Even though most of the previous studies
of PEMFC focused on the normal operating conditions, the subzero
temperature conditions (cold start) are also considered in this
section to provide a complete view of the state of water. After this
section, the normal operating conditions are focused on and the
details of the cold start processes are explicitly discussed in Section
11. Note that unless otherwise specified, the water content
mentioned starting from Section 3 (except Section 11) represents
the non-frozen water content, since no frozen water is present in
ionomer in normal operating conditions. The states of water in
different components of PEMFC are presented in Table 1, and the
detailed descriptions are given in the following subsections.
Table 1
State of water in each component of PEMFC.
Cell component
Specific location/
material
State of water
Membrane
Ionomer
GDL
Pore region
Free, Freezable, Non-freezable (the free
and freezable water are all possible to
freeze at subzero temperatures)
Vapour, Liquid, Icea
CL
Pore region
Ionomer
Vapour, Liquid, Icea
Free, Freezable, Non-freezable (the free
and freezable water are all possible to
freeze at subzero temperatures)
Flow channel
Everywhere
Vapour, Liquid, Icea (the ice formation
might be neglected)
a
Only present at subzero temperatures.
2.1. In membrane
As shown in Fig. 3, the ionomer of membrane and CL consists of
hydrophobic backbones, flexible perfluorocarbons (gas permeable),
and hydrophilic clusters with HþSO3 (the region in the middle of the
ionomer structure demonstrated in Fig. 3). The SO3 are bound to the
material structures and difficult to move, and there is an attraction
between the Hþ and SO3 for each HþSO3 . The hydrophilic clusters
with HþSO3 can absorb large quantities of water to form hydrated
hydrophilic regions. In the hydrated hydrophilic regions, the Hþ are
relatively weakly attracted to the SO3 and can move more easily. The
hydrated hydrophilic regions can be considered as dilute acids,
explaining why the membrane needs to be well hydrated (hydrated
regions must be as large as possible) for appreciable proton conductivity, and the SO3 can be considered as the proton exchange sites
since the Hþ often move between the SO3 . The water absorption level
of ionomer is often represented as the number of water molecules per
SO3 referred to as the water content (l). In a well hydrated Nafion
membrane, there will be about 20 water molecules for each SO3 , and
the proton conductivity can reach higher than 10 S m 1. The thickness
of the present Nafion membrane ranges from 25 (Nafion 211) to 175
(Nafion 117) mm, and the size of the hydrophilic region that can
contain water is on the level of nanometre. The water concentration
ðcH2 O ; kmol m 3 Þ inside the ionomer of membrane and CL can be
correlated with the water content (l):
l ¼
EW
rmem
cH2 O
(7)
where rmem (kg m 3) is the density of dry membrane (ionomer), and
EW (kg kmol 1) the equivalent weight represented by the dry mass
of the membrane (ionomer) over the number of moles of SO3 [1]:
EW ¼
Dry ionomer mass in g
Mole of proton exchange sites SO3
(8)
Generally, with high EW, the mechanical and thermal strengths
of membrane are high; and with low EW, the number of proton
exchange sites is high, resulting in high proton conductivity. EW is
usually equal to 1100 kg kmol 1 (Nafion 112, 115 and 117) or
2100 kg kmol 1 (Nafion 211 and 212) for Nafion membranes. In
considering all the operating conditions (normal and cold start) and
in the hydrophilic regions, liquid water, water bound to Hþ (e.g.
H3Oþ) and ice are all possible to be present. The most suitable
classification of the water in ionomer is non-frozen water and
frozen water [62,63], which is based on the observations of the
freezing behaviour of water by differential scanning calorimetry
(DSC) and nuclear magnetic resonance (NMR) [64e68]. DSC has
been used to determine the amounts of the different types of water,
and the maximum allowed amount of non-frozen water in Nafion
membrane at different subzero temperatures has been reported. It
has been found that there is a certain amount of water (about 4.8
water content) that does not freeze [64]. The non-frozen water can
be further classified into non-freezable, freezable and free water
[69e71], based on how tight they are bound to the sulphuric acids
(HþSO3 ). The non-freezable water is mostly tightly bound to
HþSO3 , and its maximum allowed amount is about 4.8 [64]. The
freezable water is loosely bound to HþSO3 and exhibits freezing
point depressions, which has been observed in [64e68]. The free
water may also appear if the water content is sufficiently high,
confirming the observations that the water freezes in ionomer at
the temperatures close to the normal freezing temperature of water
(0 C) if the water content is high [64e68]. The nanometre
confinement of water in the small pores may also lead to freezing
point depressions due to the enhanced surface dynamics of water.
Therefore the free water may still possess slightly lower freezing
K. Jiao, X. Li / Progress in Energy and Combustion Science 37 (2011) 221e291
points than bulk water. An empirical correlation to calculate the
maximum allowed amount of non-frozen water content has been
given in [62,63] based on the experimental measurement in [64]:
lsat
8
>
<¼
¼
>
:
if T < 223:15 K
h4:837
1:304 þ 0:01479T 3:594 10
5T2
if 223:15 K T < TN
if T TN
> lnf
i
1
(9)
where lsat and lnf are the saturation (maximum allowed nonfrozen water content) and non-frozen water content; and T and TN
are the local temperature and the normal freezing temperature of
water (273.15 K), respectively. The units of T and TN are all in K. It
should be noticed that no further water phase change was
detected when the temperature is lower than 50 C [64], and the
saturation water content remains at about 4.8. This amount of
water content corresponds to the non-freezable water. For the
temperature range from
50 to 0 C (223.15e273.15 K),
the maximum allowed amount of non-frozen water increases with
the increment of temperature. Moreover, water in ionomer does
not freeze if the temperature is higher than the normal freezing
temperature of water (T TN), therefore the saturation water
content is always higher than the non-frozen water content in the
ionomer in this temperature range. The difference between the
maximum allowed water content and the local non-frozen water
content can be considered as the driving force for the water phase
change in ionomer, i.e. if lnf is larger than lsat, water will freeze
until the local equilibrium state is reached (lnf ¼ lsat). The proton
conductivity of Nafion ionomer is usually calculated based on the
correlation reported in [72], and by considering the non-frozen
water content rather than the total water content, the correlation
becomes
kion ¼ 0:5139lnf
1
0:326 exp 1268
303:15
1
T
(10)
where kion (S m 1) is the proton conductivity and T (K) the
temperature. It should be noticed that Equation (10) is originally
correlated based on experimental measurements between 30 and
80 C, and the original equation overestimates the ion conductivity
at subzero temperatures (sharper decrements of the conductivity
were observed [64e66]). However, by considering the freezing of
membrane water at subzero temperatures, Equation (10) with nonfrozen membrane water content rather than total membrane water
Membrane Conductiv ity, S m
-1
14
λ = 15
12
10
8
λ = 10
6 Freezing points
4
λ=5
2
0
-50
0
50
100
Temperature, oC
Fig. 5. Effects of water freezing on proton conductivity of Nafion ionomer [62].
231
content provides a more reasonable agreement with the experimental measurements in [64]. Fig. 5 shows the changes of proton
conductivity at different total membrane water contents, it can be
observed that the proton conductivity starts dropping fast when
the membrane water starts freezing.
The non-frozen water in ionomer can also be classified into
surface and bulk water [73,74], or vapour and liquid [75e77]. The
classification of surface and bulk water is related to the strength of
the interactions between water and HþSO3 . Surface water strongly
interacts with HþSO3 and is mostly likely present close to the SO3
[78]. Bulk water is mainly identified as liquid which are loosely
bound to SO3 . The surface and bulk water, and the non-freezable,
freezable and free water might be able to be interrelated. The classification into vapour and liquid water in membrane is actually an
assumption based on the state of water in the adjacent phase outside
the membrane. In this paper, the classification of water into nonfreezable, freezable and free water is used since it comprehensively
covers all the operating conditions and is easy to understand. In
addition, the state and phase change of water in the different PFSA
polymer membranes features the similar characteristics, and only
water in Nafion membrane is discussed in this paper. For PEMFC
with the other kinds of membranes, the state of water as well as the
transport processes could be different, e.g. HT-PEMFC with PBI based
membranes is discussed explicitly in Section 12.
2.2. In gas diffusion layer
The GDL is usually carbon paper or carbon cloth, and the
porosity is around or higher than 0.5. Water can exist in the pore
regions of GDL in forms of vapour, liquid and ice. As shown in Figs.
3a and 4 under normal operating conditions, water exists in forms
of vapour and liquid, and the condensation and evaporation occur
depending on the local operating condition. For cold start as presented in Figs. 3b and 4 and as mentioned earlier, the formation of
liquid water can be almost neglected since it freezes to ice, and it
should be noticed that ice and liquid water may still co-exist when
the local cell temperature increases or decreases to the freezing
point of water, resulting in ice melting or liquid freezing. Therefore,
water usually exists in forms of vapour and ice in the pore regions
of GDL (ice is formed if the local water vapour pressure is higher
than the local water saturation pressure) for PEMFC cold start. The
pressure inside PEMFCs is usually between 1 to several atm, and the
freezing point of water in this pressure range can be treated as
constant (0 C). Because the operating pressure of PEMFC is always
equal to or greater than the atmospheric pressure, the sublimation
process (phase change from ice directly to vapour) can be safely
neglected. The difference between the local temperature and the
freezing temperature of water (around 0 C in GDL) indicates the
phase change direction between liquid and ice: ice will melt to
liquid if the local temperature is higher than the freezing point, and
vice versa, and liquid and ice may co-exist during such phase
change processes. The difference between the water saturation
pressure and vapour pressure indicates the phase change direction
between vapour and liquid (above the freezing point), and between
vapour and ice (below the freezing point). When the local
temperature is above the freezing point of water, if the vapour
pressure is higher than the saturation pressure, vapour will
condense to liquid, otherwise liquid will evaporate to vapour; when
the local temperature is below the freezing point of water, if the
vapour pressure is higher than the saturation pressure, vapour will
desublimate to ice, however, as mentioned earlier that ice will not
sublimate to vapour under the operating conditions in PEMFC even
when the vapour pressure is lower than the saturation pressure. By
checking the experimental data tabulated in [79], it is found that
the correlation provided by Springer et al. [72] provides acceptable
232
K. Jiao, X. Li / Progress in Energy and Combustion Science 37 (2011) 221e291
agreement with the experimental data in the temperature range
from 50 to 100 C:
2:1794 þ 0:02953ðT
Saturation water content
273:15Þ
9:1837 10
5
ðT
273:15Þ2
þ 1:4454 10
7
ðT
273:15Þ3
15
ð11Þ
where psat (Pa) is the saturation pressure of water and T (K) the
temperature.
2.3. In catalyst layer
Water Content
log10 ðpsat =101; 325Þ ¼
20
10
lequil ¼
0:043 þ 17:81a 39:85a2 þ 36:0a3 if 0 a 1
14:0 þ 1:4ða 1Þ if 1 < a 3
Equilibrium water content
0
-30
-25
-20
-15
-10
-5
0
o
Temperature, C
Fig. 6. Temperature dependence of the equilibrium water content (lequil) for the phase
change between the non-frozen water in the ionomer of CL and the vapour in the pores
of CL, and the saturation water content (lsat) for the phase change between the nonfrozen water in the ionomer and the ice in the pores of CL or the frozen water in the
membrane layer for the temperature range from 30 to 0 C (note that in the calculation of equilibrium water content shown in this figure, the water vapour amount in
the pores of CL is assumed to be equivalent to a water vapour activity of 1 at 30 C
and liquid water is not considered because it freezes to ice at subzero temperatures;
also note that the two lines, equilibrium water content and saturation water content,
separate this figure into four regions as marked in the figure, and for the non-frozen
water content in the ionomer, lnf, falling in these different regions, the water phase
change processes are different, as detailed in Table 2) [63].
between the non-frozen membrane water and vapour occurs when
non-humidified inlet gases are supplied [62,63]; only the phase
change between the non-frozen membrane water and liquid occurs
in cathode CL with fully humidified inlet gases [80]; and only the
phase change between the non-frozen membrane water and
vapour occurs in anode CL no matter the inlet gases are humidified
or not [62,63,80], due to the fact that water is not produced in
anode and the EOD effect dries out the anode.
Based on the assumptions mentioned above, Fig. 6 shows the
temperature dependence of the equilibrium water content (lequil)
for the phase change between the non-frozen water in the ionomer of CL and the vapour in the pores of CL, and the saturation
water content (lsat) for the phase change between the non-frozen
water in the ionomer and the ice in the pores of CL or the frozen
water in the membrane layer for the temperature range from 30
to 0 C. Note that in the calculation of the equilibrium water
(12)
where lequil is the equilibrium water content (the water content in
ionomer corresponding to the amount of surrounding water), and
a the water activity in the pore regions, defined as
Xvp pg
a ¼
þ 2slq
psat
Intersection point
3
4
5
The most complex scenario occurs in CL. As shown in Fig. 3, the
carbon and platinum particles together with part of the ionomer
(polymer electrolyte) have to mix together to form the reaction
sites. The carbon and platinum particles have to be present as the
catalyst and for electron transport, and the ionomer has to be
present for proton transport. The volume fraction of the ionomer in
CL ranges from about 0.2 to 0.4, and the porosity of the CL ranges
from about 0.2 to 0.5. The thickness of CL is typically around
0.01 mm. The state of water in the pore regions of CL is the same as
in the pore regions of GDL. The water in the ionomer of CL is also
the same as in the membrane. It should be mentioned that due to
the presence of small pores in CL (on the level of nanometre), the
surface dynamics of water is enhanced, and the freezing point of
water may be depressed to about 1 C in such small pores [62,63].
In the CL and at the interface between the CL and membrane,
both the liquid water and vapour can be desorbed or absorbed by
the ionomer to or from the pore regions of CL. The non-frozen
membrane water, liquid or vapour can freeze or desublimate to ice,
as shown in Fig. 4. The different water phase change processes can
occur simultaneously or separately. It should be noticed that the ice
in CL in Fig. 4 represents the ice in both the ionomer and pore
regions. To simplify the phase change processes, one assumption
made in the previous studies [62,63] is that when the water in the
ionomer of CL freezes, it leaves the ionomer and only forms ice in
the pore regions of CL, based on the experimental observation that
the frozen water does not contribute to the proton transport in
ionomer [64]. Equation (9) can be used to determine whether the
water in the ionomer freezes to ice or not (if the non-frozen water
content is higher than the saturation water content, it will freeze).
The equation first used in [72] can be used to indicate the mass
transfer (phase change) direction between the water in the ionomer and the vapour/liquid in the pore regions:
2
1
(13)
where Xvp is the mole fraction of water vapour in the pore regions,
pg (Pa) the pressure of the gas mixture in the pore regions, psat (Pa)
the water saturation pressure, and slq the liquid water volume
fraction in the pore regions.
As mentioned in Section 1.3, when non-humidified gases are
supplied, the presence of liquid water may be neglected, on the
other hand, liquid water will mostly likely be present in cathode CL
when fully humidified gases are supplied. Based on this premise,
further assumptions to simplify the water phase change processes
between the non-frozen membrane water and vapour/liquid have
been made [62,63,80]. It is assumed that only the phase change
Table 2
Water phase change processes that involve the ionomer in CL for the non-frozen
water content in the ionomer (lnf) in the different regions and on the different
dividing lines of Fig. 6 [63].
For the non-frozen water
content in the different
regions of Fig. 6
Phase change between
the non-frozen water
in the ionomer and
the vapour in the pores
The non-frozen water
in the ionomer freezing
to the ice in the pores
In Region 1
In Region 2
In Region 3
In Region 4
On the line between
Regions 1 and 2
On the line between
Regions 2 and 3
On the line between
Regions 3 and 4
On the line between
Regions 4 and 1
At the intersection point
Occur
Occur
Occur
Occur
Stop
Occur
Occur
Stop
Stop
Occur
Occur
Stop
Stop
Stop
Occur
Stop
Stop
Stop
233
K. Jiao, X. Li / Progress in Energy and Combustion Science 37 (2011) 221e291
6
water and vapour can co-exist, and the condensation and evaporation can happen based on the difference between the vapour and
saturation pressures (as described in Section 2.2); for cold start
conditions, since water usually freezes in CL and GDL first and stick
on the solid materials that is difficult to move, and usually nonhumidified gases are supplied for PEMFC cold start to reduce the
amount of ice formation, therefore the chance of ice formation in
flow channel is much lower than in CL and GDL, and might be
neglected.
4
2.5. Summary
2
The states of water in different components of PEMFC are
elaborated in Section 2 and summarized in Table 1. In the ionomer of membrane and CL, water is classified into free, freezable
and non-freezable water, based on how tight they are bound to
the sulphuric acids. In the pores of GDL and CL, and in flow
channel, both vapour and liquid states are possible to be present.
At subzero temperatures, water is able to freeze in all the
components. The freezing points of water in the pores of GDL and
CL and in the ionomer of membrane and CL are depressed due to
the enhanced surface dynamics in the small holes. The freezing
point of water in the ionomer of membrane and CL is further
lowered because water is bound to the sulphuric acids. Even
though water is classified into different types, whether phase
equilibriums of water exist in different components remain
debated. This is mainly due to the presence and arbitrary transport of liquid water, especially in the heterogeneous structures of
CL and GDL.
Equilibrium Water Content
14
Equation 12 [72]
Measurement at 30 oC [81]
o
Measurement at 80 C [82]
12
10
8
0
0
0.2
0.4
0.6
0.8
1
Water Activity
Fig. 7. Equilibrium water content as a function of water vapour activity for Nafion
membrane at temperatures of 30 C and 80 C [72,81,82].
content shown in this figure, the water vapour amount in the
pores of CL is assumed to be equivalent to a water vapour activity
of 1 at 30 C and liquid water is not considered because it freezes
to ice at subzero temperatures. In Fig. 6, the equilibrium water
content decreases monotonically with temperature, because the
water vapour amount in the pores of CL has been assumed to be
saturated at 30 C, correspondingly the water vapour becomes
more and more unsaturated as temperature is increased, leading
to the decrease in equilibrium water content. On the other hand,
the saturation water content increases monotonically with
temperature since more non-frozen water can be maintained in
the ionomer to avoid freezing. The two curves in Fig. 6 divide the
figure into four regions, and for the non-frozen water content in
the different regions and on the different dividing lines, the two
water phase change processes (from the non-frozen membrane
water to ice, and between the non-frozen membrane water and
vapour) can either occur or stop simultaneously, or one process
can also occur when the other stops, as illustrated in Table 2. The
principal driving forces for these water phase change processes in
the CL are the differences between the non-frozen water content
in the ionomer and the equilibrium water content, and between
the non-frozen water content and the saturation water content.
It should be mentioned that Equation (12) is derived based on
the experiential measurement at 30 C. As shown in Fig. 7
[72,81,82] comparing the equilibrium water contents at two
different temperatures (30 and 80 C), although the equilibrium
water content increases slightly with temperature increment at low
water vapour activity, it decreases with increasing temperature at
high water activity, especially for the hydration from saturated
water vapour. The drying out of the ionomer at elevated temperature suggests that the membrane needs to be better hydrated at
higher temperature. In addition, the measurements in [81,82] also
revealed that the equilibrium water content is higher when the
membrane is immersed to liquid water than that exposed to
saturated water vapour, and this is the reason that the water
activity is usually calculated to be greater than 1 when liquid water
is present for calculating the equilibrium water content (Equations
(12) and (13)). The details about the ionomer absorption and
desorption processes are given in Section 3.3.
3. Mechanism of water transport
With the different materials and local operating conditions in
the different components of PEMFC, the mechanisms of water
transport are different as well. In this section, the mechanisms of
water transport in the different cell components are described one
by one to provide a complete view of water transport in PEMFC.
Note that starting from this section, only the normal operating
conditions are considered, and the cold start processes are explicitly described in Section 11. Therefore, unless otherwise specified,
the water content mentioned starting from this section (except
Section 11) represents the non-frozen water content, since no
frozen water is present in ionomer in normal operating conditions.
3.1. In membrane
Effective membrane hydration is of paramount importance for
reducing the ohmic loss of PEMFC for all the PFSA polymer
Hydrophobic
polytetrafluoroethylene
(PTFE) backbone
F F
F F
C C
C C
F F n O Fm
m=1
n = 6 ~ 10
F C F
F C CF3
Polymer side chain
O
m
F C F
F C F
2.4. In flow channel
The state of water in flow channel is similar to as in GDL. As
shown in Figs. 3 and 4, for normal operating conditions, liquid
Sulphuric acid (SO3− H +)
O S O − H+
O
Fig. 8. Schematic of chemical structure of Nafion.
234
K. Jiao, X. Li / Progress in Energy and Combustion Science 37 (2011) 221e291
At a low water content:
At a high water content:
SO3−
H+
H2O
Polymer chain
1 nm
Fig. 9. Schematics of micro-structural features of Nafion for low and high water contents (note that the free volume containing water at a low water content is smaller than at a high
water content, and such difference is neglected in this figure).
membranes. As mentioned earlier, in an operating PEMFC, the
membrane close to the anode side may dry out quickly due to the
EOD effect, therefore balancing the various water transport
processes to evenly hydrate the membrane is critical, and this
requires the understanding of the various water transport mechanisms in membrane. In this subsection, the proton transport in
membrane is first elaborated (because it largely couples with water
transport), followed by the explanations of the various water
transport mechanisms in membrane (diffusion, EOD effect and
hydraulic permeation). In addition, the reactant transport in
membrane and membrane expansion are described as well.
3.1.1. Proton transport
The proton transport has significant influence on the water
transport in membrane; on the other hand, the proton transport
also largely depends on the membrane environment (water
SO3−
H+
content). Therefore, the proton transport is described first in this
subsection to guide the following discussions on the water transport in membrane. Generally, the ionomer (polymer electrolyte)
proton conductivity follows the Arrhenius Law [83]. In fact, Equation (10) for calculating the proton conductivity is essentially
a modified Arrhenius equation based on experimental measurements, and it shows that the proton conductivity strongly depends
on the water content and temperature. Fig. 8 presents the schematic of chemical structure of Nafion. It consists of PTFE backbone
providing the mechanical stability with polymer side chains with
sulphuric acid for promoting proton conduction. As mentioned
earlier, with large amount of SO3 , the proton conductivity is high,
but the mechanical and thermal strengths are low; and with water
absorbed in the membrane, the Hþ becomes weakly attracted to the
SO3 , resulting in easier proton transport. Fig. 9 shows the schematics of micro-structural features of Nafion for both the low and
Polymer chain
Fig. 10. Schematic of direct proton transport between polymer chains of Nafion.
K. Jiao, X. Li / Progress in Energy and Combustion Science 37 (2011) 221e291
high water contents. The solid lines in this figure represent the
polymer chains (the hydrophobic backbones and polymer side
chains shown in Fig. 8), where those directly connected to the SO3
are the polymer side chains (also shown in Fig. 8). Each polymer
side chain attracts one Hþ forming sulphuric acid. The Hþ is relatively weakly bound to the polymer side chains. There are free
volumes surrounded by the polymer chains, on the scale of nanometre, and water can be absorbed into these spaces (the hydrophilic regions discussed with Fig. 3 earlier). It should be mentioned
that the size of these spaces increases with the increment of water
content, and such difference between the spaces at the low and
high water contents is neglected in Fig. 9.
As shown in Fig. 9 at the low water content, the water in the
membrane is mostly likely bound to the slide chains. This kind of
water may be classified as the non-freezable water [62e64,69e71]
235
or surface water [73,74], as mentioned in Section 2.1. This part of
water is strongly bound to the charged sites (SO3 ). In this scenario,
proton has to transfer through the void volume from one charged
site to another. Since the Hþ is relatively weakly bound to the SO3 , it
is possible for it to jump from site to site directly. The mechanism of
proton transport from one charged site to another directly is
demonstrated in Fig. 10 (this is the most common type of proton
transport in solid conductors). It can be noticed that the polymer
side chain can actually vibrate in the free volume, and the movement of the polymer side chain can physically reduce the distance
for the Hþ transport. When one extra Hþ is present at the polymer
side chain from left hand side (the first picture from the left of
Fig. 10), the two polymer side chains may attract each other and
therefore vibrate. Again, it should be noticed that increasing the
SO3 enhances the proton transport by reducing the distance
Fig. 11. Schematics of proton transport in water (a: vehicle mechanism; b: hopping mechanism).
K. Jiao, X. Li / Progress in Energy and Combustion Science 37 (2011) 221e291
content and conductivity, it has been observed that the membrane
conductivity drops sharply when the water content is lower than 5
[65,66,95,96], and the membrane conductivity starts becoming
appreciable when the water content is higher than 2, these values
may also be used to explain the percolation threshold for the
hopping mechanism for proton transport.
If no water is present in a membrane, the proton can only transfer
directly between the charged sites, resulting in very low proton
conductivity. If water is present in a membrane but the amount is
below the percolation threshold, the proton transfer mechanisms will
be mainly the direct transfer between the charged sites and vehicle
mechanism (with the increment of water content, the water diffusion
coefficient is also higher, therefore the vehicle mechanism becomes
more dominating). When the amount of water is higher than the
percolation threshold, the dominating proton transport mechanisms
are the vehicle and hopping mechanisms. It has been reported that for
intermediate and low hydration levels, the vehicle mechanism still
dominates the proton transport, and for water content higher than 13,
the hopping mechanism dominates [97e100].
3.1.2. Diffusion of water
Due to the concentration gradient, water can diffuse through the
void space of membrane. Since water is produced in cathode CL,
resulting in more water on the cathode side, the diffusion of water
is therefore usually from cathode to anode. The diffusional nonfrozen water flux in membrane, Jnmw,diff (kmol m 2 s 1), due to the
concentration gradient is a vector quantity and can be written as
Jnmw;diff ¼
Dnmw Vcnmw ¼
rmem
EW
Dnmw Vlnf
(14)
20
-2
m s
-1
where Dnmw (m2 s 1) is the diffusion coefficient of non-frozen
water in membrane, cnmw (kmol m 3) the non-frozen water
concentration in membrane; and as mentioned earlier, rmem
(kg m 3) is the density of dry membrane, and EW (kg kmol 1) the
equivalent weight. The negative sign in Equation (14) presents the
fact that the diffusional flux is always in the direction of decreasing
the concentration. Note that since only the normal operating
condition is considered in this section, the non-frozen water
represents all the water in membrane.
The diffusion coefficient is sensitively dependent on the
membrane hydration. Experimentally, self-diffusion coefficient is
relatively easily measured by tracking the tracer in a homogeneously
hydrated membrane, due to the random molecular motion. Generally,
the water self-diffusion coefficient is similar in the PFSA polymer
-10
between the SO3 , but the mechanical and thermal strengths are
reduced. The vibration of the polymer side chains enhances the
proton transport, explaining why the polymer membranes exhibit
higher proton conductivity than most of the other solid proton
conductors (without vibration of the polymer side chains) (e.g.
ceramics).
Proton may also transfer by hitching a ride on water by forming
hydrogenewater ions (e.g. H3Oþ, H5Oþ
2 or something similar). The
diffusion of the hydrogenewater ions happens from high to low
proton concentration regions, therefore facilitating proton transport. This kind of proton transport is called the vehicle mechanism
(also called the vehicular diffusion) [84]. However, the diffusion of
the hydrogenewater ions may be retarded due to the hydrogen
bonding [85] (i.e. H3Oþ may bond with other water molecules and
therefore difficult to move). For a well hydrated membrane, more
water exists in the membrane, as shown in the second picture in
Fig. 9. If the amount of water is high enough to connect the polymer
side chains, the proton may also transfer directly from one water
molecule to another, and the water molecules in this case are
essentially the charged sites. This kind of proton transfer is called
the hopping mechanism (also called the Grotthuss mechanism or
structure diffusion) [86e90]. These two mechanisms are demonstrated in Fig. 11. The vehicle mechanism can be relatively easily
understood as the diffusion of the hydrogenewater ions, as shown
in Fig. 11a. The hopping mechanism can be explained with the help
of Fig. 11b. The transferring proton can reside on a water molecule
forming a hydronium ion (H3Oþ), and an Eigen-ion (H9Oþ
4 ) can be
formed since the H3Oþ is bound to three neighbouring water
molecules by forming hydrogen bonds, as shown in the top left
picture of Fig. 11b. The proton can then transfer to a location
symmetrically between two water molecules, and a Zundel-ion
(H5Oþ
2 ) can be formed, as shown in the top right and bottom
pictures of Fig. 11b. After that, another H9Oþ
4 can be formed but at
a different location than the previous H9Oþ
4 . Such transformations
þ
between the H9Oþ
4 and H5O2 are caused by the hydrogen bond
forming and breaking processes, which essentially results in proton
transport.
It should be noticed that the vehicle mechanism mainly depends
on the diffusion coefficient of water (or hydrogenewater ions) in
membrane, which depends on the water content and temperature
(explained later in Section 3.1.2). The hopping mechanism takes
place if enough amount of water is present in membrane to form
continuous network between the charged sites. This can be
explained by using the percolation theory, by which enough free
volume in a solid material has to be ensured for possible fluid flow
from one side to another. The minimum amount of volume fraction
required is called the percolation threshold [91,92]. The percolation
threshold for the hopping mechanism is not determined yet, even
though it is believed to be between the water contents of 1 and 7.
The water content in Nafion membrane can change from
0 (completely dry) to 22 [93], and for PEMFC operating at 80 C, the
maximum membrane water content is considered to be about 16.8
[72,93]. Assuming that all the free volumes in a membrane are
occupied by water at its maximum water content (22), based on the
percolation threshold for randomly distributed equiaxed ellipsoids
(aspect ratio is 1), where the percolation does not occur until the
volume fraction of the conducting phase is 28.54% [92], the
percolation threshold for the hopping mechanism in membrane is
therefore about 6.28 water content, which seems to be high. It was
also claimed that the percolation threshold is about 5% of the
maximum water content [94], which is only about a water content
of 1.1. Another approximation is based on the measured proton
conductivity, when the proton conductivity becomes appreciable,
the water content ranges from 2 to 5. Even though Equation (10)
shows a linear relationship between the membrane water
Fickian D iffus ion Coefficient, 10
236
15
10
Equation 17 [105]
5
Equation 16 [104]
0
0 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17
Water Content
Fig. 12. Fickian diffusion coefficients of water in Nafion membrane at different water
contents and at 80 C by using different correlations [104,105].
237
K. Jiao, X. Li / Progress in Energy and Combustion Science 37 (2011) 221e291
membranes in terms of the magnitude and the trend of variations
with the water content and temperature. Due to the small spaces for
water diffusion, and the hydrogen bond which retards the water
movement, the self-diffusion coefficient in Nafion is merely about
1
2416
>
303
>
>
>
: 10 10 exp 2416 1
303
> 10
10 exp
ðl 2Þ
i
1 h
0:87 3 lnf þ 2:95 lnf 2 ; ð2 < l 3Þ
T
i
1 h
2:95 4 lnf þ 1:642454 lnf 3 ; ð3 < l 4Þ
T
1
2
3
2:563 0:33lnf þ 0:0264lnf 0:000671lnf ; ðl > 4Þ
T
four times lower than the value in bulk liquid water when hydrated by
saturated vapour. The self-diffusion coefficient has been measured
experimentally at different temperatures and water contents [95,101]
using pulsed gradient NMR spectroscopy. The self-diffusion coefficient is measured when the membrane hydration is uniform and
homogeneous for the entire membrane often referred to as the
intradiffusion coefficient, which is applicable for the completely
hydrated membrane in PEMFC operations. In practice, the membrane
during the dynamic operation of PEMFC may be partially dried out on
the anode side and yet still maintain full hydration on the cathode
side. In the presence of such water gradient, the appropriate coefficient describing the water diffusion through such membrane is the
interdiffusion (or Fickian diffusion) coefficient, which is related to the
self-diffusion coefficient. For systems where the transport number of
electrons is zero or unity, the intro- and Fickian diffusion coefficients
(Dnmw,I and Dnmw,F) are related through the “Darken factor” [102]
Dnmw;F
"
#
vlnðaÞ
¼ Dnmw;I
vln lequil
|fflfflfflfflfflfflfflfflfflfflffl{zfflfflfflfflfflfflfflfflfflfflffl}
(15)
Darken factor
where a is the water activity when measuring the intradiffusion
coefficient, and lequil the equilibrium membrane water content. The
Darken factor can be obtained by taking the reciprocal of the
differential of Equation (12). Based on the experimental measurement in [101] at 30 C and the activation energy of water diffusion
in Nation membrane [103], and by transforming the intradiffusion
coefficient to Fickian diffusion coefficient, the following correlation
has been developed [72,104] to calculate the Fickian diffusion
coefficient of non-frozen water in membrane (Equation (16)),
where the units of Dnmw,F and Dnmw are all m2 s 1, and the unit of
T is K. Note that unless otherwise specified, the diffusion coefficient of water in membrane represents the Fickian diffusion
coefficient in this paper. Similarly, based on the same experimental data in [101] and similar methods, the following correlation has been developed [105].
Dnmw ¼ Dnmw;F
¼
(
h
i
1 exp 2346
3:1 10
;
nf exp 0:28lnf
T
0 < lnf < 3
h
i
ð17Þ
;
4:17 10 8 lnf 161exp lnf þ 1 exp 2346
T
3 lnf < 17
7l
where the units of Dnmw,F and Dnmw are all m2 s 1, and the unit of
T is K.
ð16Þ
used for PEMFC modeling. It can also be noticed that the two correlations all show that the maximum diffusion coefficient occurs at the
water content of 3, caused by the transformation (Equation (15)) of
the diffusion coefficient, and such sharp change of the diffusion
coefficient also increases the difficulty for numerical simulations.
3.1.3. Electro-osmotic drag effect
The EOD coefficient, nd, depends on the water content of
membrane. Water content varies across a Nafion membrane
because of several factors. Perhaps most important is the fact that
protons traveling through the pores of Nafion generally drag water
molecules along with them. Actually, protons travel in the form of
hydronium complexes (H3Oþ) or something similar as explained
earlier. For simplicity, however, it is straightforward to define the
EOD coefficient in terms of the number of water molecules per
proton. In other words, EOD coefficient is defined as the ratio of
mole-of-water per mole-of-proton transported through the
membrane in the absence of concentration and pressure gradient.
The water flux (Jnmw,EOD, kmol m 2 s 1) due to the EOD is
Jnmw;EOD ¼ nd
Iion
F
(18)
where Iion (A m 2) is the ionic current density, and F
(9.6487 107 C kmol 1) the Faraday’s constant.
For Nafion membranes, nd has been measured to be 2.5 at
a water content of 22, and 0.9 at a water content of 11 [81]. Another
measurement showed that it is 1.4 for the water contents from 5 to
2.5
Electro-os motic D rag Coefficient
Dnmw ¼ Dnmw;F ¼
8
2:692661843 10 10 ;
>
>
>
1
>
10
>
< 10 exp 2416 303
Fig. 12 compares the diffusion coefficients by using the two
correlations at 80 C. Apparently, the difference between the two
correlations is not negligible, and there is no conclusion indicating the
more accurate one. In fact, the two correlations have all been widely
2
1.5
1
Equation 20 [108]
0.5
Equation 19 [72]
0
0
2
4
6
8
10 12 14 16 18 20 22
Water Content
Fig. 13. EOD coefficients of Nafion membrane at different water contents by using
different correlations [72,108].
238
K. Jiao, X. Li / Progress in Energy and Combustion Science 37 (2011) 221e291
14, and gradually decreases to zero for the water contents from 5 to
0 [106]. It was also reported that it is 1 for the water contents from
1.4 to 14 [107]. nd has also been reported for several other PFSA
polymer membranes [108]. Correlations have been developed
based on the experimental measurements for numerical models,
one shows a linear relationship [72]:
nd ¼
2:5lnf
22
(19)
and another one shows a stepwise correlation [108]:
nd ¼
lnf 14
0:1875lnf 1:625 ðotherwiseÞ
1
DO2
(20)
Fig. 13 compares these two correlations and the difference is not
negligible, even though both the correlations have been widely
used for PEMFC modeling.
3.1.4. Hydraulic permeation
The water flux (Jnmw,hyd, kmol m 2 s 1) associated with the
hydraulic permeation of water due to pressure gradient is
Jnmw;hyd ¼
cnmw
Knmw
mnmw
Vpnwm ¼
lnf
rmem Knmw
Vp
EW mlq
(21)
where Knmw (m2) is the permeability of non-frozen water in ionomer, cnmw the non-frozen water concentration in ionomer
(kmol m 3), mnmw (kg m 1 s 1) the dynamic viscosity of non-frozen
water in ionomer (the property of liquid water is often used
instead), and pnmw (Pa) the pressure of non-frozen water in ionomer. It is in the direction of deceasing pressure, represented by the
negative sign. The permeability of non-frozen water in membrane
is mainly associated with water content because the pore size in
membrane increases with the increment of water content, as
described in Section 3.1. The following correlation has been
developed to calculate Knmw (m2) [109,110].
Knmw ¼ 2:86 10
20
lnf
preferred in order to minimize cell performance losses related to
mass transfer resistance (or the depletion of the reactants at the
reaction sites). This requires optimization of the ionomer fraction in
CL to ensure high proton conductivity in CL while maintaining
enough amount of reactant transport. Like water in the membrane,
the diffusion of reactant gas is much larger than hydraulic permeation, and the diffusion coefficient ðDO2 mem ; m2 s 1 Þ for oxygen
in fully hydrated Nafion 117 for oxygen is [111]
(22)
Apparently, the permeability is very low. In order to provide
additional means for the reduction of water in the cathode and
hydrate the membrane close to the anode, cells may be differentially
pressurized such that the oxidant gas is supplied at a higher pressure
than the fuel gas in the anode. Thus, the predominant direction of
the diffussional and hydraulic water fluxes can be arranged opposite
to that of EOD water flux, balancing the water in the membrane. In
addition, the mechanical strength of the membrane needs to be
considered when pressurizing the anode and cathode differentially,
as well as the parasitic power requirement.
3.1.5. Reactant transport
The requirement for the transport of reactant gases through the
membrane is self-conflicting: on one hand, low diffusion coefficients, hence low rates of the reactant transfer, through the
membrane are mandatory to separate the fuel and oxidant gas from
mixing in order to avoid the degradation of cell performance and
the occurrence of potential hazards; on the other hand, the electrochemical reaction for electric energy generation in the CLs is
heterogeneous, occurring at the surface of the catalyst which is
surrounded by the ionomer (polymer electrolyte) (as shown in
Fig. 3). The ionomer covering the catalyst surface is essential for the
protons to be transported away avoiding reaction product accumulation in the anode CL and to transfer the protons as reactants
for the cathode reactions. This creates a significant challenge for the
reactant gases (hydrogen and oxygen) to reach the catalyst surface
and high values of hydrogen and oxygen diffusion coefficients are
mem
and for hydrogen ðDH2
is [109]
DH2
mem
10
¼ 2:88 10
¼ 4:1 10
1
exp 2933
313
mem ;
7
exp
m2 s
1Þ
1
T
(23)
in fully hydrated Nafion 117
2602
T
(24)
where T is temperature with the unit of K. It can be noticed that the
diffusion coefficients of reactant gases in membrane are very low
(about 10 6 of the bulk diffusion coefficients), therefore the
transport of reactants through the membrane is usually neglected
in PEMFC studies.
It should be mentioned that the diffusion of oxygen and
hydrogen through the membrane occurs after the oxygen and
hydrogen gas have dissolved in the hydrated membrane, following
the Henry’s Lay [1]:
pi
Hi
ci ¼
(25)
where ci (kmol m 3) is the reactant concentration on the
membrane side (i represents hydrogen or oxygen), pi the partial
pressure (Pa) of the corresponding reactant on the gas side, and Hi
(Pa m3 kmol 1) the corresponding Henry’s constant. For oxygen
dissolving in hydrated Nafion 117, the Henry’s constant for oxygen
in membrane ðHO2 mem ; Pa m3 kmol 1 Þ is [109]
HO2
mem
¼ 101:325exp
666
þ 14:1
T
(26)
where T is the temperature with the unit of K. Equation (26) yields
about 2.04 107 Pa m3 kmol 1 at the typical PEMFC operating
temperature of 80 C. Hydrogen is a weak function of temperature
and can be considered as constant: 4.56 107 Pa m3 kmol 1 for the
operating temperature range of PEMFC. The Henry’s constants are
1.45 108 and 1.26 108 Pa m3 kmol 1 for hydrogen and oxygen in
liquid water, respectively, much higher than in hydrated membrane
[112,113]. Apparently, the reactants are more soluble in membrane
than in liquid water. In fact, the reactants are more soluble when
the membrane is dry, however, the diffusion of the reactant gases is
faster when the membrane is better hydrated [114], and again, the
slow diffusion of the reactants in membrane may be neglected.
3.1.6. Membrane expansion
As mentioned earlier, the pore size in PFSA polymer membranes
increases when water is absorbed, in this case the membranes also
expand in volume. As a result of the volume change, the concentration of the fixed charged sites also changes, depending on water
content. For Nafion membrane, the concentrations of charged sites
and water (ccs and cH2 O , kmol m 3) can be expressed as [72]
ccs ¼
r
1
1
; c
¼ lccs ¼ l mem
EW 1 þ 0:0126l H2 O
EW 1 þ 0:0126l
rmem
(27)
where rmem (kg m 3) is the density of dry membrane, EW
(kg kmol 1) the equivalent weight, and l the total water
content.
239
K. Jiao, X. Li / Progress in Energy and Combustion Science 37 (2011) 221e291
For this subsection (3.1), and as mentioned previously, only the
Nafion membrane is focused on, which is not the ideal membrane
but the best or standard so far. More information about the other
membranes such as Flemion, Gore and so on is also available in
literature [100,115].
3.2. In gas diffusion layer
As the physical support of membrane and CL, GDL is attached to
the outside of CL and the whole structure (membrane, CL and GDL)
is compressed to form the membrane electrode assembly (MEA).
Carbon paper or carbon cloth with thicknesses between 0.2 and
0.4 mm is widely used as the GDL because of the high porosity
(usually higher than 0.5) that facilitates the reactant transport
towards the catalyst sites. Also the very low electrical resistance of
carbon paper or carbon cloth makes the ohmic loss within an
acceptable range. PEMFC at high current density operation is of
particular interest to vehicular applications for obtaining high
power density, as long as with sufficient fuel and oxidant supply,
the “concentration or mass transport loss” shown in Fig. 2 comes
into play due to the excessive liquid water build up. As mentioned
earlier and shown in Fig. 3a, liquid water blocks the porous pathways in CL and GDL thus causing hindered oxygen transport to the
reaction sites, such water flooding phenomenon is perceived as
the chief mechanism leading to the limiting current behaviour in
the cell performance, and this is the reason that the carbon paper or
carbon cloth is usually treated to be hydrophobic by coating a layer
of PTFE on them to expel water. Therefore, effective water removal
in GDL without affecting the membrane hydration is critical, and
this requires understanding the water transport mechanism in GDL.
In this subsection, the various water transport mechanisms in GDL
(diffusion, convection and capillary effect) are explained, as well as
the water condensation and evaporation.
3.2.1. Diffusion and convection
In GDL and CL, the flow exhibits porous and tortuous structures
on the micro- and nanometre length scales, in which the convective
forces are resisted. As a result, the flow in GDL can be diffusion
dominated, convection dominated, or mixed, depending on the
design of the flow channel. Convection refers to the bulk motion of
a fluid (under action of a mechanical force), and diffusion refers to
the transport of a species due to a concentration gradient. In
PEMFC, the convective force that dominates the convective transport is the pressure at the flow channel inlets. High flow rate can
ensure good distribution of reactants (and effective water removal)
but may require unacceptable high driving pressures or lead to
other problems. The concentration gradients that dominate diffusive transport are from species consumption/production in CL: the
reactant consumption and water product result in reactant delivery
and water removal.
Fig. 14 shows the flow characteristics in GDL with different flow
channel designs. For parallel flow channel design, only small inlet
pressures are needed for reactant flow because the flow distance
from flow channel inlet to outlet is short, and the pressures are
relatively evenly distributed for each straight channel. Since very
small pressure gradient is present in GDL, resulting in slow
convective flow, the flow is mostly likely diffusion dominated in
GDL. Serpentine flow channel design needs higher inlet pressures
than parallel design because the distance from inlet to outlet is
longer, resulting in larger pressure gradient and cross flow from
channel to channel through GDL directly, therefore the flow can be
dominated by both diffusion and convection. Interdigitated design
forces all the flow through GDL, resulting in the largest pressure
gradients among the three flow channel designs shown in Fig. 14
and the flow becomes convection dominated in GDL.
In the porous and tortuous flow structures of GDL and CL, the
movement of the gas molecules can be restricted by the pore walls,
lowering the diffusional flux. To account for such diffusion resistance, a modified or effective diffusion coefficient based on the
porosity and tortuosity can be used [116].
Deff
¼ Di
i
3
s
(28)
where Deff
i and Di represent the effective and bulk diffusion coefficients of gas species i, respectively; and 3 and s are the porosity
and tortuosity of GDL or CL, respectively. The porosity is defined as
the percentage of void volume in the total volume (void and solid
volumes), and the tortuosity describes the additional impedance to
Parallel design
Flow field
GDL
Diffusion
dominated
flow
Serpentine design
Flow field
GDL
Interdigitated design
Flow field
GDL
Fig. 14. Flow characteristics in GDL with different flow channel designs.
Convection
dominated
flow
240
K. Jiao, X. Li / Progress in Energy and Combustion Science 37 (2011) 221e291
diffusion caused by a tortuous or convoluted flow path. The Bruggeman correlation provides the relationship between porosity and
tortuosity as [117]
s¼ 3
0:5
(29)
therefore Equation (28) can be simplified as
Deff
¼ Di 31:5
i
(30)
Other correlations have also been developed [118e122] to
calculate the effective diffusion coefficient in porous medium, and
there is no significant difference among the different correlations.
Therefore the simplicity of the Bruggeman correlation made it
become the most popular one. Considering the blockage of liquid
water, it can be assumed that liquid water has the similar effect as the
solid materials [62,63], therefore Equation (30) can be modified as
Deff
¼ Di 31:5 1
i
slq
1:5
(31)
Different values of the exponential for the liquid water volume
fraction (slq) term in Equation (31) have also been reported, e.g.
a value of 2 was suggested in [121].
The convective flow through GDL (and CL) by pressure gradient
is mainly affected by the permeability of the GDL (and CL). The
pressure gradient caused due to the porous and tortuous structure
can be calculated for gas and liquid phases:
Vpg;lq ¼
mg;lq !
Kg;lq
u g;lq
Klq ¼ K0 s4:0
lq
slq
4:0
Dp ¼ s
1
1
þ
R1 R2
(35)
where the surface curvature is represented by the two radii (R1 and
R2, m) of the curved surface (interface) in two orthogonal directions.
Considering liquid water transport in PEMFC cathode, the
following non-dimensional numbers can be used to evaluate the
importance of surface tension effect.
Reynolds number :
Re ¼
rlq Ulq L
mlq
(36)
Capillary number :
Ca ¼
mlq Ulq
slq air
(37)
(32)
!
where p (Pa), m (kg m 1 s 1), K (m2) and u (m s 1) are the pressure,
dynamic viscosity, permeability and velocity, respectively; and the
subscripts g and lq represent gas and liquid phases, respectively. Note
that Equation (32) is only valid when the gravitational force can be
neglected. For large velocity flow (e.g. with serpentine and interdigitated flow channel), the right hand side of Equation (32) can be added
as an extra resistance force to standard momentum conservation
equations for the porous media (detailed in Section 6). The intrinsic
permeability of GDL is typically on the level of 10 12 m2, and the gas
phase and liquid phase permeabilities (Kg and Klq, respectively, m2)
depend on the intrinsic permeability of the porous materials (K0, m2)
and the local volume fraction of liquid water. The gas phase and liquid
phase permeabilities can be calculated as [62,63]
Kg ¼ K0 1
the force acting to minimize the free energy (minimize the forces at
the interface) by decreasing the area of the interface, resulting in the
droplet squeezing itself together until it has the locally lowest surface
area possible. Surface tension can be represented by a fluid property,
surface tension coefficient (s, N m 1), representing the tensional force
along a line on the interface (therefore having of unit of N m 1). If
there is no pressure difference across the interface, the interface
remains flat. If the pressure on one side is greater than the other side,
the interface is curved to the low pressure side (as for the droplet and
air example mentioned before, the shape of water droplet indicates
that the interface is curved to the air side). The surface tension coefficient (s, N m 1), pressure difference (Dp, Pa) and surface curvature
can be related by the YoungeLaplace equation [126]:
(33)
Weber number :
Bond number :
Bo ¼
2L
rlq Ulq
¼ Re$Ca
slq air
(38)
rair gL2
rlq
slq
(39)
air
where rlq, Ulq, L, mlq, slq-air, rair and g are the liquid water density, liquid
water velocity, characteristic length, liquid water dynamic viscosity,
liquid water surface tension coefficient when exposed to air, air
density and gravity, respectively; and by using the typical values (at
80 C and 1 atm) for these parameters of 990 kg m 3, 10 5 m s 1,
8 10 5 m (the pore diameter in GDL), 3.5 10 4 kg m 1 s 1,
0.063 N m 1, 1 kg m 3 and 9.81 m s 2, respectively, the Reynolds
number (inertia force divided by viscous force) is calculated to be
0.0023, the Capillary number (viscous force divided by surface
(34)
where slq represents the volume fraction of liquid water in the
pores. Other values of the exponentials in Equations (33) and (34)
have also been reported in the range from 3.0 to 5.0 [80,123e125].
3.2.2. Surface tension and wall adhesion effects in porous media:
capillary effect
Surface tension is a force, acting only at the interface between
liquid and liquid, liquid and gas or liquid and vacuum. For example, for
a liquid water droplet in air, the inter-molecular forces on the water
molecules inside the droplet are balanced, however, at the interface
between the droplet and air, the inter-molecular forces (attractive
forces) on the droplet side are larger than on the air side. The water
molecules at the interface are therefore subject to an inward force of
inter-molecular attraction which is only balanced by the liquid
water’s resistance to compression. This results in a pressure difference
across the interface (the pressure in this case is higher on the droplet
side due to the larger inward inter-molecular force and the resistance
to compression). In this case, the surface tension can be understood as
We ¼
d
d
Gas
Gas
R
θ
θ
R
θ
R
Liquid water
Liquid water
Hydrophilic (θ < 90o)
pg > plq
Hydrophobic (θ > 90o)
pg < plq
Fig. 15. Two-phase behaviours in small pores with different surface wettabilities.
K. Jiao, X. Li / Progress in Energy and Combustion Science 37 (2011) 221e291
tension) is calculated to be 5.6 10 8, the Weber number (inertia
force divided by surface tension) is calculated to be 1.3 10 10, and
the Bond number (gravitational force divided by surface tension) is
calculated to be 9.9 10 4. Apparently, the surface tension effect
plays an important role in liquid water transport in PEMFC, using the
typical values for liquid water in CL results in the same conclusion.
Note that the liquid water velocity might become much higher when
strong convective flow occurs in GDL (as mentioned in Section 3.2.1).
Assuming that the interface is identical to a sphere surface, the
R1 and R2 in Equation (35) become identical to the radius of the
sphere radius (R, m), and Equation (35) can be simplified to
2s
Dp ¼
R
(40)
In GDL (and CL), liquid water transport is strongly affected by
the pore walls due to the small pore sizes, and the wall adhesion
effect becomes significant in this scenario. The surface wettability
of GDL (and CL) therefore plays a significant role in liquid water
transport. Fig. 15 shows the two-phase behaviours in small pores
with different surface wettabilities (hydrophobic and hydrophilic)
of the pore walls. It can be noticed that the angles between liquid
water and pore walls are different with different surface wettabilities. The angle q in Fig. 15 represents such angles, and is usually
called the contact angle. The contact angle can be considered as
a measure of the surface wettability. In the case in Fig. 15, for q less
than 90 , the surface is hydrophilic, and for q greater than 90 , it is
hydrophobic. A hydrophobic surface is simply more effective in
facilitating liquid transport than a hydrophilic surface due to the
reduced contact area between liquid and wall. This is why that the
GDL and CL are usually treated to be hydrophobic by adding PTFE to
them to expel water. With the different surface wettabilities of the
pore walls, it can be noticed in Fig. 15 that the interfaces between
liquid water and gas are significantly different, and the shape of the
interface is determined by the surface wettability. The hydrophilic
pore walls result in an interface that is curved to the liquid water
side (the pressure on the gas side is therefore higher than on the
liquid water side), and vice versa. The pressure difference can drive
the flow of liquid water and gas in the small pores, and such
movement in small pores is defined as the capillary motion, or
simply say that such movement is caused by the capillary effect.
The capillary effect is essentially the combined effects of surface
Capillary Pressure, kPa
0
Gas diffusion layer
-5
-10
Catalyst layer
-15
-20
-25
0
0.2
0.4
0.6
0.8
1
241
tension and wall adhesion in small pores. As illustrated in Fig. 15,
the sphere radius R (m) at the interface can be calculated by using
the pore diameter d (m) and the contact angle q as
R ¼
d
2cosq
(41)
The pressure difference across the interface in this case is also
called the capillary pressure (pc, Pa), which can be defined as
pc ¼ pg
plq ¼
4slq cosq
d
(42)
where pg (Pa) and plq (Pa) are gas and liquid water pressures, and slq
(N m 1) the surface tension coefficient of liquid water (exposed to
air or oxygen in cathode, and exposed to hydrogen in anode).
Apparently, the capillary pressure (pc, Pa) is an important
parameter affecting liquid water transport in PEMFC as a function
of liquid water surface tension coefficient (slq, N m 1), contact
angle (q), porous structure (represented by the porosity 3 and the
intrinsic permeability K0, m2), and liquid water volume fraction
(slq), and these parameters can be related based on the Leverett
function [127,128]:
(
0:5 h
2
1:42 1 slq
2:12 1 slq
3 i
if q < 90
þ1:26 1 slq
0:5 h
i
slq cosq K30
1:42slq 2:12s2lq þ1:26s3lq if q > 90
slq cosq
pc ¼
3
K0
(43)
Equation (42) shows that the liquid water pressure can be
calculated by using the gas phase and capillary pressures. The liquid
!
water velocity ( u lq , m s 1) by neglecting the gravity effect can be
calculated based on liquid water pressure by using Equation (32).
Fig. 16 shows the changes of capillary pressures with liquid
water volume fractions in GDL and CL by using Equation (43). For
the calculations in Fig. 16, the surface tension coefficient of liquid
water is 0.063 N m 1 (when expose to air); the contact angles in
GDL and CL are all 110 ; the porosities of GDL and CL are 0.6 and 0.3,
respectively; and the intrinsic permeabilities of GDL and CL are
10 12 and 10 13 m2, respectively. It can be noticed that negative
capillary pressures (liquid water pressure is higher than gas pressure) are obtained because the GDL and CL are all hydrophobic.
Note that Equation (43) was originally derived based on experimental data of homogeneous soil or a sand bend with uniform
wettability, which are different from the GDL and CL structures in
PEMFC. Other experimental measurements have been carried out
recently trying to assess the real situation in PEMFC, and the other
correlations for calculating capillary pressure are available in
[80,123,129e133]. However, due to the differences in the
measurement approaches, facilities, experimental conditions, and
the materials being investigated, the results do not agree with each
other very well. Therefore Equation (43) is still widely used for
PEMFC studies. In addition, the diffusion coefficient of liquid water
(Dlq, m2 s 1) and the relationship between liquid water and gas
!
!
velocities ( u lq and u g , m s 1) have also been derived based on the
capillary pressures [134] in GDL and CL, which concentrate significantly for modeling water transport in PEMFC:
Dlq ¼
Kg dpc
mlq dslq
(44)
Liquid Water Volume Fraction
Fig. 16. Capillary pressures in GDL and CL at different liquid water volume fractions
calculated by using Equation (43) (the surface tension coefficient of liquid water is
0.063 N m 1 (when expose to air); the contact angles in GDL and CL are all 110 ; the
porosities of GDL and CL are 0.6 and 0.3, respectively; and the intrinsic permeabilities
of GDL and CL are 10 12 and 10 13 m2, respectively).
Klq mg !
!
!
u lq ¼ i u g ¼
ug
Kg mlq
(45)
where i is named as the interfacial drag coefficient, representing
the ratio of liquid and gas velocities. i has been assumed to be 1 (gas
242
K. Jiao, X. Li / Progress in Energy and Combustion Science 37 (2011) 221e291
velocity equal to liquid velocity) to solve liquid water transport in
flow channel by assuming that liquid water only forms very small
droplets, however, the accuracy of this assumption remains
debated [80].
3.2.3. Condensation and evaporation of water
The condensation and evaporation of water in PEMFC depend
on the local mass and heat transfer conditions. From kinetic theory
[135,136], assuming an ideal gas and neglecting interactions
between individual molecules, the net mass transfer of the evaporation and condensation can be estimated using the Hertze
KnudseneLangmuir equation, as demonstrated in [137]:
Sv
l
0
rffiffiffiffiffiffiffiffiffiffiffiffi
MH2 O B
pvp
¼ Avp=lq
@2cond pffiffiffiffiffiffiffi
2pR
Tvp
p
1
lq C
2evap qffiffiffiffiffiffi
A
Tlq
(46)
where Sv-l (kg m 3 s 1) is the mass transfer rate of phase change
between vapour and liquid water, Avp/lq (m 1) the liquid/vapour
specific interfacial area (interfacial area per unit volume) which
depends on the volume fraction of liquid water, MH2 O
(18 kg kmol 1)
the
molecular
weight
of
water,
R
(8314 J kmol 1 K 1) the universal gas constant, 2cond and 2evap the
condensation and evaporation rate coefficients, pvp and plq the
vapour and liquid water pressures (Pa), and Tvp and Tlq the vapour
and liquid water temperatures (K). A comprehensive investigation
of the condensation and evaporation process is rather complicated
and needs to be performed in the surrounding regions of the liquid/
vapour interface on the molecular level. For the sake of simplicity,
for PEMFC modeling on the macroscopic level, it is impractical to
incorporate such processes and a revised form of the equation can
be used [134,137,138]:
Sv
l
Avp=lq
¼
RT
2ce
pvp
psat
(47)
where 2ce is the analogous condensation/evaporation rate and it
reads
2ce ¼ Gce
sffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi
RT
2pMH2 O
(48)
where Gce is an uptake coefficient that accounts for the combined
effects of heat and mass transport limitations in the vicinity of the
liquid/vapour interface. From the analysis of [121], this coefficient is
about 0.006. The specific liquid/vapour interfacial area is calculated
as
(49)
Avp=lq ¼ Gsurf Apore
where Apore is the pore surface area per unit volume which varies
from 1.3 107 to 3 107 m 1 for different GDL materials [130]. Gsurf
is an accommodation coefficient similar to Gce. The study of [139]
showed that Gsurf rarely exceeds 20% for spherical droplets with
small amount of liquid water. The ranges of Gce and Gsurf were also
estimated to be 0.001e0.006 and 1e20% for PEMFC operations
[137].
The water condensation/evaporation dynamics are limited by
the mass transport in the vicinity of the vapour/liquid interface, and
therefore Equation (47) was also modified as [137]
Sv
l
Shce Dvp
¼ Apore
d
pvp
psat
RT
(50)
where d (m) is the characteristic length for water diffusion, Dvp
(m2 s 1) the mass diffusivity of water vapour, and Shce the dimensionless number accounting for mass transport capability during
condensation/evaporation, and it is analogous to the Sherwood
number for mass transfer, calculated as [137]
Shce ¼ Gce Gsurf
sffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi
RT
d
2pMH2 O Dvp
(51)
The values of d and Dvp are not important here since they will
cancel out in Equation (50). d can be calculated based on the
diameter of the solid material of the porous media (e.g. the fibre
diameter of the GDL), dsl (m) based on the model in [121]:
d ¼ 4dsl
(52)
It has also been estimated that the range of Shce is from 0.00204
to 0.245 [137].
To differentiate the condensation and evaporation processes,
a Langmuir type correction with the porosity and local liquid
should be considered [137]:
Sv
l
¼
8
<
pvp psat
Shcond Dvp
3 1 slq
Apore
RT
d
if pvp > psat ðcondensationÞ
pvp psat
Shevap Dvp
3slq
Apore
RT
d
if pvp < psat ðevaporationÞ
ð53Þ
:
Shcond and Shevap are the phase transfer rates of condensation and
evaporation ranging from 0.00204 to 0.245 [137]. 3 and slq are the
porosity and liquid water volume fraction, respectively. Equation
(53) can be further simplified by using constant overall phase
change rates gcond and gevap (s 1):
Sv
l
¼
(
gcond 3 1
gevap 3slq
slq
ðpvp
ðpvp
psat Þ
RT
psat Þ
RT
if pvp > psat
if pvp < psat
ðcondensationÞ
ðevaporationÞ
(54)
Equation (54) is the most widely used equation to calculate the water
condensation and evaporation rates in PEMFC modeling on the
macroscopic level. However, the values of gcond and gevap from 1:0 to
104 s 1 have all been used in the previous studies [80,140]. Since the
water phase change rates are strongly affected by the local conditions
such as mass and heat transfer, the accuracy of the mass transfer rate of
phase change calculations on the macroscopic level remains debated.
3.3. In catalyst layer
On both sides of membrane, CLs usually form in terms of carbon
supported platinum powders as the catalyst embedded in part of
the membrane ionomer, as shown in Fig. 3. With the presence of
both the pore regions and ionomer, the most complex water
transport occurs in CL. Since the water transport in ionomer and
pores has been described in Sections 3.1 and 3.2, respectively, they
are not repeated in this subsection. It should be noticed that the
pores in CL are much smaller than in GDL, resulting in lower
porosity and permeability. The platinum particles in CL are typically
in the range from 1.5 to 2.5 nm, while the carbon support particles
are in the size range from a few mm to about 20 mm [1]. With
different platinum/carbon and ionomer fractions, the pore diameters range from several nm to about 1 mm [141]. The Knudsen
numbers (Kng and Knlq) for both the gas and liquid phases can be
calculated by the following equations [142].
mg
lg
Kng ¼
¼
d
pg
sffiffiffiffiffiffiffiffiffiffi
pRT 1
2Mg d
(55)
243
K. Jiao, X. Li / Progress in Energy and Combustion Science 37 (2011) 221e291
Knlq ¼
dlattice
lq
(56)
d
where lg (m) is the mean free path (average distance between collisions) of gas phase, d (m) the pore diameter, mg (kg m 1 s 1) the
dynamics viscosity of gas phase, pg (Pa) the pressure of gas phase, R
(8314 J kmol 1 K 1) the universal gas constant, T (K) the temperature,
the
Mg (kg kmol 1) the molecular weight of gas phase, and dlattice
lq
average distance between liquid water lattices. By using a pore
diameter of 10 nm for CL, the properties of air (the dynamic viscosity
is 2.08 10 5 kg m 1 s 1 and the molecular weight is 29 kg kmol 1),
the operating pressure and temperature of 101,325 Pa and 353.15 K,
and the average distance of 0.3 nm between the liquid lattices, the
Knudsen numbers for gas and liquid water are calculated to be about
8.2 and 0.03, respectively. Note that the macroscopic top-down
approach (by solving the continuity, NaviereStokes and other equations) is only applicable when the Knudsen number is lower than 10 1
with slip boundary conditions on walls and lower than 10 3 without
slip boundary conditions on walls, suggesting that the macroscopic
top-down approach cannot be used with the real micro-structure of
CL (also cannot be used with the nanometre pores in membrane).
The main mechanism of gas diffusion is essentially the collision
between gas molecules, as occurring in GDL, CL and flow channel.
However, in the extremely small pores in CL, as analyzed before by
calculating the Knudsen numbers, another mechanism of gas
diffusion occurs in CL as well, which is called the Knudsen diffusion
due to the collision between gas molecules and walls. The diffusion
coefficients related to the mechanism of collision between gas
molecules (the binary diffusion coefficient, DBi , m2 s 1) and related
to the collision between gas molecules and walls (the Knudsen
2
1
diffusion coefficient, DK
i , m s ) are [143,144]
DBi ¼ DB;ref
i
DK
i ¼
T
Tref
!1:5
pref
p
(57)
1 8RT 0:5
d
3 pMi
(58)
where DB,ref
(m2 s 1) is the reference binary diffusion coefficient at
i
the reference temperature (Tref, K) and pressure (pref, Pa); p (Pa) and
T (K) are the local pressure and temperature ; R (8314 J kmol 1 K 1),
Mi (kg kmol 1) and d (m) are the universal gas constant, molecular
weight of gas species i, and the pore diameter; and the subscript i
represents different gas species. The binary diffusion occurs in GDL,
CL and flow channel, and the Knudsen diffusion is only significant
in CL, the combined diffusion coefficient (Di, m2 s 1) can be
summarized as
Surrounded by water vapour
F
F
F
F
F F
8 B
< Di ðin GDL and flow channelÞ
1
Di ¼
: 1B þ 1K
ðin CLÞ
D
D
i
As mentioned earlier, the effective diffusion coefficient needs to
be further calculated in considering the porous and tortuous flow
structures of GDL and CL as well as the liquid water formation
(Equations (28)e(31)).
The water transfer between the ionomer and pore regions in CL
(membrane water absorption/desorption) plays an important role
in PEMFC because it determines the hydration/dehydration of
membrane. It has been observed that the amounts of water
absorption from liquid water and from saturated vapour are not the
same (from liquid water is higher than from saturated vapour).
Such phenomenon was initially reported by Schroeder in 1903,
hence the phenomenon has been called Schroeder’s paradox [145].
Evidence suggests that the PFSA polymer membrane surface is
strongly hydrophobic when it is in contact with water vapour
(whether saturated or not), and it becomes hydrophilic when in
contact with liquid water, as illustrated in Fig. 17. When a liquid
water droplet advances on the ionomer surface, wetting more areas
of the surface, the hydrophilic sulphuric acid moieties initially
inside the ionomer (when in contact with the water vapour on the
surface) spring out towards the liquid water now spread on the
surface, thus making the surface more hydrophilic and more water
absorption occurs when liquid water wets the surface. On the other
hand, water absorption from the vapour phase involves water
condensation on the hydrophobic surface, leading to less water
uptake. As mentioned in Section 2.3, this is the reason that the
water activity is usually calculated to be greater than 1 when liquid
water is present for calculating the equilibrium water content
(Equations (12) and (13)).
It should be mentioned that the humidification of the ionomer is
a very slow process, especially with water vapour. It has been
shown that the time scale for the membrane to reach its absorption
equilibrium state in humid air is on the order of 100e1000 s
[146,147] or even longer [148]. The following equation has been
used to calculate the mass transfer rate of the phase change (water
transfer) between the non-frozen water in ionomer and vapour/
liquid in pores (the membrane absorption/desorption rate, Sn-v,n-l,
kmol m 3 s 1) [80,149e151]:
Sn
v;n l
¼ gn
Surrounded by liquid water
− +
SO3− H + SO3 H
SO3− H + SO3− H +
− +
SO3 H
SO3− H +
rmem
EW
lnf
lequil
During operation
(in case surrounded by both
vapour and liquid)
F
F
F SO3− H + SO − H +
3
SO3− H +
SO3− H +
−
Ionomer surface
v;n l
(60)
where gn-v,n-l (s 1) represents the various phase change (water
transfer) rates; rmem (kg m 3) is the density of dry membrane, and
SO3− H + SO − H +
3
SO3 H +
− +
SO3− H +
SO3 H
SO3− H +
(59)
i
SO3− H +
SO3− H +
Ionomer surface
Ionomer surface
Fig. 17. Illustration of PFSA membrane surface morphology when it is in contact with vapour and liquid water [1].
K. Jiao, X. Li / Progress in Energy and Combustion Science 37 (2011) 221e291
3.4. In flow channel
Flow channel provides pathways for distributing reactant and
removing product water in PEMFC. The flow in flow channel is
convection dominated, and the driving force is the pressure at the
flow channel inlets. Considering a flow velocity of 5 m s 1 in
cathode flow channel with a typical cross section of
0.001 m 0.001 m (the characteristic length is therefore 0.001 m),
and using the values of 1 kg m 3 and 2.075 10 5 kg m 1 s 1 as the
density and dynamic viscosity of air, respectively, the Reynolds
number in flow channel is calculated to be about 241 by using
Equation (36), indicating that the flow in flow channel is mostly
likely laminar. By further considering the values of 990 kg m 3,
3.5 10 4 kg m 1 s 1 and 0.063 N m 1 as the density, dynamics
viscosity and surface tension coefficient (when expose to air) of
liquid water, respectively, it can be calculated from Equations (37)
to (39) that the inertia force, viscous force, surface tension (as
well as wall adhesion) and gravitational force all need to be
accounted for in flow channel. In fact, to facilitate liquid water
removal in flow channel, the channel walls can be modified to be
hydrophobic to minimize the wall adhesion effect. Since the various
transport mechanisms as well as the flow characteristics related to
different flow channel designs have been described in Sections
3.1e3.4, they are not repeated here.
3.5. Summary
The mechanisms of water transport in different components of
PEMFC are elaborated in Section 3. The transport of water in the
ionomer of membrane and CL involves diffusion, EOD and hydraulic
permeation, and it is largely coupled with proton transport. In the
pores of GDL and CL and in flow channel, the transport of water
vapour involves diffusion and convection. The capillary force also
plays an important role in liquid water transport in the pores of GDL
and CL. The Knudsen diffusion can be neglected in the pores of GDL
and in flow channel, but needs to be considered in the much
smaller pores of CL. Phase change processes take place until the
equilibrium states are achieved. However, as mentioned in Section
2.5, whether phase equilibriums of water exist remain debated
because of the presence and arbitrary transport of liquid water,
especially in the heterogeneous structures of CL and GDL.
4. Experimental observation
Presently available experimental techniques are excellent tools
for investigating the transport phenomena in PEMFC. The current
and high frequency resistance (HFR) distribution measurements
provide important information about the reactant delivery, product
removal and water distribution in membrane. The distributions of
the different gas species can be obtained from the species
concentration measurements at different locations of PEMFC. The
temperature distribution measurements and various water visualizations provide valuable information to guide better thermal and
water management of PEMFC. Not only help understand the
transport phenomena, the experimental observations also provide
valuable data to guide more complex and accurate numerical
modeling of PEMFC. This section reviews the various experimental
work published in literatures, the representing results obtained
from the experimental measurements are shown as well.
4.1. Current distribution measurement
The printed circuit board technique demonstrated in [152,153]
was first used by Cleghorn et al. [154] to the measure the current
distribution in a PEMFC. In the measurement of [154], a segmented
current collector on the anode side with different flow fields
separated was used, the anode GDL and CL were also segmented
corresponding to the current collector. This approach is called the
partial electrode approach with flow field, GDL and CL all
segmented, it allows for mapping of the current distribution on the
electrode surface to investigate the reaction kinetics at different
locations directly. Stumper et al. [155] demonstrated three methods
for current distribution measurement of PEMFC, including the
partial electrode approach as described in [154], the subcell
approach and the current distribution mapping approach. First, the
partial electrode approach involves the segmentation of flow field,
GDL and CL, therefore determining the local current density
behaviour of the electrode. Second, the subcell technique involves
placing small subcells at specific locations in a main cell and
isolating them, therefore the performance of the desired location
(at a very small scale) can be measured from the subcells. The third
is the current distribution mapping approach, in this approach the
current distribution is measured from the flow field plate with
unmodified MEA, e.g. shunt resistors normal to an unmodified MEA
surface were located between the flow field plate and a buss plate,
voltage sensors could passively determine the potential drop across
each resistor, and via Ohm’s law, current distribution through the
flow plate was determined [155]. All the current distribution
measurements can be generally categorized into these three
approaches, even though the different measurement techniques
were used for the same approach. The first and third approaches
1
-2
EW (kg kmol 1) the equivalent weight; and lnf and lequil are the nonfrozen membrane water content and equilibrium water content
(Equation (12)), respectively. In Equation (60), the mass transfer rate
is assumed to be proportional to the difference between the local
ionomer water content and the equilibrium value. Generally, the
study of water absorption/desorption is still relatively new and
many characteristics for this process remain unclear. A constant
value of 1.3 s 1 was suggested for gn-v,n-l [80,149].
The study in [147] showed that the physical mechanism of
membrane absorption is different from that of desorption which is
mainly limited by the interfacial mass transport. Water absorption
process presents a two-step behaviour: the initial 35% of water
absorption is described by the same interfacial transport rate
coefficient as that of desorption, while for the value above 35%,
water absorption is controlled by the dynamics of membrane
swelling and relaxation. It is found that the absorption process is 10
times slower than that of desorption in the second stage.
Current D ensity, A cm
244
Cell voltage = 0.8 V
Cell voltage = 0.65 V
Cell voltage = 0.5 V
0.8
0.6
0.4
0.2
0
0
0.2
0.4
0.6
0.8
1
Fractional Distance from Cathode Inlet
Fig. 18. Measured current density along fractional distance from cathode flow channel
inlet at different cell voltages [168].
245
K. Jiao, X. Li / Progress in Energy and Combustion Science 37 (2011) 221e291
were all widely adopted, and the second one is relatively less
popular, due to the fact that the manufacture of the special MEA
and flow field plate is complex and great care is needed to ensure
proper alignment [155].
The partial electrode approach has been continuously used and
improved [156e160], mainly by increasing the number of the
segments to improve the resolution. Natarajan and Nguyen
[159,160] segmented the CLs, GDLs and flow fields for both the
cathode and anode, and therefore the cell voltage/current density
for each segment can be measured and controlled separately. As
mentioned earlier, the partial electrode approach generally features
the ease of use and the current behaviour on the electrode surface
(reaction area) can be directly measured. However, the spatial
resolution is also limited by the size of the segmented electrode,
and it is also important to utilize a non-segmented electrode to
preserve the true operating characteristics. The current distribution
mapping approach using unmodified MEA has been widely adopted
due to such reasons [155].
Rather than using shunt resistors for the current distribution
mapping approach [155], Wieser et al. [161] developed a technique
utilizing a magnetic loop array embedded in a current collector plate
for the current mapping with non-segmented electrode. Similar to
[155], but utilizing small current collector pins connected to the
different locations on a flow field plate and to high-resolution
resistors for current measurements, the current mapping distribution approach was also used by Noponen et al. [162,163], and similar
measurements were also done by Brett et al. [164,165] and Mench
and Wang [166,167]. Even though the current mapping distribution
approach allows using unmodified MEA, the flow field plate still
needs modification, which may still result in unrealistic operating
characteristics. Developing new experimental approach and allowing accurate current distribution measurement without affecting
cell performance therefore becomes the greatest target for future
work.
Fig. 18 shows the measured current density along the fractional
distance from cathode flow channel inlet at different cell voltages
[168] by using the experimental methods developed in [166,167].
For measuring the current density shown in Fig. 18, the cell was
operating at 80 C and the inlet relative humidities are 100% and
50% for anode and cathode corresponding to the operating
temperature, respectively; and the stoichiometry ratios are 2 and
1.5 for supplying air and hydrogen corresponding to the operating
current density, respectively. It can be noticed that when the cell is
operating at a high voltage (0.8 V), the variation of current density
is insignificant; and for the intermediate and low cell voltages (0.65
and 0.5 V), the current density is higher and the variations are more
significant. The variation of current density is caused by the
combined effects of reactant consumption, membrane hydration/
dehydration and water production. Therefore investigating the
membrane hydration level and water concentration by measuring
the HFR and species concentration distribution is needed to interpret the measured current density distribution.
4.2. High frequency resistance distribution measurement
By using the experimental approaches for measuring current
distribution described in Section 4.1, HFR distribution can be
measured simultaneously [154,157,162,163,165,168], mainly by
using electrochemical impedance spectroscopy (EIS). The
membrane resistance contributes most significantly to the total
ohmic resistance, and the other resistances do not vary much
during operation. Therefore, the HFR distribution measurement can
be used to estimate the membrane hydration level at different
locations and to explain the measured current distribution data.
Fig. 19 shows the measured HFR along the fractional distance
from cathode flow channel inlet at different cell voltages [168]. In
fact, the HFR distribution shown in this figure was measured
simultaneously with the current distribution measurement in
Fig. 18. The two figures together show that at most of the locations
with low HFRs (high membrane hydration level), the corresponding
current densities are high, and vice versa. Fig. 19 also shows that the
locations with the lowest HFRs (highest membrane hydration level)
are at the inlet and outlet, due to the humidification of the supplied
reactants (hydrating the membrane most significantly at inlet) and
the accumulation of product water (hydrating the membrane most
significantly at outlet). In addition, the corresponding HFR increment to the current density drop at the outlet in Fig. 18 cannot be
observed in Fig. 19, suggesting that the current density drop at the
outlet is perhaps due to the concentration or mass transport loss.
Therefore, measuring the gas species concentration distribution can
help understand the results shown in Figs. 18 and 19 more clearly.
4.3. Gas species concentration measurement
Measuring the distribution of reactants provides the information about reactant delivery, and measuring the water vapour
distribution helps understand the measured HFR distribution
(related to membrane hydration/dehydration). The water amount
0.5
0.14
Water V apour Mole Fraction
Hig h Frequency Resistance, Ω cm
2
Cell voltage = 0.8 V
Cell voltage = 0.65 V
Cell voltage = 0.5 V
0.4
0.3
0.2
Cell voltage = 0.8 V
Cell voltage = 0.65 V
Cell voltage = 0.5 V
0.1
0
0.12
0.1
0.08
0.06
0
0.2
0.4
0.6
0.8
1
Fractional Distance from Cathode Inlet
Fig. 19. Measured HFR along fractional distance from cathode flow channel inlet at
different cell voltages [168].
0
0.2
0.4
0.6
0.8
1
Fractional Distance from Cathode Inlet
Fig. 20. Measured water vapour mole fraction in cathode flow channel along fractional
distance from inlet at different cell voltages [168].
246
K. Jiao, X. Li / Progress in Energy and Combustion Science 37 (2011) 221e291
was first measured at the flow channel outlet by collecting liquid
water and condensing water vapour [169,170]. Knowing the water
supply and removal rates at the inlets and outlets on both sides, the
rate of membrane hydration/dehydration and the water transfer
rate through membrane can be estimated. Measuring the gas
species concentrations at different locations of PEMFC was first
carried out by Mench et al. [171]. In the measurement of [171], gas
sampling ports were placed at different locations along anode and
cathode flow channels, and a micro gas chromatograph (GC) system
was used to measure the gas species concentrations corresponding
to the gas sampling ports at different locations. More advanced
experimental measurements to simultaneously measure current,
HFR and gas species concentration distributions have also been
developed [168,172,173], and the simultaneously measured results
are more valuable and can be more easily understood than a single
set of measurement data.
Corresponding to the same cell and operating condition for Figs.
18 and 19, Fig. 20 shows the measured water vapour mole fraction
in cathode flow channel along the fractional distance from inlet at
different cell voltages [168]. It can be noticed that due to the water
production and EOD causing water migrating from anode to
cathode, the water vapour mole fraction increases along cathode
flow channel from inlet. The increasing water vapour mole fraction
along cathode flow channel in Fig. 20 and the dramatically
changing HFR in Fig. 19 do not agree with each other well, because
the water mole fraction in anode perhaps decreases along flow
direction, indicating that simultaneous measurements of gas
species concentrations on both sides are desirable.
4.4. Temperature distribution measurement
Water phase change, membrane hydration/dehydration and
electrochemical reaction kinetics are significantly affected by local
temperature in PEMFC. Therefore, measuring temperature distribution provides valuable information for thermal and water
management. Temperature distribution can be measured by using
infrared (IR) cameras [158,174,175] (the thermography technique).
Wang et al. [174] designed a PEMFC with an optical window on the
anode side allowing IR light. Two-dimensional temperature
distribution on the MEA surface was obtained under different
operating conditions. Shimoi [175] applied the thermography
technique to an operating test cell in a manner similar to [174] as
well. The thermography technique was also applied simultaneously
with other measurements such as current distribution and liquid
80
Anode
Cathode
o
Temperature, C
75
70
65
60
0
0.2
0.4
0.6
0.8
1
Fractional Location Along Flow Channel
Fig. 21. Measured temperature in anode and cathode flow channels along flow
direction [181].
water visualization (the optical window allows both the temperature measurement and liquid water visualization) [158]. The main
drawback of thermography technique is that it requires major
modifications to cell design and component material due to the
requirement of optical window.
Another way to measure temperature distribution is inserting
micro-thermocouples at different locations of PEMFC. Mench et al.
[176] measured the temperature distribution at different positions in
a MEA by embedding eight micro-thermocouples. Such measurement is not easy to conduct due to the PEMFC configuration, and it is
also difficult to prevent the destruction of the thermocouple when
clamping the cell. Vie and Kjelstrup [177] measured the temperature
profile in the MEA of a PEMFC by using micro-thermocouples. It was
shown that the temperature gradient across the MEA surface is not
negligible. The measurement in [178] placed micro-thermocouples
in the lands (ribs) of bipolar plate (BPs) in direct contact to GDL
surface along the flow direction for both the anode and cathode.
Similarly, temperature distribution measurements in PEMFC by using
micro-thermocouples were also carried out in [179,180].
Recently, Alaefour et al. [181,182] conducted non-destructive
temperature distribution measurements for a PEMFC with straightparallel flow channel design. 23 micro-thermocouples were
embedded in the arrays of blind holes along the flow channels and
lands (ribs). Temperature distributions have been obtained for two
principle directions: parallel and normal to the direction of flow
channel for both the anode and cathode. The obtained results clearly
indicated that the temperature distribution inside PEMFC is very
sensitive to operating current density. Almost uniform temperature
distribution inside PEMFC was observed at low current densities,
and the temperature variations were considerable at high current
densities. Fig. 21 shows the measured temperature in anode and
cathode flow channels along the flow direction [181] for the cell
operating at 0.6 V. It can be noticed that the highest temperature
locations are close to the middle along the flow direction, possible
explanation is that the membrane is well hydrated due to product
water accumulation there and the reactant concentrations are still
sufficiently high. Simultaneous measurements of current, HFR, gas
species and temperature distributions have not been conducted yet
to the best of the authors’ knowledge, which are expected to provide
more valuable information for thermal and water management.
4.5. Water visualization
Since water management is one of the most important issues for
PEMFC, investigation of detailed water behaviours inside PEMFC is
therefore important. Experimental methods for investigating water
behaviours include: direct imaging on liquid water in transparent
PEMFCs [183e189], neutron radiography/tomography [190e198]
and X-ray micro-tomography [199,200].
Similar to the temperature distribution measurement using IR
cameras mentioned in the previous subsection, optical window is
also required for direct imaging on liquid water in PEMFC. ChargeCoupled Device (CCD) camera is often used to capture detailed
liquid water movement through optical window. Tuber et al. [183]
visualized liquid water transport in cathode flow channels of
a transparent PEMFC, and it was found that the air stoichiometry
ratio, temperature, inlet air relative humidity and GDL property all
have non-negligible influence on liquid water transport. Not only
focusing on liquid water transport in cathode flow channel, the
water emerging process from cathode GDL surface was visualized
by Wang and co-workers [184,185], it was reported that water
droplets emerging from the cathode GDL surface only appear at
preferential locations, and can grow to a size comparable to the
flow channel dimension under the influence of surface adhesion.
Rather than only focusing on the liquid water transport in cathode,
K. Jiao, X. Li / Progress in Energy and Combustion Science 37 (2011) 221e291
Spernjak et al. [186] and Ge and Wang [187] also investigated liquid
water transport in anode flow channel, the visualizations in
[186,187] all confirmed that the anode and cathode electrode
wettabilities have significant impacts on the presence of liquid
water in anode (i.e. increasing the hydrophobicity of cathode
electrode and hydrophilicity of anode electrode results in more
liquid water in anode due to the enhanced water transfer from
cathode to anode through membrane). Such observation explains
the reason that using hydrophobic MPL between cathode CL and
GDL could result in better cell performance since membrane can be
better hydrated. The effects of various operating conditions on
liquid water transport in cathode flow channel have also been
studied by Hussaini and Wang [188].
In fact, building transparent PEMFC and direct imaging on liquid
water is the only available experimental method to capture the
detailed liquid water flow behaviours (the neutron and X-ray
methods to be discussed later on this subsection cannot distinguish
the state of water). However, most of the transparent materials are
not electrically conductive: if the BP is purely transparent, the cell
may not work because electrons cannot be transferred. Therefore,
present transparent PEMFCs still use electrically conductive materials as the land to contact the GDL directly, and the only visible
place is inside the flow channel, this is why the experimental
investigations reported in [183e188] all showed real-time liquid
water behaviours in flow channels but they all neglected the cross
flow under the land area. Park and Li [201] conducted both
numerical and experimental investigations on the cross flow
through the GDL, and they reported that the pressure drop could be
reduced by up to 80% due to the cross flow through the GDL under
the land area. Therefore, experimental investigations of detailed
water transport behaviours caused by such cross flow under the
land are necessary for better water management of PEMFCs. Based
on this premise, a transparent PEMFC with serpentine flow channel
247
design and with both the optical land and flow channel was build by
Jiao et al. [189], promising liquid water visualization in both the flow
channel and under the land. The GDL thickness was carefully
controlled by inserting metal shims with different thicknesses in
parallel with GDL. Fig. 22 shows the visualization of liquid water
transport in both the flow channel (Fig. 22a) and under the land
(Fig. 22b) [189]. The cell was initially flooded with liquid water, and
then the water removal characteristics were investigated. Fig. 22a
shows that liquid water films sticking on flow channel walls are hard
to be removed, and Fig. 22b shows liquid water flowing under the
land. The visualization in Fig. 22b confirms that the cross flow under
land plays an important role in liquid water removal.
Neutron method relies on the nature of neutron beam: it could
detect organic hydrogen-containing substances, and this feature is
suitable for PEMFCs since water is the only substance that could be
detected. However, it is difficult to use neutron method to distinguish between liquid water and vapour. In addition, for neutron
radiography method [190e195], the through-plane location of
water (to indicate water in cathode, anode, membrane, CL, GDL, flow
channel) is also difficult to be determined since only two-dimensional images for the two in-plane directions can be obtained. The
neutron imaging results from Geiger et al. [190] and Pekula et al.
[191] both showed significant water concentration in the flow
channels, especially at the downstream and at the serpentine
corners. Hickner et al. [192] observed that the amount of water
accumulation changes dramatically with current density and
increasing the reactant flow rate also facilitates water removal.
Owejan et al. [193] showed neutron images for an interdigitated
flow field, and they reported that water accumulation in the GDL
reduced the GDL permeability significantly. Zhang et al. [194]
reported that GDL properties such as wettability, porosity etc.
affect the characteristics of water removal significantly. With
a special design of serpentine flow channels on both sides, the
Fig. 22. Visualization of liquid water transport (a) in the flow channel and (b) under the land of a transparent PEMFC with serpentine flow channel design [189].
248
K. Jiao, X. Li / Progress in Energy and Combustion Science 37 (2011) 221e291
Fig. 23. Neutron images at different current densities with 3 min duration for each load (a: current density is increased up to 0.75 A cm
1 A cm 2) [195].
neutron imaging conducted by Park et al. [195] was able to distinguish the water in anode and cathode flow channels, and the
dynamic response of PEMFC together with water distribution was
analyzed. Satija et al. [196] used neutron tomography method to
reconstruct three-dimensional water distribution for an inactive
PEMFC, however, real-time imaging was only obtained in twodimensional by using neutron radiography method. Hickner et al.
2
; b: current density is increased up to
[197] visualized the water profile on a cross section of an operating
PEMFC by using neutron radiography method. At different operating current densities, the water amounts in MEA and flow channels were estimated based on the operating conditions. Based on
neutron imaging method, Turhan et al. [198] analyzed the throughplane liquid water accumulation, distribution and transport in
different components of a PEMFC with different levels of channel
K. Jiao, X. Li / Progress in Energy and Combustion Science 37 (2011) 221e291
a
5. Overview of numerical models
1
Increasing current density
Decreasing current density
Cell V oltag e, V
0.9
0.8
0.7
0.6
0.5
0.4
0
0.2
0.4
0.6
0.8
1
-2
Current Density, A cm
b
1
Increasing current density
Decreasing current density
0.9
Cell V oltag e, V
249
0.8
0.7
0.6
0.5
0.4
0
0.2
0.4
0.6
0.8
1
Current Density, A cm-2
Fig. 24. Dynamic cell performance corresponding to the test conditions of Fig. 23 (a:
current density is increased up to 0.75 A cm 2; b: current density is increased up to
1 A cm 2) [195].
wall hydrophobicity. X-ray method can also be used to detect water
[199,200], however, X-ray method is relatively unstable because the
X-ray beam is easy to be scattered and absorbed by electrons, and
this feature makes X-ray method less popular than Neutron method.
Fig. 23 [195] shows the neutron images at different current
densities with 3 min duration for each load, it can be noticed from
Fig. 23 that more water is accumulated at the serpentine corners,
which agrees with the liquid water visualization shown in Fig. 22a
[189]. The water distributions at the same current density when
increasing and decreasing the load are also different, corresponding
to the test conditions of Fig. 23, the cell dynamic performance
shown in Fig. 24 indicates different cell voltages at the same current
density when increasing and decreasing the load. The different cell
voltages are caused by the dynamic liquid water transport blocking
the reaction sites, indicating that effective control of liquid water in
PEMFC is essential to achieve consistent cell performance.
4.6. Summary
Experimental observations on the distributions of current, HFR,
gas species, temperature and water are reviewed in Section 4. The
distributions of many parameters can be obtained based on the
presently available experimental techniques, and simultaneously
measuring more parameters with minimum modification of cell
and system design is the primary target of future experimental
observations.
As reviewed in Section 4, many experimental studies have been
conducted to observe the various transport phenomena in PEMFC.
However, due to the drawbacks such as difficulty to perform the
different experimental measurements simultaneously, unrealistic
operating conditions due to the modified cell and system designs
for experimental measurements, and high cost for materials and
testing instruments, numerical modeling of PEMFC is therefore
critical for better understanding of transport phenomena in PEMFC.
Up to early 2000s, excellent reviews on the modeling work for
PEMFC mainly focusing on the macroscopic (top-down) first-principle-based models can be found in [143,202], and on the
researches related to the proton conductor of PEMFC (e.g. Nafion
membrane) in [100,203]. In this paper, both the first-principlebased (from atomistic to full cell levels) and rule-based models are
comprehensively reviewed. The first-principle-based models rely
on solving a set of governing partial differential equations including
both the top-down (e.g. solving continuity, NaviereStokes and
other equations) and bottom-up (e.g. solving Boltzmann equations)
approaches, and the rule-based models depend on applying physical rules to simplified or real physical structures. The first-principle-based models are categorized into three levels in this paper
based on the characteristic length and numerical methods:
microscale, mesoscale and macroscale, as shown in Fig. 25 and
detailed in Section 5.1. This paper therefore provides a comprehensive review on most of the PEMFC related models on all the
levels of scale. In this section, the numerical models, including both
the first-principle-based and rule-based models, are summarized
to guide the detailed discussions on the different models in
Sections 6e12. A summary of the level of scale for the first-principle-based models is given in Section 5.1, followed by a discussion
on the model development history in Section 5.2.
5.1. Level of scale
Most of the previously developed PEMFC models are first-principle-based, from atomistic to full cell levels. The number of rulebased model is much less, and such models mainly focused on estimating liquid water transport in PEMFC electrode by applying physical rules to simplified or real electrode micro-structures. Therefore, in
this paper, only the first-principle-based models are classified into
different levels of scale, which are microscale, mesoscale and
macroscale. Fig. 25 illustrates the levels of scale for first-principlebased PEMFC models with representing phenomena and numerical
methods. The first-principle-based models depend on either the topdown or bottom-up approach. The top-down approach relies on
solving a set of governing partial differential equations essentially
based on the continuum assumption, and these partial differential
equations include continuity, NaviereStokes and/or other equations
for conservations of the macroscopic properties such as mass,
momentum, energy and so on. The bottom-up approach includes the
molecular dynamics (MD), off-lattice pseudo particle (such as dissipative pseudo particle and Monte Carlo (MC)), lattice gas (LG) and
lattice Boltzmann (LB) methods, by solving the partial differential
equations for the motions of a set of single molecules (or atoms) or
a set of pseudo particles (the pseudo particle represents a group of
molecules). The name of “top-down” is obtained because this
approach relies on the solutions of the macroscopic properties, and
the solutions for the motions of molecules/atoms or pseudo particles
pursue “bottom-up” strategies, so-called the bottom-up approach.
The top-down approach (the continuum method shown in
Fig. 25) can only be applied for small Knudsen number regimes (on
the macroscale defined in this paper, as shown in Fig. 25). By
neglecting the real micro-structures of GDL, CL and membrane, the
250
K. Jiao, X. Li / Progress in Energy and Combustion Science 37 (2011) 221e291
Fig. 25. Level of scale for PEMFC modeling with representing phenomena and numerical methods (only first-principle-based numerical methods are shown).
K. Jiao, X. Li / Progress in Energy and Combustion Science 37 (2011) 221e291
top-down approach can be applied to all the cell components by
considering modeled effective transport properties (such as the
effective diffusion coefficient in Equation (30)) and/or other
assumptions (such as homogeneous material of GDL, CL and
membrane, or assuming simplified CL and membrane structures
such as the agglomerate models). In fact, most of the previously
developed models rely on using the effective transport properties
and assuming homogeneous materials, therefore all the cell
components can be accounted for simultaneously, allowing full cell
modeling. By considering the real micro-structures of GDL, CL and
membrane, as analyzed in Section 3.3, the small pores on the level
of nanometre in CL and membrane prohibit the use of the topdown approach, while the top-down approach can still be used in
GDL with the pores on the level of micrometre. For the bottom-up
approaches, the MD and pseudo particle simulations can be used to
investigate the transport phenomena and material self-organization in membrane and CL on the microscale level (as defined in this
paper and shown in Fig. 25), which provide the fundamental
understanding on the transport mechanisms to guide the analysis
on the grander scales. The office-lattice pseudo particle, LG and LB
methods solve the motions of pseudo particles which include
a number of real molecules, therefore the computational time can
be saved and the simulations can be performed on a larger scale
(e.g. the mesoscale defined in this paper, as shown in Fig. 25). The
current computational power also allows the LG and LB simulations
performed on the macroscale. In fact, the LG and LB methods allow
the micro-structure to be more easily considered than the
continuum method (the top-down approach), and the LB method
therefore has been used to investigate the liquid water transport in
the pores of GDL, CL and membrane and in flow channel. Hence the
LG and LB methods can be considered as the scale-bridging method
which can be used to capture the transport phenomena on the
multi-scales. Detailed discussions on the various numerical models
will be given in Sections 6e12 with their applications to different
components of PEMFC.
5.2. History of model development
The numerical models developed by Springer et al. [72,204] and
Bernardi and Verbrugge [109,205] in early 1990s are usually referred
to as the pioneering modeling work for PEMFC. These models are
essentially one-dimensional models considering the membrane, CL
and GDL based on the continuum (top-down) approach by solving
the conservation equations by assuming homogeneous materials
and using effective transport properties. The water diffusion
through the membrane and the effects of water content on the
membrane conductivity are accounted for by Springer et al. [72,204],
while Bernardi and Verbrugge [109,205] assumed constant water
content (fully hydrated) across the membrane. Effective transport
properties are used such as the effective diffusion coefficient by
using the Bruggeman correlation. In these one-dimensional models
[72,109,204,205], the fundamental framework and most of the
fundamental formulations for PEMFC modeling based on the
continuum (top-down) approach are established and have been
widely used in many of the later numerical studies. Following
[72,109,204,205], Nguyen and White [206] and Fuller and Newman
[207] developed pseudo two-dimensional models by further
considering the flow channel with the along-the-channel direction,
which considers the effect of inlet water humidity and temperature
distributions, providing more detailed water and thermal management capability. After that, more models which are similar to
[72,109,204e207] have also been developed [208e210].
As interest grew in fuel cells in late 1990s, more and more
numerical models were developed. Yi and Nguyen [211,212] and
Gurau et al. [213] all developed two-dimensional models to explore
251
more detailed transport phenomena in PEMFC, these models illustrated the utility of multi-dimensional models in the understanding
of the internal conditions of PEMFC, such as the reactant and water
distributions. Based on the continuum (top-down) approach,
simplified CL structures were proposed by assuming that the ionomer and platinum/carbon particles form large agglomerates on the
level of micrometre, so-called the agglomerate models, which had
been developed by Gloaguen and Durand [214], Bultel et al.
[215e217] and Marr and Li [218]. Investigations on the water and
proton transport through the membrane have also been carried out
based on the continuum (top-down) approach [219e223]. Not only
on the macroscale, physical models with assumed randomly
distributed mesoscale pores were studied to relate the water content
and membrane conductivity [73], and MD simulations [224,225]
were also conducted to study the self-organization of membrane.
In 2000s, multi-dimensional models based on the continuum
(top-down) approach and solving a complete set of conservation
equations (continuity, NaviereStokes, Energy and so on) coupled
with electrochemical reactions, were developed by many
researchers. Computational fluid dynamics (CFD) code (such as the
commercial code Fluent, Star-CD, CFX, CFD-ACEþ through their
user coding capability) based on finite volume or finite element
methods were modified and used to develop such models, and
more complex geometry and transport phenomena were able to be
investigated. With three-dimensional geometry considered, the
models of Dutta et al. [226,227], Zhou and Liu [228], Berning et al.
[229], Mazumder and Cole [230], Lee et al. [231], Um and Wang
[232] and Wang and Wang [233] are considered as the pioneering
work in this field, these models mainly considered a single flow
channel with the major components (e.g. flow channel, GDL, CL,
membrane, and some with BP), and some of them accounted for the
transient calculation as well, such as in [233]. Large scale simulations considering multi-channel or small stacks were also carried
out [234e239]. Note that the multi-dimensional models in
[211e213,226e239] neglected liquid water formation by only
solving a single equation for vapour and liquid water, therefore
super saturated water vapour was observed, and some of the
models calculated the amount of liquid water during the post
processing (e.g. [226,227,231]). Real two-phase models have also
been developed, in which vapour and liquid water move at
different velocities, and the other liquid water effects were
accounted for, such as the surface tension and water flooding
effects. Generally, two major models, two-fluid model and mixture
model, were widely used and considered as the real two-phase
models. The two-fluid model is mainly attributed to the work of
Nguyen and co-workers [80,134,138,240,241], Djilali and coworkers [242e244], Mazumber and Cole [245] and Wu et al.
[137,246]; and the mixture model was mainly developed by Wang
and co-workers [247e255] and You and Liu [256,257], Note that
here “single-phase” and “two-phase” represent the state of water
only in pores of GDL and CL and in flow channel, therefore “singlephase” indicates that only water vapour is considered in these
regions, and “two-phase” means that both the vapour and liquid
water are accounted for in these regions. However, the state of
water in the ionomer of the membrane and CL is different (as discussed in Section 2), and a separate equation was used to solve the
water transport in ionomer (e.g. [80,137,246], the treatment of
water/proton transport in membrane will be detailed in Section 7).
Perhaps “multiphase” is more accurate in presenting the state of
water. The two-fluid model solves the mass, momentum and
species transport conservation equations for the gas mixture, with
an extra liquid water transport conversation equation; and the
mixture model solves the mass and momentum and species
transport conservation equations for the two-phase mixture
(mixture of liquid water and gas) mainly based on the mass-
252
K. Jiao, X. Li / Progress in Energy and Combustion Science 37 (2011) 221e291
averaged properties of the two-phase mixture. Both models are
able to predict the velocities for both the vapour and liquid water. In
consideration with phase change of water, the mixture model
essentially accounts for an infinitely large (instantaneous) phase
change process, while the phase change process in the two-fluid
model can be modeled at different rates (constant phase transfer
rates were assumed in previous studies). Detailed discussions on
the multiphase multi-dimensional models for PEMFC based on
continuum (top-down) approach are given in Section 6. Recently,
multiphase multi-dimensional modeling of PEMFC starting from
subzero temperatures (cold start) was also carried out, by Wang
and co-workers based on the mixture model framework [258e260]
and by Jiao and Li based on the two-fluid model framework
[62,63,261]. In the work of [258e260], liquid water formation was
not considered therefore the simulations were limited at subzero
temperatures, and the water freezing in ionomer was not accounted for as well. Both the ice and liquid water formations, as well as
water freezing/melting processes in ionomer were all included in
the models in [62,63,261], therefore these models provided the
capability for simulations in the whole temperature range (from
subzero to normal operating temperatures).
The two-fluid and mixture models can only be used to estimate
the two-phase concentrations (e.g. the volume fractions of liquid and
gas phases), and the detailed liquid water transport behaviours
cannot be investigated because these models do not allow interface
tracking between liquid water and gas. Therefore, in 2000s, rather
than focusing on increasing the size of the computational geometry,
numerical models focusing on more detailed transport characteristics were also developed. Based on the volume-of-fluid (VOF) model,
the liquid water dynamics in a single serpentine flow channel was
investigated by Quan et al. [262], and in small PEMFC stacks with
straight-parallel and serpentine-parallel flow channels by Jiao et al.
[11,12]; and the effects of surface wettability of flow channel on
liquid water behaviours were also investigated [263,264]. The water
transport behaviours in both the simplified [265e267] and real [268]
micro-structures of GDL were investigated by Jiao and co-workers by
using the VOF model, in which the effects of surface wettability were
investigated as well [266,268]. The development of agglomerate
models considering simplified structures of CL was also continued in
2000s [269e271] based on continuum (top-down) approach.
Models based on bottom-up approach were also developed in
2000s for investigating liquid water transport in GDL and CL. With
micro-structures of GDL and CL, LB method was used to simulate the air
flow to obtain the GDL properties (such as permeability) [272,273], and
to simulate liquid water transport in GDL and CL [274e277].
Construction of realistic micro-structures of GDL and CL is an essential
prerequisite for such simulations, and this can be achieved by either 2D
or 3D imaging (such as by using X-ray and magnetic resonance microtomography) or stochastic models (generating micro-structures by
computational simulations). Rather than focusing on first-principlebased models, rule-based models were also developed, such as full
morphology (FM) models [274,275,278] and pore-network (PN)
models [274,279] mainly for investigating liquid water behaviours in
GDL (such as liquid water effects on capillary pressure). Models based
on bottom-up approach on microscale and mesoscale were also
developed in 2000s, mainly focused on the understanding of water and
proton transport and material self-organization. MD based on
quantum mechanical [78,280,281] and classical [282e290] theories
and pseudo particle [291,292] simulations were conducted for Nafion
membrane. Quantum mechanical MD and MC simulations have been
performed to study the elementary reaction processes on catalyst
surfaces [293e298]. Mesoscale simulations to evaluate key factors for
CL fabrication by investigating material self-organization (sizes of
platinum/carbon/ionomer agglomerates and pores) were also carried
out [299].
5.3. Summary
Section 5 provides an overview of the numerical models for PEMFC
to guide the detailed discussions in the following sections. The review
shows that numerical models based on different principles and physical rules on all the levels of scale have been extensively developed. In
the following sections, full cell models based on continuum (top-down)
approach are described in Section 6 to provide a complete view for the
various transport processes, followed by Sections 7e12 focusing on
a specific cell component or operating condition in each section.
6. Multi-dimensional multi-component multiphase model
with full cell geometry
With the currently available computational power, continuum
(top-down) approach by considering homogeneous materials of
GDL, CL and membrane with modeled effective transport properties
is perhaps the only way to model a full PEMFC (including major or all
cell components). By solving a number of conservation equations in
a multi-dimensional computational domain, the multi-component
multiphase transport with electrochemical reactions and electron/
proton transport has been successfully modeled in the past decade,
and as reviewed in Section 5, the two major types of such models are
the two-fluid model and mixture model. The names of two-fluid and
mixture are obtained based on the modeling approaches of water
(vapour and liquid) transport in the pores of GDL and CL and in flow
channel. However, different modeling approaches of water transport in ionomer have also been used and can be classified into three
groups, namely, the hydraulic (or convective) model [74,205,229],
diffusive model [72,300e302], and chemical potential model
[75,149,303e307]. The hydraulic model assumes a fully humidified
membrane therefore the major water transport mechanism
becomes convection (because the pores in membrane is enlarged by
water). However, in an operating PEMFC the ionomer close to anode
usually dries out quickly and water is produced in cathode, resulting
in uneven water distribution, therefore the hydraulic model has only
been used for the early days’ modeling work, and is rarely considered
nowadays. The diffusive model accounts for diffusive water transport in ionomer by incorporating with experimentally measured
diffusion coefficient of water. Therefore unevenly distributed water
can be accounted for in the diffusive model, which models a more
realistic condition. In the diffusive/hydraulic models, the proton
concentration is assumed constant over the ionomer domain and
therefore the proton transport can be easily solved by using Ohm’s
law (water content only affects proton conductivity of ionomer). In
the chemical potential model, however, the concentration of mobile
protons is assumed to vary with the water concentration. Utilizing
the dusty fluid model [149,303], generalized StefaneMaxwell
equations [305e307], or concentrated solution theory [75,304], the
water and proton transport are strongly coupled and solved simultaneously. The chemical potential model might be considered as
a superclass of diffusive/hydraulic models; the diffusive/hydraulic
models are only valid in certain situations (constant proton
concentration), while the chemical potential model is a more
comprehensive approach which applies to a much larger range.
Nevertheless, the present chemical potential models are invariably
confined to the membrane region along with many simplifications.
Further, several parameters and correlations related to this model
class remain unknown, such as the diffusion coefficient of hydrogenewater ions (e.g. hydronium), the interaction properties of water
and proton, etc. Hence, the application of the chemical potential
model in full cell modeling needs to be explored further. Therefore,
most of the two-fluid and mixture models still assume constant
proton concentration in membrane, and the diffusive model is most
widely used. The EOD causing water migrating from anode to
K. Jiao, X. Li / Progress in Energy and Combustion Science 37 (2011) 221e291
cathode, as well as the convective transport of water through
membrane (usually can be neglected but might be considered when
the cell is pressurized differentially) can also be accounted for by
adding extra terms on the water transport equation. The details
about modeling water transport in membrane are given in Section 7.
For water transport in pores of CL and GDL and in flow channel,
three models based on continuum (top-down) approach have been
used, which are the two-fluid model, mixture model and VOF
model. The two-fluid model solves individual sets of equations for
each phase while the interaction among different phases is
explicitly taken into account through limited phase transfer terms.
The mixture model is a kind of single-fluid model. It solves a single
set of conservation equations for the phase mixture assuming
phase equilibrium, and the volume fractions of the phases, as well
as the relative velocity among different phases are estimated
subsequently. In PEMFC modeling, the two-fluid and mixture
models are usually simplified by combining the momentum and
other equations with the help of Darcy’s law and a capillary pressure function. Power law relations and Leverett J-functions were
widely used to calculate capillary pressure (e.g. [247,308]), and
recent studies [129e133] also provide alternative correlations.
Compared to the mixture model, the advantage of the two-fluid
model is that only one extra equation for liquid saturation is added,
253
while allowing for the simulation of non-equilibrium phase transfer processes. Rather than using the two-fluid and mixture models,
two-phase water transport in pores of CL and GDL and in flow
channel can also be modeled by using the VOF model. The biggest
advantage of the VOF model is its ability to trace the trajectory of
the liquid water movement. However, due to the nature of the
extremely small time-steps and intensive computing time related
to VOF methods, its application so far has been restricted to
investigating the liquid behaviour in the electrode [265e268] or
flow channels [11,12,262e264]. Full geometry PEMFC models that
incorporate the VOF approach have also been conducted [309,310],
but only with very limited time instances considered due to the
computational power limitation. It should be noticed that the
transport in real micro-structure of CL cannot be model by any topdown models, due to the presence of nanometre pores.
In summary, the “chemical potential þ VOF” approaches may
finally evolve to be the main features of the next generation of
PEMFC models. At the current stage, however, the “diffusive þ twofluid/mixture” type models still dominate. Therefore, in this section,
only the two-fluid and mixture model are described, all with diffusive model for water transport in ionomer. The computational
domain required for such full cell models is first introduced. Then
the two-fluid model and mixture model are introduced one by one,
followed by the specifications of boundary condition and numerical
implementation. The comparison between the two models is given
as well with representing simulation results.
6.1. Computational domain
The requirements of computational domain for the different full
cell models are the same. Fig. 26 shows the sample computational
domain and mesh including all major cell components (BP, flow
channel, GDL, CL and membrane) for a single straight PEMFC. It can
be noticed that the micro-structures of GDL, CL and membrane are
all neglected and they are considered as homogeneous layers.
Therefore, structured mesh can be applied to the whole computational domain, which improves both the computational accuracy
and efficiency. Usually 10 10 layers of grid are needed on the cross
sections of each layer normal to the flow direction, and the number
of grid along the flow direction depends on the flow channel length/
geometry and other conditions (e.g. 100 layers of grid along the flow
channel direction is usually considered sufficient for a single straight
PEMFC). In fact, the number of grid layers along the through-plane
direction (normal to membrane surface) cannot be too high in CL,
due to the fact that the CL is very thin (usually around 0.01 mm),
which may cause too large aspect ratios of the computational cell
and lead to computational instability. The number of grid cell for the
single straight PEMFC shown in Fig. 26b is 76,000, such computational domain allows both the steady and unsteady simulations on
a single computer. For multi-channel and stack simulations, parallel
computing is an effective way to handle large number of grid cell. For
example, using 24 processors in parallel is considered sufficient for
the number of grid cell on the level of 106.
6.2. Two-fluid model
The two-fluid model presented in this section can be concisely
summarized by the following conservation equations [80,134,137,
138,240e246].
Mass of gas mixture (solved in flow channel, GDL and CL):
Fig. 26. Sample (a) computational domain and (b) mesh including all major cell
components for full cell modeling.
v
3 1
vt
!
slq rg þ V$ rg u g ¼ Sm
(61)
Momentum of gas mixture (solved in flow channel, GDL and CL):
254
K. Jiao, X. Li / Progress in Energy and Combustion Science 37 (2011) 221e291
Table 3
Source terms in the conservation equations (Equations (61)e(68)) except those for water (for two-fluid model) [137].
Domain
Sm
Su
Si
Sele
Sion
ST
BP
Flow channel
GDL
0
Svp
Svp
0
0
0
0
0
0
0
0
0
0
0
kVfele k2 keff
ele
Spc
kVfele k2 keff
þ Spc
ele
ja
ja jhact j þ kVfele k2 keff
þ kVfion k2 keff
ion þ Spc
ele
mg !
Kg
Anode CL
Cathode CL
Kg
ug
SH2 ¼
mg !
SO2 þ Svp
Membrane
ug
mg !
SH2 þ Svp
Kg
0
ug
SO2 ¼
0
ja
M
2F H2
jc
M
4F O2
0
!
! !
rg !
rg !
ug
ug ug
v
þ V$
2
vt 3 1 s
lq
32 1 slq
!
!!
!T
!
ug
ug
þV
¼ Vpg þ mg V$ V
3 1 slq
3 1 slq
!!
!
u
2
m V V$ g
þ Su
3 g
3 1 s
(62)
Gas species (solved in flow channel, GDL and CL, i represents
hydrogen, oxygen or vapour):
!
slq rg Yi þ V$ rg u g Yi ¼ V$ rg Deff
i VYi þ Si
(63)
Liquid water (solved in flow channel, GDL, CL):
v 3slq rlq
vt
!
þ V$ irlq u g ¼ V$ rlq Dlq Vslq þ Slq
(64)
Non-frozen membrane water (the total amount of water in
membrane for normal operating conditions) (solved in membrane
and CL):
rmem v ulnf
EW
vt
¼
rmem
EW
V$ Deff
nmw Vlnf þ Snmw
(65)
Energy (solved in whole computational domain):
v
vt
rCp
eff
T
fl;sl
þ V$
rCp !
u
eff
T
fl
¼ V$ keff
fl;sl VT þ ST
(66)
Electronic potential (solved in CL, GDL, BP):
0 ¼ V$ keff
ele Vfele þ Sele
(67)
0 ¼ V$ keff
ion Vfion þ Sion
(68)
Ionic potential (solved in CL, membrane):
The above equations are closely coupled through the right hand
side source terms, which either stem from the electrochemical
Table 4
Source terms in the conservation equations (Equations (61)e(68)) for water (for
two-fluid model) [137].
Domain
Sv-l
Sv-l
Sv l þ Sn
Cathode CL
Sv
þ Sn
Membrane
Slq
Snmw
0
0
v MH2 O
Sv-l
Sv-l
Sv-l
v MH2 O
Sv-l
jc
2F
0
Svp
Flow channel
GDL
Anode CL
0
l
0
Sn
v
n
d eff
k Vf
F ion ion
n
þ V$ d keff
Vf
F ion ion
þ V$
Sn
v
jc
jc
0
lq
v
3 1
vt
ja
0
jc T D S
þ jc jhact j þ kVfele k2 keff
þ kVfion k2 keff
ion þ Spc
ele
2F
2 eff
kVfion k kion
reactions or from the interfacial mass transfer among different
phases. The expressions of these source terms have been summarized in Tables 3 and 4. The related mass transfer functions
accounting for the water phase change and membrane absorption/
desorption processes are described in Section 3, and therefore are
not repeated in this section. Most of the constitutive and empirical
formulas for transport properties, mass transfer rate and other
parameters for closing the conservation equations have also been
given in Sections 2 and 3, and therefore only those important and
not previously mentioned are described in this section.
It should be noticed that superficial velocity is used and all the gas
species are assumed to be ideal gas for the conservations equations
shown above (Equations (61)e(68)). By assuming incompressible
flow, the viscous force terms in the momentum equations for gas
mixture (Equations (62)) can be further simplified. Strictly speaking,
the diffusion terms in the gas species conservation equations
(Equation (63)) are only valid for binary diffusion (when only two
gas species are present). For PEMFC without impurities (e.g. CO), the
only gases in anode are hydrogen and water vapour, which is
acceptable; however, oxygen, water vapour and nitrogen are all in
gas phase in cathode when air is supplied, which requires StefaneMaxwell formulation for the diffusion terms. However, most of
the previous full cell models used simplified binary (Fick’s law)
diffusion formulation (the fraction of nitrogen is only estimated
from the fractions of oxygen and water vapour because the total
fraction is 1). The gas mixture pressure and velocity can be solved
through its mass and momentum equations (Equations (61) and
(62)), and the liquid water pressure, velocity and diffusion coefficient are derived from the capillary pressure and gas mixture pressure and velocity, as described in Section 3.2 and shown in Equations
(42)e(45). Therefore, only one extra equation to solve the liquid
water volume fraction is needed to account for liquid water transport. It should be noticed that this approach still needs better
correlations to relate the gas and liquid water pressures and velocities to better predict water transport in flow channel, and that is
why most of the previous models neglected liquid water transport in
flow channel. Previous study has assumed that the gas and liquid
water velocities are the same [80] (the interfacial drag coefficient i in
Equation (64) is equal to 1), which remains debated especially when
large liquid water droplets are present. The effect of liquid water
blockage on reaction rate in CL is usually represented by using
a linear relationship applied to the ButlereVolmer equation:
ja ¼ 1
jc ¼
1
0
10:5
c
2aa F a
H
@ 2A
h
exp
slq jref
0;a
RT act
cref
H2
c
O2
slq jref
0;c ref
cO2
exp
exp
4aa F c
hact þ exp
RT
2ac F a
h
RT act
(69)
4ac F c
h
RT act
(70)
255
K. Jiao, X. Li / Progress in Energy and Combustion Science 37 (2011) 221e291
where j (A m 3) is the reaction rate, slq the liquid water volume
fraction, c (kmol m 3) the concentration of gas species, a (0.5 for
anode and cathode) the transfer coefficient, F (9.6487 107 C kmol 1)
the Faraday’s constant, R (8314 J kmol 1 K 1) the universal gas
constant, T (K) the temperature, and hact (V) the overpotential (activational voltage loss). The subscripts/superscripts a and c represent
anode and cathode, respectively; and ref and 0 represent reference
states. Equations (69) and (70) indicate that when CL is fully flooded
by liquid water (slq ¼ 1), the reaction rate becomes 0.
The water transport in ionomer is solved in a single conservation equation (Equation (65)). This equation is solved in both
membrane and CL, membrane is considered to be full of ionomer
(the ionomer volume fraction u is equal to 1) and CL is partially
occupied by ionomer (u usually ranges from 0.2 to 0.4). The nonfrozen water diffusion coefficient in ionomer in CL therefore also
needs to be corrected in considering the volume fraction of ionomer, for example, the Bruggeman correlation can be used
1:5
Deff
Dnmw
nmw ¼ u
Deff
nmw
2
Heat is generated from the electrochemical reactions (activational
heat and reversible heat), electron and ion transport (ohmic heat),
and water phase change (latent heat), as shown in the sources
terms in Table 3. The non-frozen water in ionomer could be
assumed to be equivalent to liquid water [137], therefore the latent
heat for ionomer absorption/desorption with water vapour could
be the same as for water condensation/evaporation, and it is zero
for ionomer absorbing/desorbing liquid water.
The conservations of electronic and ionic potentials are given
in Equations (67) and (68), respectively. Ohm’s law is used for
such processes, by assuming that the concentrations of electron
and ion are constant in the electron conductive (e.g. BP, platinum
and carbon powers) and proton conductive (ionomer) materials.
This assumption is usually valid for electron transport but
remains debated for proton transport. The interaction between
proton and water in ionomer may need to be accounted for, and
the details will be given in Section 7. The corresponding source
terms (reaction rates) are added to the conservation equations, as
shown in Table 3. It should be noticed that the transient terms
are neglected in the conservation equations for the electronic
potential and ionic potential, the reason can be explained with
the help of Table 5. Table 5 shows an analysis of the time
constants of the fundamental transient phenomena in PEMFC,
which are gas transport, liquid water transport, non-frozen water
transport in ionomer, electrochemical double layer charging and
discharging, and heat transfer. Different components of PEMFC
are used to estimate the typical values of different time
constants. GDL is used to calculate the time constants of gas
transport and liquid water transport because the diffusion
dominated mass transport in GDL is slower than the convection
dominated mass transport in flow channel. Water transport in
ionomer typically occurs in membrane so that membrane is
considered to calculate the time constant. Electrochemical
double layer charging and discharging takes place in CL so that CL
is considered. Heat transfer is typically slow in membrane due to
its low heat conductivity, so that the time constant calculation for
heat transfer is conducted for membrane. Based on the calculated
values in Table 5, it can be noticed that the time constant of
electrochemical double layer charging and discharging is much
smaller than the other time constants, explaining why the transient terms can be safely neglected in the conservation equations
for the electronic potential and ionic potential.
(71)
1
where
(m s ) is the effective non-frozen water diffusion
coefficient in ionomer, and Dnmw (m2 s 1) the bulk non-frozen water
diffusion coefficient in ionomer (as given in Equations (16) and (17)).
It can be noticed from Table 4 that the product water is assumed to
be the non-frozen water in ionomer, by adding a source term (the
term jc/(2F) in Table 4) to the non-frozen membrane water conservation equation, due to the fact that water is produced at the interface of three-phase contact (ionomer, catalyst and reactants). The
Vfion Þ in Table 4)
water flux term due to EOD (the term V$ððnd =FÞkeff
ion
is therefore also added as a source term to the non-frozen membrane
water conservation equation. In fact, with the present model,
different water production (vapour, liquid water and water in ionomer) can be easily implemented by placing the water production
term in the corresponding conservation equations. Similarly, the
various water phase change processes can also be implemented by
simply adding source terms to the water conservation equations, as
shown in Table 4. The water phase change and ionomer absorption/
desorption rates can be controlled by adjusting the values of the
source terms. Therefore, the non-equilibrium water phase change
and ionomer absorption/desorption can be easily modeled. In fact,
rather than solving a conservation equation for non-frozen water in
ionomer, the water vapour conservation equation can be modified to
account for water transport in ionomer, by assuming equilibrium
ionomer absorption/desorption. This approach is introduced in
Section 6.3 together with mixture model.
The energy equation (Equation (66)) is solved in whole
computational domain, to fully account for heat generation and
transfer processes. Effective heat capacities and thermal conductivities need to be considered to include all the materials (gas
mixture, liquid water, ionomer, catalyst layer and other materials).
6.3. Mixture model
The mixture model can be concisely presented by the following
conservation equations [247e257].
Mass of gas and liquid water mixture (solved in flow channel,
GDL and CL):
Table 5
Time constants of various transient phenomena in PEMFC.
Phenomenon
Typical value
Gas transport
In GDL: dGDL z200 mm; Deff
g z10
Time constant
Liquid water transport
In GDL: dGDL z200 mm; Dlq z10
Non-frozen water transport in ionomer
In membrane: dmem z50 mm; rmem z1980 kg m 3 ;
Dlnf z10; Iz1 A cm 2 ; EWz1100 kg kmol 1 ; Fz96; 487 C mol
Electrochemical double layer
charging and discharging
In CL: dCL z10 mm; CL specific area ðaÞz105 m
Heat transfer
In membrane: dmem z50 mm; ðrCp Þeff
mem z1650 kJ m
5
m2 s
d2GDL
1
Deff
g
6
m2 s
1
d2GDL
Dlq
Electric capacity ðCÞz0:2 F m
keff
mem z1 W m
1
K
1
2;
1
1;
keff
z50 S cm 1 ; keff
ion z0:1 S cm
ele
3
K
1
;
1
z0:004 s
z0:04 s
2F dmem Dlnf rmem
z17 s
I EW
1
1
d2CL aC eff þ eff z0:2 ms; Too small, can be ignored.
kele
kion
d2mem ðrCp Þeff
mem
keff
mem
z0:004 s
256
K. Jiao, X. Li / Progress in Energy and Combustion Science 37 (2011) 221e291
Table 6
Source terms in the conservation equations (Equations 66-68 and 72-74) (for mixture model) [251, 252]
Domain
Su
Si
Sele
Sion
ST
BP
Flow Channel
0
0
0
0
0
0
0
0
kVfele k2 keff
ele
Spc
SH2 ¼ V$ðYH2 Jcap;lq Þ
0
0
kVfele k2 kele þ Spc
SH2 O ¼ V$ððYvp Ylq ÞJcap;lq Þ
SO2 ¼ V$ðYO2 Jcap;lq Þ
0
0
kVfele k2 kele þ Spc
ja
ja jhact j þ kVfele k2 keff
ele
m!
Anode GDL
K0
u
m!
Cathode GDL
K0
u
SH2 O ¼ V$ððYvp
m!
Anode CL
K0
u
SH2 ¼
eff
eff
Ylq ÞJcap;lq Þ
ja
ja
2F MH2
eff
þkVfion k2 kion þ Spc
þV$ðYH2 Jcap;lq Þ
SH2 O ¼
K0
Ylq ÞJcap;lq Þ
jc
4F MO2 þ
þV$ððYvp
m!
Cathode CL
eff
V$ðnFd kion Vfion ÞMH2 O
u
SO2 ¼
jc
jc
jc T DS
2F
SH2 O ¼
þ jc jhact j
eff
þkVfele k2 kele
V$ðYO2 Jcap;lq Þ
jc
2F MH2 O
þkVfion k2 keff
ion
þ Spc
þV$ðnFd keff
Vfion ÞMH2 O
ion
Membrane
þV$ððYvp
0
0
Ylq ÞJcap;lq Þ
v
!
ð3rÞ þ V$ðr u Þ ¼ 0
vt
0
(72)
Momentum of gas and liquid water mixture (solved in flow
channel, GDL and CL):
!
!!
ru u
v ru
þ V$
¼
vt 3
32
!!
!
!T
u
u
Vp þ mV$ V
þV
3
3
!
2
u
mV V$
þ Su
3
3
(73)
Species (solved in flow channel, GDL, CL and membrane,
i represents hydrogen, oxygen or water (vapour and liquid water
are accounted for together)):
v eff
3 rYi þ V$ðbr!
u Yi Þ ¼ V$ rg Deff
i g VYi
vt i
g
þ Si
(74)
Energy (solved in whole computational domain): can be represented by Equation (66)
Electronic potential (solved in CL, GDL, BP): can be represented
by Equation (67)
Ionic potential (solved in CL, membrane): can be represented by
Equation (68)
The source terms for the above conservation equations are given
in Table 6. It can be noticed that rather than solving the mass and
momentum equations for gas mixture as in the two-fluid model,
the mass and momentum conservations for the gas and liquid
water mixture is solved in the mixture model. Therefore, the source
term for the mass conservation equation is 0 (Equation (72)). For
incompressible flow, the transient term for the mass conservation
equation (Equation (72)) can be neglected, and the viscous stress
terms in the momentum conservation equations (Equation (73)) can
be simplified. The density (r, kg m 3) and dynamics viscosity (m,
kg m 1 s 1) of gas and liquid water mixture are calculated based on
the properties and volume fractions of gas mixture and liquid water as
r ¼ rlq slq þ rg 1
m ¼ h
slq
rlq slq þ rg 1
(75)
slq
i.
Klq =nlq þ Kg =ng
where rg (kg m 3) and rlq (kg m 3) are the densities of gas mixture
and liquid water, respectively; ng (m2 s 1) and nlq (m2 s 1) are the
kinematic viscosities of gas mixture and liquid water, respectively;
and Kg (m2), Klq (m2) and K0 (m2) are the gas, liquid water and
intrinsic permeabilities, respectively, as described in Section 3.2 and
Equations (33) and (34). Rather than solving a conservation equation
for liquid water transport as in the two-fluid model, the liquid water
volume fraction (slq) in the mixture model is calculated as
slq ¼
cH2 O
rlq =MH2 O
c
sat
(77)
csat
where cH2 O (kmol m 3) and csat (kmol m 3) are the vapour/liquid
water mixture and saturation water (can be obtained by using
saturation pressure of water by ideal gas law) concentrations,
respectively; and MH2 O (18 kg kmol 1) is the molecular weight of
water. As shown in the source terms of gas species in Table 6, the
capillary effect on gas species transport is also accounted for, and by
neglecting the effect of gravity, the capillary liquid water flux (Jcap,lq,
kg m 2 s 1) can be calculated as
Jcap;lq ¼
Yg Ylq
m
K0
(76)
(78)
rK0 Vpc
!
and liquid water velocity ( u lq , m s
1
) can be obtained as
!
rlq !
u lq ¼ Jcap;lq þ Ylq r u
(79)
where pc (Pa) is the capillary pressure, as described in Section 3.2
!
and Equations (42) and (43). u (m s 1) is the two-phase mixture
velocity. Yg and Ylq are the relative mobilities of gas mixture and
liquid water, given as
Ylq ¼
Klq =nlq
Klq =nlq þ Kg =ng
and
Yg ¼ 1
Ylq
(80)
In flow channel, it might be assumed that gas and liquid water
velocities are the same (as mentioned earlier with two-fluid model),
and better correlations need to be further explored. The relationship
!
between the two-phase mixture velocity ( u , m s 1), gas mixture
!
!
velocity ( u g , m s 1) and liquid water velocity ( u lq , m s 1) is
!
!
r!
u ¼ rg u g þ rlq u lq
eff
kVfion k2 kion
0
(81)
Rather than solving a separate conservation equation for water
transport in ionomer as in the two-fluid model, the water transport
in ionomer is obtained from the water (vapour and liquid mixture)
257
K. Jiao, X. Li / Progress in Energy and Combustion Science 37 (2011) 221e291
transport equation by modifying the transport properties. The
eff eff
effective porosities ð3eff
H2 O ; 3H2 ; 3O2 Þ and effective diffusion coefficient
of water vapour (Deff
H2 O
g,
m2 s 1, the subscript g indicates that only
vapour phase water is considered) are modified by considering
both pores and ionomer as
3eff
H2 O ¼ 3 þ u
rmem RT dlequil
EW psat
eff
3eff
H2 ¼ 3O2 ¼ 3 1
Deff
H2 O
g
slq
(82)
da
1:5
1
¼ Deff
vp ¼ 3
þ u1:5
(83)
slq
1:5
Dvp
rmem RT dlequil
EW psat
da
Dnmw
ð84Þ
where 3 and u are the volume fractions of pores and ionomer,
respectively; rmem (kg m 3) and EW (kg kmol 1) are the dry density
and equivalent weight of membrane; R (8314 J kmol 1 K 1), T (K) and
psat (Pa) are the universal gas constant, temperature and water
saturation pressure, respectively; lequil and a are the equilibrium
water content in ionomer and water activity in pores, and dlequil/da
can be obtained by differentiating Equation (12); and Dvp (m2 s 1)
and Dnmw (m2 s 1) are the bulk diffusion coefficients of water vapour
and the diffusion coefficient of non-frozen water in ionomer. By
neglecting the hydrogen and oxygen transport in ionomer, the last
term on the right hand side of Equation (84) can be neglected to
calculate the effective diffusion coefficients for hydrogen and oxygen.
This method can also be applied to the two-fluid model described in
Section 6.2.
In the mixture model presented this subsection, both the liquid
water volume fraction in pores and water content in ionomer can be
estimated by proper correlations. Without solving separate
conservation equations of liquid water and absorbed water in ionomer, the mass transfer rates of water phase change and ionomer
absorption/desorption cannot be implemented (as they are implemented in the two-fluid model described in Section 6.2 by adding
source terms to the different conservation equations). The estimated
amounts of liquid water and water content are based on the equilibrium condition without considering the non-equilibrium mass
transfer. Therefore, the mixture model presented in this subsection
can be regarded as the two-fluid model described in Section 6.2 with
infinitely large water phase change and ionomer absorption/
desorption rates (to ensure equilibrium states). The two-fluid model
features a more straightforward concept, and provides the capability
to model the water phase change and ionomer absorption/desorption processes at different rates; and the mixture model is more
computational efficient with less amount of conservation equations.
It should be noticed that the equilibrium approach described through
Equations (82)e(84) can also be implement to the two-fluid model by
modifying the corresponding effective porosities and diffusion
coefficients in the gas species conservation equations.
6.4. Boundary conditions and numerical implementation
The boundary conditions for the two models can be specified in
similar ways. At inlets of flow channels, mass flow rates (or velocity or
mass flux etc.), volume/mass fractions of the different gas species and
liquid water, and temperatures are often specified, and pressures are
usually defined at outlets of flow channel. On the outer surfaces of
computational domain, either temperatures or heat flux conditions
can be specified for heat transfer into or out of computational domain.
For the boundary conditions of electronic and ionic potentials, it
is worthwhile to be mentioned that two methods have been used, as
shown in Fig. 27. In method 1 in Fig. 27, the electronic potential at the
anode BP end surface is set at fele ¼ 0, and it is set to be the cell
voltage at the cathode BP end surface, fele ¼ Ecell. The overpotential
(activational voltage loss) in anode CL is simply the difference
between the electronic and ionic potentials, haact ¼ fele fion, while
the overpotential in cathode CL is calculated as hcact ¼ fele fion Er,
where Er (V) is the theoretical reversible cell voltage (as described in
Section 1 and Equation (4)). It can be calculated from the modified
form of the Nernst equation by assuming that the overall cell reaction is at thermodynamic equilibrium:
Vr ¼
Fig. 27. Two different methods in specifying boundary conditions for electronic and
ionic potentials [137].
Dgref
2F
þ
Dsref
2F
T
Tref
!
!0:5 #
"
pO2
p H2
RT
þ ln
2F
pref
pref
(85)
Here Dgref (J mol 1) and Dsref (J mol 1 K 1) are the changes of
Gibbs free energy and entropy for the overall reaction per mole of
hydrogen at reference temperature (Tref, K) and pressure (pref, Pa). F
(96,487 C mol 1) is the Faraday’s constant, R (8314 J kmol 1 K 1) the
universal gas constant, and T (K) the temperature. pH2 (Pa) and pO2
(Pa) are the partial pressures of hydrogen and oxygen, respectively,
and the values at flow channel inlets or the averaged values in CL
have all been used for calculating the reversible voltage. In method 2
in Fig. 27, a zero electronic potential is set at cathode BP end surface,
fele ¼ 0. While at anode BP end surface, the total cell potential loss is
imposed, fele ¼ htotal ¼ Er Ecell. The corresponding overpotentials
in anode and cathode CLs are all hact¼fele fion.
No observable differences in results between these methods in
Fig. 27 were found [137]. The sample distributions of potential/overpotential corresponding to the two methods are shown in Fig. 28.
Generally, the potential distribution from method 1 is physically more
meaningful since it demonstrates the real potential distributions
within the cell. In contrast, the potential distribution from method 2 is
more intuitive since it reveals in a straightforward manner of the
258
K. Jiao, X. Li / Progress in Energy and Combustion Science 37 (2011) 221e291
φele
φion
0.6
0.5
Po tential, V
0.4
0.3
ηc = φele − φ ion − Er
0.2
0.1
0
-2
a
0.7
Current D ens ity, A cm
a
0.94
Liquid production
Vapour production
0.92
0.9
Equivalent to
single phase model
0.88
Equivalent to
mixture model
0.86
ηa = φele − φ ion
0.84
-0.1
10
-3
-0.2
CL
Membrane
CL
-1
10
10
0
1
2
10
10
0.6
3
1.2
Production in ionomer
Liquid production
Current D ensity, A cm
-2
φele
φion
0.5
10
-1
Anode
b
Overpotential, V
-2
Water Phase Change Rate, s
Cathode
b
10
ηa = φele − φion
0.4
0.3
1.1
1
Equivalent to
equilibrium aproach
0.9
ηc = φele − φion
0.2
0.8
0.1
10
-2
10
-1
10
0
10
1
Water Phase Change Rate, s
0
Cathode
CL
Membrane
CL
Anode
Fig. 28. Sample distributions of potential/overpotential corresponding to boundary
conditions in Fig. 27 (a: corresponding to method 1 in Fig. 27; b: corresponding to
method 2 in Fig. 27) [137].
potential loss from each component of the cell. It was also found that
method 2 is more computational efficient than method 1 [137].
The model equations described for the two-fluid and mixture
models are often implemented into CFD codes (such as the
commercial code Fluent, Star-CD, CFX, CFD-ACEþ) based on finite
volume or finite element methods, through their user coding
capability. Therefore, the multi-dimensional geometry can be easily
handled, and the numerical iterations can be properly stabilized.
6.5. Two-fluid model vs mixture model
As described in Sections 6.1e6.4, the main difference between
the two-fluid and mixture models is that a separate conservation
equation of liquid water is solved in the two-fluid model, and the
liquid water distribution is estimated from the two-phase mixture
properties in the mixture model. The phase equilibrium is always
maintained in the mixture model (instantaneous water evaporation and condensation), and the non-equilibrium water phase
change can be accounted for by the two-fluid model by setting
different source terms in the conservation equations. Even though
the non-equilibrium ionomer absorption/desorption approach is
described with the two-fluid model, and the equilibrium ionomer
absorption/desorption approach is presented with the mixture
10
2
-1
Fig. 29. Effects of water phase change rate (a) between vapour and liquid water and (b)
between water in pores and water in ionomer [137].
model, it is important to be noticed that these two approaches can
all be applied to the two models.
By using different values of the water phase change rate, g (s 1),
in equations (54) and (60) for the two-fluid model, the effects of
mass transfer rate for water phase change and ionomer absorption/
desorption processes can be examined. Fig. 29 compares the effects
of water phase change rate between vapour and liquid water and
between water in pores and water in ionomer on the performance of
a single straight PEMFC running at a constant voltage. In Fig. 29a, the
product water is assumed to be vapour or liquid water corresponding to the two lines in this figure. When water is assumed to be
produced in vapour state, the current density decreases significantly
with the increment of water phase change rate, due to the fact that
the liquid water formation can be neglected at very low water phase
rate, therefore the water vapour production condition together with
very low water phase change rate can be considered to be equivalent
to a single-phase model (formation of liquid water is neglected). On
the other hand, when water is assumed to be produced in liquid
state, the current density is not very sensitive to the water phase
change rate. When the water phase change rate is high enough (on
the level of 103 s 1), the current densities corresponding to the two
water production mechanisms become similar, due to the fact that
the phase equilibrium state can be achieved with large water phase
rate. Therefore, the two-fluid model with large water phase change
rate between vapour and liquid water can be regarded as equivalent
to the mixture model (assumes phase equilibrium between vapour
and liquid water). Water is assumed to be produced in liquid state in
K. Jiao, X. Li / Progress in Energy and Combustion Science 37 (2011) 221e291
259
Fig. 30. Evolution of liquid water volume fraction upon the change of anode and cathode inlet relative humidities from 100% to 50% [252].
Fig. 31. Multi-channel simulation results (a: water content at membrane/CL interface in cathode; b: current density distribution in membrane; c: temperature distribution in
membrane) [238].
260
K. Jiao, X. Li / Progress in Energy and Combustion Science 37 (2011) 221e291
pores or absorbed in ionomer in Fig. 29b, it can be noticed that the
production water in ionomer mechanism is more sensitive to the
water phase rate for ionomer absorption/desorption than the liquid
water production mechanism. With large enough water phase
change rate (on the level of 102 s 1), both water production mechanism result in similar results, and they can be considered to be
equivalent to the equilibrium ionomer absorption/desorption
approach described in Section 6.3.
With the two-fluid and mixture models, transient simulations to
investigate liquid water evolutions upon change of operating
condition can be conducted. Fig. 30 shows the evolution of liquid
water volume fraction upon the change of anode and cathode inlet
relative humidities from 100% to 50% based on the mixture model
with equilibrium ionomer absorption/desorption [253]. By
reducing the inlet relative humidities, it can be observed that the
amount of liquid water decreases, and at the time instance of 1.75 s
after the relative humidity change, almost all the liquid in electrodes under flow channel is removed, but it is still present under
the lands. Large scale simulations considering multi-channels or
stacks can also be carried out based on such continuum (top-down)
approach by assuming homogeneous materials and effective
transport properties. Fig. 31 shows the multi-channel simulation
results with inlet gas humidities all at 50%. (liquid water formation
was neglected in the simulations), which are the water content at
membrane/CL interface in cathode and current density and
temperature distributions in membrane. It can be observed from
Fig. 31a that the water content at membrane/CL interface in
cathode increases along the flow direction mainly due to the
product water accumulation. Correspondingly, the current density
distribution in Fig. 31b matches well with Fig. 31a: the current
density is higher at the locations with higher water content (less
ohmic resistance). With higher current density, the reaction rate is
also higher, resulting in faster heat generation, therefore the
temperature distribution in Fig. 31c matches with the water
content and current density distributions in Fig. 31a and b.
6.6. Summary
The continuum (top-down) full cell models for PEMFC by
considering homogeneous materials of GDL, CL and membrane
with modeled effective transport properties are reviewed in
Section 6. Such models can be generally classified into two-fluid
model and mixture model. The phase equilibrium is always maintained in the mixture model, and the non-equilibrium water phase
change can be accounted for by the two-fluid model by setting
different phase change rates. Such multi-dimensional multicomponent multiphase full cell models are powerful tools to
investigate the transport phenomena simultaneously in different
cell components and to evaluate various cell designs and material
properties for the cell sizes from single straight channel to multichannel or stacks based on the available computational power.
Therefore, such models have attracted many attentions. However,
modeling a specific cell component with more realistic microstructures and more accurate formulations provides deeper
insights for various transport phenomena and material properties.
Therefore, in the following Sections (Sections 7e10), various
models applied to specific cell components are introduced.
7. Modeling water transport in membrane
Both the top-down and bottom-up models have been extensively carried out for modeling water and proton transport in
membrane. As mentioned previously, the top-down approaches
mainly include diffusive, chemical potential and hydraulic models.
The bottom-up models were used to learn water and proton
transport mechanisms and effects of water on ionomer self-organization. Physical models were also developed, and simplified
structures of hydrophilic pores in ionomer at different hydration
levels were proposed in such models to provide theoretical basis for
explaining the water transport mechanism in ionomer.
7.1. Macroscopic approach
7.1.1. Diffusive model
The diffusive model can be explained with dilute solution theory
[311], by considering ionomer as solvent and water and proton as
solute species. The dilute solution theory assumes that the solute
species are dilute enough that the interaction between them can be
neglected, and only the interaction between solute and solvent is
accounted for. The flux of solute species (water and proton) in
solvent (ionomer) can be expressed by using the NernstePlanck
equation [311]:
Ji ¼
zi wi Fi Vfion
!
Di Vci þ ci u mem
(86)
Here the subscript i represents different solute species (water
and proton). Ji (kmol m 3 s 1) is the superficial flux, zi the charge
number or valence, wi (m2 kmol J 1 s 1) the mobility, F
(9.6487 107 C kmol 1) the Faraday’s constant, fion (V) the ionic
potential, Di (m2 s 1) the diffusion coefficient, ci (kmol m 3) the
!
concentration, and u mem (m s 1) the superficial velocity of solvent
(the ionomer). The relationship between wi (m2 kmol J 1 s 1) and Di
(m2 s 1) is given by the NernsteEinstein equation [311e313]:
(87)
Di ¼ RTwi
1
1
where R (8314 J kmol K ) is the universal gas constant, and T (K)
the temperature.
For water transport, the first term on the right hand side of
Equation (86) becomes zero because water has no valence. For proton
transport, as mentioned earlier, the diffusive model assumes constant
concentration of proton in ionomer, therefore the second term on the
right hand side of Equation (86) becomes 0. Further, due the fact that
!
the ionomer does not move ð u mem ¼ 0Þ, therefore the last term on
the right hand side of Equation (86) becomes 0 for both water and
proton transport. As a result, the simplified NernstePlanck equation
(Equation (86)) for proton transport becomes Ohm’s law:
Iion ¼
(88)
kion Vfion
2
1
where Iion (A m ) and kion (S m ) are the ionic current density and
conductivity, respectively. The simplified NernstePlanck equation
(Equation (86)) for water transport becomes Fick’s law, and by
further considering the EOD effect (Equation (18)), The water flux
(Jnmw, kmol m 2 s 1) becomes
Jnmw ¼
Dnmw Vcnmw þ nd
Iion
F
(89)
Here the subscript, nmw, represents the non-frozen water in
ionomer (the only water in ionomer in normal operating condition). nd is the EOD coefficient, and F (9.6487 107 C kmol 1) the
Faraday’s constant.
It can be noticed that the diffusive model described in this
subsection is actually used in the two-fluid and mixture models
described in Section 6. In fact, the diffusive model has become the
most successful top-down model for water transport in ionomer for
PEMFC modeling since its initial application by Springer et al.
[72,204].
7.1.2. Chemical potential model
In the chemical potential model, the interactions between water,
proton and ionomer are all accounted for. This can be achieved by
utilizing the concentrated solution theory [75,314], generalized
261
K. Jiao, X. Li / Progress in Energy and Combustion Science 37 (2011) 221e291
StefaneMaxwell equations [305e307,315], dusty fluid model
[149,303,316], or the irreversible thermodynamics [317,318]. With
the consideration of multi-component (water, proton and ionomer)
interactions, these approaches give the fluxes in the following
forms (or similar expressions):
Iion ¼
Jnmw ¼
nd kion
VPion
F
I
anmw VPnmw þ nd ion
F
kion Vfion
due to EOD effect. The model in [307] sheds light on full cell
modeling incorporating the chemical potential model for water
transport in ionomer, and with further exploration of the transport
parameters, the chemical potential model is expected to be the
dominating top-down model replacing the diffusive model for
water transport in ionomer.
(90)
(91)
where Pion (J kmol 1), anmw (kmol2 J 1 m 1 s 1) and Pnmw
(J kmol 1) are the chemical potential of hydrogen ion (proton),
water transport coefficient, and chemical potential of non-frozen
water, respectively. The major difficulty for the chemical potential model is the lack of known transport parameters, such as
Pion, anmw and Pnmw. It can be noticed that by replacing the
chemical potential (P) by concentration (c), and by replacing the
water transfer coefficient (a) by diffusion coefficient (D), Equations (89) and (91) become identical, and one more term
considering the multi-component interaction in Equation (90)
presents than in Equation (88). Such replacement has been
used as one way to estimate the unknown transport parameters,
(e.g. [204,319e321]), and another way is to estimate the transport parameters based on available experimental data (e.g.
[304,305,307,322]).
The difficulty in obtaining the transport parameters makes the
chemical potential model less popular than the diffusive model,
and the chemical potential model was also relatively rarely applied
for full cell modeling. It is worthwhile to be mentioned that
Baschuk and Li [307] have recently applied the chemical potential
model to full cell modeling based on StefaneMaxwell equations. In
this model, the membrane, CL, GDL and flow channel were all
included, and protons were assumed be bound to water molecules
to form hydroniums, therefore the transport of hydronium is
accounted for rather than proton, (similar approach can also be
found in [149,305]). The multi-component transport is systematically accounted for in the whole computational domain in [307].
Fig. 32 shows the simulated water activity in different cell
components with fully humidified inlet gases on both sides [307],
the water activity in membrane and CL can be transferred to water
content by using Equation (12). It can be observed from Fig. 32 that
water accumulates at the downstream of cathode, and the anode
dries out along the flow direction. The highest water activity is
observed in cathode CL close to cathode outlet due to water
production and accumulation, and the lowest water activity is
observed at the membrane/CL interface in anode near flow outlet
7.1.3. Hydraulic model
In the diffusive and chemical potential models described in
Sections 7.1.1 and 7.1.2, the convective transport caused by pressure
gradient across membrane is not accounted for. As described in
Section 3, water enlarges the pores in ionomer to allow larger amount
of convective transport. The hydraulic model originates from the
models developed by Bernardi and Verbrugge [109,205]. In [109,205],
the membrane was assumed to be fully hydrated allowing the
maximum possible convective transport. The dilute solution theory
was utilized based on the NernstePlanck equation, and the Schlogl’s
equation [323,324] has been used to calculate the water flux due to
pressure gradient and EOD effect. The models in [109,205] also
assumed constant gas volume fractions in membrane, allowing
convective gas transport as well. After that, more models utilizing the
approach in [109,205] have been developed [74,213,229,325e327].
Generally, the hydraulic model neglects the diffusive transport and
accounts for the convective transport water in ionomer, and the water
flux can be represented by the follow equation
Jnmw ¼
Knmw
mnmw
Vpnwm þ nd
Iion
F
(92)
where Knmw (m2) is the permeability of non-frozen water in ionomer, mnmw (kg m 1 s 1) the dynamic viscosity of non-frozen
water in ionomer (the property of liquid water is often used
instead), and pnmw (Pa) the pressure of non-frozen water in
ionomer.
It should be noticed that the membrane close to anode side
often dries out quickly due to EOD effect during PEMFC operation.
Therefore, the fully hydrated membrane assumption remains
questionable. In fact, the water flux due to pressure gradient is only
considerable when the cell is differentially pressurized or most of
the membrane is well hydrated, and most models only accounted
for the diffusive and EOD caused water transport. Therefore, the
hydraulic model neglecting the diffusive water transport essentially
reflects unrealistic conditions.
7.1.4. Combinational model
When a PEMFC is differentially pressurized for anode and
cathode, both the diffusive and convective water transport may need
to be accounted for, and the most straightforward way is to add all
the terms (diffusive, convective and EOD) together, forming the
combinational diffusive/hydraulic model [211,255,303,328e332]
Jnmw ¼
Fig. 32. Simulated water activity in different cell components with fully humidified
inlet gases on both sides [307].
cnmw
Dnmw Vcnmw
cnmw
Knmw
mnmw
Vpnwm þ nd
Iion
F
(93)
It should be noticed that the permeability of non-frozen
water in ionomer, Knmw, increases with the increment of water
content, as shown in Equation (22) and other experimental
measurements (such as [333]). Another approach to account for
both the diffusive and convective water transport mainly
focused on modifying the chemical potential model with the
pressure effect [304,322], and the effects of convective water
transport on proton transport can be accounted for in the
modified chemical potential model. However, due to the fact that
some transport parameters remain unclear, further exploration
is needed for this approach.
262
K. Jiao, X. Li / Progress in Energy and Combustion Science 37 (2011) 221e291
Fig. 33. Snapshots of final micro-structures in hydrated Nafion membrane at different water contents (water, hydronium and side chain domains are show in green, and
hydrophobic domains are shown in red) [346]. (For interpretation of the references to colour in this figure legend, the reader is referred to the web version of this article.)
7.2. Bottom-up approach and physical models
7.2.1. Modeling ionomer self-organization
As mentioned in Sections 2 and 3, with the changes of amount and
state of water in PFSA polymer membranes, the ionomer morphology
changes as well, resulting in different water and proton transport
characteristics. Even though different simplified micro-structures of
ionomer were proposed (detailed in Section 7.3), bottom-up molecular simulations are the only computational methods to investigate
the details of chemical structure. Therefore, quantum mechanical
[78,224,280,281,334] and classical [282e290,335] MD and pseudo
particle [291,292,336e340] simulations have been extensively
carried out for this purpose.
The MD simulations in [224,334] showed the formation of
percolating channels with water in ionomer, which agrees with the
simplified micro-structure of Nafion ionomer proposed in [341,342].
However, the percolating water channel was observed breaking very
fast after formation in the MD simulations in [282]. Both the water
being strongly and loosely bound to sulphuric sites were observed in
the MD simulations in [224,335], and frequent exchange between
the two kinds of water at high water contents was also presented in
[335]. Similarly, other MD simulations (e.g. [285,287e290]) were
also performed to study the ionomer morphology and state of water.
The MD simulations on atomistic level mentioned in the previous
paragraph are useful tools for investigating the water phase segregation in ionomer. The current computational power does not allow
such simulations to represent the chemical structures in a sufficiently large length scale and long time scale. Pseudo particle
simulations are therefore needed by representing a group of molecules as a single molecule with appropriate approximations to save
computational time. Coarse grained (CG) MD (CGMD) simulations
were conducted in [291,294] by presenting a group of atoms as a CG
segment, and it was reported that the shape of hydrophilic regions
changes from spherical to elliptical when the water content
increases from 6 to 8. MC simulations by using a reference interaction site model [343] were carried out in [336]. In [336], each CF2 and
CF3 were presented by a pseudo atom. Threes layers were observed
for Nafion ionomer, which are a central water-rich region and two
outer layers of side groups strongly associated with water molecules.
A linear relationship between ionomer swelling and water content
was also reported in [336]. Dissipative particle dynamics (DPD) was
also used to study the chemical structure of Nafion ionomer
[337e339] by considering the motion of pseudo particles governed
by Newton’s equations [344,345]. It was found that the hydrophilic
domain size increases linearly with water content [339]. LB simulations based on the morphologies generated in [339] were also
performed to simulate water transport in ionomer, and it was
confirmed that the permeability of water in ionomer increases with
water content [340].
Fig. 33 shows the snapshots of the final micro-structures in
hydrated Nafion membrane at different water contents obtained
from CGMD simulations [346]. It can be observed that the hydrophilic domains (water, hydronium and polymer side chain) form
clusters embedded in the hydrophobic domains (hydrophobic
backbone). For the simulations in Fig. 33, it was also reported that
the typical water filled channel sizes are 1, 2 and 4 nm at the water
contents of 4, 9 and 15, respectively.
7.2.2. Modeling proton transport
ab initio quantum mechanical MD simulations are usually
needed to study the mechanism of proton transport, however, the
current computational power does not allow to perform full MD
simulations of proton/water transport in the real structure of
Nafion membrane [346]. The structural complexity needs to be
simplified, and the rarity of proton transfer events (need large time
scale) needs to be efficiently sampled [346e351]. MD simulations
to investigate the proton transfer between water and charged sites
were conducted in [99,281,352] by considering several polymer
side chains attached to a single hydrophobic backbone, therefore
the detailed polymerewater interaction could be studied. The MD
simulations in [353] found that the flexibility of polymer side chain
could be important for proton transport at low hydration level.
Similarly, Fig. 34 shows the elementary interfacial proton transport
and configuration energy along the reaction path obtained from
MD simulations at low hydration level [346], and it can be noticed
that tilting and rotation of the acceptor/donor sites facilitate the
elementary interfacial proton transport.
7.2.3. Physical models with simplified membrane micro-structure
Based on experimental observations (such as X-ray, NMR and IR)
and percolation theory, Hsu and Gierke [341,342] postulated
a simplified membrane micro-structure, as shown in Fig. 35, a principal spherical and minor cylinder-like hydrophilic micro-structure
was assumed. Even though the proposed micro-structure remains
debated, it has become the most enduring Nafion morphology. This
micro-structure in three-dimension suggested a percolation
threshold of 15% (about 3.3 water content) for the hopping proton
transport descried in Section 3.1 (the minimum water content to
form a continuous water pathway in ionomer), which reasonably
agrees with experimental measurement (proton conductivity
becomes appreciable when water content ranges from 2 to 5).
Dissimilar to the spherical-cylinder micro-structure proposed in
[341,342], irregular three-phase [354], lamella [355], sandwiched,
[356] and channel-like [357] micro-structures were also proposed.
K. Jiao, X. Li / Progress in Energy and Combustion Science 37 (2011) 221e291
263
Fig. 34. Elementary interfacial proton transfer and configuration energy along the reaction path (the nearest neighbour distance of C atoms, dCC, is about 0.7 nm, and the activation
barrier is about 0.5 eV) [346].
Similarly, all these models provided low percolation thresholds as
the critical feature of proton conductivity in ionomer. Based on the
simplified membrane micro-structure in [341,342], a PN model was
also developed [73]. This model assumed that the hydrophilic
clusters are spherical pores connected by cylindrical channels. The
numbers and sizes of the pores and channels are estimated based on
water content, and therefore the relationship between proton
conductivity and water content was obtained, and a percolation
threshold of 20% was suggested (about 4.4 water content).
difficulty in obtaining the transport parameters makes the chemical
potential model less popular than the diffusive model, and the
chemical potential model was also relatively rarely applied for full
cell modeling. The modified diffusive model by adding an EOD term
is still the most popular model so far. The bottom-up models have
been developed to study water and proton transport mechanisms
and effects of water on ionomer self-organization. Simplified
structures of hydrophilic pores in ionomer at different hydration
levels have also been proposed to provide theoretical basis for
explaining the water transport mechanism in ionomer.
7.3. Summary
The different models for water transport in membrane are
reviewed in Section 7. The review shows that both the top-down
and bottom-up models have been extensively developed. The topdown models can be classified into diffusive, chemical potential
and hydraulic models. Even though the chemical potential model
solves the water transport in the most comprehensive way, the
5 nm
SO3
−
SO3
−
SO3−
SO3−
SO3−
SO3− SO3−
SO3−
SO3−
SO3−
4 nm
SO3
1 nm
SO3−
−
SO3−
SO3− SO3−
SO3−
SO3−
SO3−
SO3−
SO3−
Fig. 35. Simplified Nafion micro-structure [341,342].
Fig. 36. A magnified picture of GDL (carbon paper) [268].
264
K. Jiao, X. Li / Progress in Energy and Combustion Science 37 (2011) 221e291
8. Modeling water transport in gas diffusion layer
Many different models have been developed to simulate transport phenomena in GDL. Generally, the different models can be
categorized into two groups, one assumes homogeneous material
of GDL, and another utilizes real or simplified micro-structure of
GDL for simulations. When GDL is assumed to be homogenous, the
micro-structure is neglected and the computational domain
becomes pure void. Effective transport properties are used to reflect
the effects of the neglected micro-structure. Such homogeneous
approach greatly simplifies the geometry generation process and
results in more stable numerical iterations. The full cell models
described in Section 6 all use this approach. The digital microstructure of GDL needs to be generated first for the second
approach. The small pores result in extremely large number of
computational grid, therefore usually only a small sample was
accounted for. This approach therefore was mostly adopted to
simulate liquid water behaviours or to numerically measure the
transport properties. In this section, the different models considering homogeneous or real/simplified GDL are introduced, the
homogeneous approach is only briefly discussed because it is
described in the previous sections, and this section mainly focuses
on the models considering real or simplified micro-structures of
GDL. In addition, both the first-principle-based and rule-based
models are included in this section.
8.1. Homogeneous approach
The homogeneous approach is described with the full cell
models in Section 6, and the various transport mechanisms are
described Sections 2 and 3, therefore the details of this approach
are not repeated here. It should be noticed that the StefaneMaxwell
formulation for gas diffusion is not used in the full cell models in
Section 6. Such approach is only valid when two gas species are
present. For PEMFC without impurities (e.g. CO), the only gases in
anode are hydrogen and water vapour, which is acceptable;
however, oxygen, water vapour and nitrogen are all present in gas
phase in cathode when air is supplied, which requires StefaneMaxwell formulation, as shown in the following equation.
VYi ¼
X Yi Jj
jsi
Yj Ji
rg Deff
i;j
used. The micro-structure can be obtained either by a threedimensional volume imaging technique or by constructing a digital
micro-structure based on stochastic models, description about such
models can be found in [358,359]. Constructing the three-dimensional image using the volume imaging technique requires the use of
non-invasive experimental techniques, such as X-ray and NMR.
With this technique, the porous material is repeatedly sectioned and
imaged automatically. Then the two-dimensional cross-sectional
images are combined to construct the three-dimensional image of
the micro-structure. This technique can be very expensive and time
consuming. It requires the use of a three-dimensional imaging
software along with a scanning electron microscope and very
accurate sectioning of the GDL. The reconstruction of the image
using digital stochastic models requires the knowledge about the
pore distribution and pore size of the micro-structure, which can be
obtained easily with a porosimeter. This technique is not as expensive and is faster than the experimental imaging technique, and has
been used and proved to be successful (e.g. [274,275,278,360]).
Assumptions are often needed for the stochastic method, for
example, for constructing the micro-structure of carbon paper, the
following assumptions are often made [360]:
1. The fibres are considered to be cylindrical, with a constant
radius and are infinitely long.
2. The fibres are allowed to overlap.
3. According to the fabrication process of carbon fibre, the fibre
system is isotropic in the material plane.
Fig. 37 shows a three-dimensional computational domain for
GDL (carbon paper) with a 90% porosity obtained by using the
stochastic method [360], which is very comparable with the image
shown in Fig. 36.
Note that simplified micro-structures of the GDL were also used,
such as the study in [121] assumed that the GDL is made up of
stacked two-dimensional random carbon fibre mats as shown in
Fig. 38. It was assumed that the fibres are infinitely long on the plane
parallel to GDL surface, and are allowed to overlap. The solid
(94)
Here the subscripts i and j represent different gas species. Y and J
(kg m 2 s 1) are the mass fraction of gas species in gas mixture and
diffusional flux. rg (kg m 3) is the density of gas mixture, and Di,jeff
(m2 s 1) the effective binary diffusion coefficient of species i in
species j. Apparently, the StefaneMaxwell formulation is much
more complex than the Fickian diffusion formulation in Section 6.
Therefore, many of the previous full cell models neglected it to only
consider the Fickian diffusion formulation.
As reviewed in Section 5, the VOF and LB are the only two firstprinciple-based models used that can track the interface between
liquid and gas phases to investigate the detailed liquid water
behaviour in PEMFC. Even though these models can all be used for
the homogeneous approach with effective transport properties, the
detailed liquid water behaviour can only be investigated when the
micro-structure of GDL is used as the computational domain.
8.2. Structure generation
A magnified picture of GDL (carbon paper) is shown in Fig. 36
[268]. It can be noticed that the GDL structure is highly anomalous,
and therefore special methods are needed to generate the digital
micro-structure of GDL. There are two common methods that can be
Fig. 37. Three-dimensional computational domain for GDL (carbon paper) with a 90%
porosity obtained by using stochastic method [360].
265
K. Jiao, X. Li / Progress in Energy and Combustion Science 37 (2011) 221e291
Fig. 38. Simplified micro-structure of GDL (a: image of Toray carbon paper; b: single screen made of overlapping fibres with square pore spaces; c: stack of fibre screens; d: pore
spaces of stacks with an arbitrary screen position shifting) [121].
structure was modeled as stacks of continuously overlapping fibre
screens. PN models often utilizes simplified micro-structures which
consist of pores, throats and solids, as shown in Fig. 39 [361]. The
flow resistances are modeled to be different in pores and throats in
PN models with such simplified micro-structures (detailed in
Section 8.5.2). The simplified micro-structures shown in Figs. 38 and
39 are much easier to be constructed than the more realistic structure in Fig. 37, and therefore were also widely adopted.
8.3. Volume-of-fluid model
The VOF model is perhaps the only top-down model that has
been used to investigate detailed liquid water behaviour in PEMFC
[11,12,262e268]. Most of the work focused on liquid water
transport in flow channel or in GDL with simplified micro-structure. With the micrometre pores in GDL, it is still possible to use the
top-down VOF model with real GDL micro-structures. To the best of
the authors’ knowledge, only Park et al. [268] recently adopted the
VOF method to simulate water transport in a real GDL.
By neglecting the phase change, electrochemical reaction, gravity
and heat transfer, the VOF model with two phases, gas and liquid
water, can be presented by the following conservation equations.
Mass of two-phase mixture:
v
!
ðrÞ þ V$ðr u Þ ¼ 0
vt
(95)
Fig. 39. Two-dimensional schematic of pore-network construction (a: relationship among pores, throats and solid; b: structure in terms of void and solid space) [361].
266
K. Jiao, X. Li / Progress in Energy and Combustion Science 37 (2011) 221e291
Momentum of two-phase mixture:
v !
!!
ðr u Þ þ V$ðr u u Þ ¼
vt
Liquid water:
T
!
!
Vp þ mV$ Vð u Þ þ V u
2
mVðV$ð!
u ÞÞ þ Su
3
(96)
v
rslq þ V$ rslq !
u ¼ 0
vt
(97)
r ¼ sg rg þ slq rlq
(98)
where r (kg m 3) and m (kg m 1 s 1) are the density and dynamic
viscosity of the two-phase mixture, respectively, and they are the
volume averaged values, such as for density:
Here the subscripts g and lq represent the gas and liquid phases,
and s the volume fraction. It can be noticed that the two phases rely
!
on the same momentum equations, and therefore the velocity ( u ,
1
m s ) and pressure (p, Pa) are shared by the two phases in each
computational cell as well. The liquid water conservation is solved
(Equation (97)), and the gas phase volume fraction can be simply
calculated since the total volume fraction is unity.
sg þ slq ¼ 1
(99)
2
2
The source term (Su, kg m s ) in the momentum equation
(Equation (96)) accounts for the surface tension effect, and it can be
calculated as
Su ¼ s
rkVsg
0:5 rg þ rlq
(100)
where s (N m 1) is the surface tension coefficient between the two
phases, and k (m 1) the surface curvature defined as
!
!
k ¼ V$!
n ¼ V$ð n w cosðqÞ þ t w sinðqÞÞ
(101)
!
where n is the unit vector normal to the interface between the two
!
!
phases; near the wall, n w and t w are the unit vectors normal and
tangential to the wall surface; and q is the contact angle. It can be
noticed that the shape of the interface between the two phases at
the wall depends on the wettability of the wall. In addition, it can be
noticed from Equations (95)e(97) that the porous media properties
such as porosity are not taken into account (as described in Section
6 for the two-fluid and mixture models), this is valid in flow
channel (pure void space) and when the micro-structure of GDL (or
CL) is represented in the computational domain.
By using the VOF model in a GDL micro-structure, Fig. 40 shows
the transient liquid water discharging from GDL with a pressure
gradient of 6.5 105 Pa m 1 and a contact angle of 135 [268]. It can
be observed that when a pressure gradient is exposed, liquid water
are being removed, and finally there is a small amount of liquid
water left which is difficult to be removed. Fig. 41 presents the
transient liquid water volume fractions in GDL with different
pressure gradients and contact angles [268]. It can be noticed that
the water removal time is significantly affected by the amount of
pressure gradient, and the GDL hydrophobicity has to be high
enough to initiate liquid water removal. Not only investigating
liquid water dynamics, the simulations in [268] also indicated that
the pressure gradient and contact angle are the two parameters
which determine the initiation of liquid water removal in GDL.
Fig. 40. Transient liquid water discharging from GDL with a pressure gradient of
6.5 105 Pa m 1 and a contact angle of 135 obtained from volume-of-fluid model (the
time step between the successive sequence plots is 500 ms) [268].
adopted to find the GDL and CL properties (e.g. permeability)
[272,273], as well as to investigate liquid water transport in GDL
and CL [274e277].
The LB method is based on a finite number of identical particles
(it is intrinsically a pseudo particle method) that go through collision and propagation successively on prefixed paths in space. For
multiphase flow simulations, the evolution of distribution function
for each phase is represented by an evolution equation [276]. For
the kth phase, it is
8.4. Lattice Boltzmann model
Based on the numerously previous development [362e373], the
LB model, originated from its predecessor LG model, has been
fik ðx þ ei dt ; t þ dt Þ
fik ðx; tÞ ¼
fik ðx; tÞ
kðeqÞ
fi
sk
ðx; tÞ
(102)
K. Jiao, X. Li / Progress in Energy and Combustion Science 37 (2011) 221e291
Liquid Water V o lume Fraction
a
1
6.5×103 Pa m-1
4
-1
6.5×10 Pa m
6.5×105 Pa m-1
6.5×106 Pa m-1
0.8
0.6
0.4
0.2
0
0
1
2
3
4
-3
Time, 10 s
Liquid Water V o lume Fraction
b
1
124 o
125 o
127.5 o
135 o
0.8
0.6
0.4
0.2
0
1
2
3
4
-3
Time, 10 s
Fig. 41. Transient liquid water volume fractions in GDL (a: for different pressure
gradients when the contact angle is 135 ; b: for different contact angles when the
pressure gradient is 6.5 105 Pa m 1) [268].
where fik(x,t) is the number density distribution function for the
kth phase in the ith velocity direction at position x and time t, and dt
is the time step. sk is the relaxation time of the kth phase, and
fik(eq)(x,t) represents the equilibrium distribution function. The
(2)
(5)
(0)
(3)
(7)
right hand side of Equation (102) is the BhatnagareGrosseKrook
(BGK) collision term [362]. The LB model is usually represented by
the numbers of dimension and velocity directions. Fig. 42 shows
the lattice structure of a D2Q9 lattice Boltzmann model [276]. Here
D2 represents two-dimensional, and Q9 means 9 directions of
velocity at each node. Other configurations such as D3Q15 and
D3Q19 were also commonly used. The interactions between fluids
and between fluids and wall (e.g. surface tension and wall adhesion) can be taken into account through the modified equilibrium
distribution functions. The macroscopic fluid phase properties can
be obtained through appropriate averaging of the particle distribution function. It should be noticed that the density ratio between
air and liquid water is about 1000 in cathode, while it is much
higher in anode (hydrogen and liquid water). However, most of the
previous LB simulations assumed that the density ratio is 1 (large
density ratio results in numerical instability), which is only valid
when the capillary effect dominates. As analyzed in Section 3.2.2,
this assumption is valid when the cross flow through GDL is not
significant (e.g. with parallel flow channel design small pressure
gradient is present in GDL), and it may be failed for large cross flow
(e.g. with long serpentine or interdigitated flow channel design
large pressure gradient is present in GDL). Even though a few
recently developed LB models [370,373] are available for high
density ratio up to 1000, their stability is not yet proven in
a complicated geometry.
Fig. 43 shows the liquid water transport through GDL obtained
from the LB simulation in [276], it can be noticed that part of water
is trapped in GDL and difficult to move, and this observation is
consistent with the water transport shown in Fig. 40 with VOF
model. It was also concluded in [276] that the pressure gradient and
GDL wettability are the dominating factors for liquid transport in
GDL, which agrees with the conclusion in [268] using VOF model.
8.5. Rule-based model
0
(6)
267
(4)
(1)
(8)
Fig. 42. Lattice structure of a two-dimensional (D2Q9) LB model [276].
The VOF and LB models introduced in the previous subsections
are first-principle-based models. Rule-based models were also used
to investigate liquid water transport in PEMFC. The most representing rule-based models are the FM [274,275,278] and PN
[274,279,361] models. Rather than solving a set of governing
equations, the rule-based models depend on applying physical
rules to simplified or real physical structures, and therefore are
more computationally efficient than the first-principle-based
models but can only be applied to certain flow conditions.
8.5.1. Full morphology model
Based on the previous development in [374,375], the FM model
has been adopted to investigate liquid water transport in GDL in
[274,275,278]. With a constructed GDL micro-structure (as
described in Section 8.2 and shown in Fig. 37), the FM model was
used to estimate the steady state liquid water distribution in GDL by
assuming that the capillary effect dominates the two-phase transport in GDL [274,275,278]. As analyzed in Section 3.2.2, this
assumption is valid when the cross flow through GDL is not
significant (e.g. with parallel flow channel design small pressure
gradient is present in GDL), and it may be unacceptable with strong
cross flow (e.g. with long serpentine or interdigitated flow channel
design large pressure gradient is present in GDL). Based on this
assumption, if liquid water enters a dry GDL from one side, the
pressure difference between the liquid water inlet and the opposite
GDL surface (the outlet) can be considered as the capillary pressure.
By further assuming that liquid water forms sphere droplets, the
capillary pressure, liquid water droplet diameter and contact angle
can be related by Equation (42). For different capillary pressures
and contact angles, different liquid water droplet diameters can be
268
K. Jiao, X. Li / Progress in Energy and Combustion Science 37 (2011) 221e291
or generalized Poiseulle law [376] or something similar. To investigate liquid water transport through GDL, the PN model in
[274,279] applied constant liquid water injection rate at one side of
a simplified network structure, constant pressure is applied to the
opposite surface (outlet), and the other boundaries are set as walls.
Liquid water can only invades to the next pore or throat when the
pressure difference between liquid water and gas is higher than the
capillary pressure. The pores and throats are invaded one by one
from liquid water inlet until a steady state liquid water distribution
is reached. Therefore the liquid water invasion process can be
investigated by using the PN model, and the liquid water distribution in GDL obtained from the PN model when a steady state is
reached is shown in Fig. 45 [279]. The PN model can account for
different flow resistances with proper calculation of pressure drop
(such as using the HagenePoiseulle law [144] or generalized Poiseulle law [376]), and therefore can be used for different flow
conditions. The transient simulation and capability for various flow
conditions are the advantages of the PN model by comparing with
the FM model. However, the PN model in [274,279] assumed that
the pores and throats are completely filled with liquid water after
invasion. This assumption remains debated because gas films may
still form at the corners between pores and throats [377,378]. The
simplified GDL micro-structure may also lead to unrealistic results.
These features make the accuracy of the PN model remain debated.
8.6. Summary
Fig. 43. Liquid water transport through GDL obtained from LB model (the Reynolds
and Capillary numbers are 0.1 and 2.3 10 5, respectively) [276].
obtained. The liquid water droplets with calculated diameters are
fitted in a generated GDL micro-structure (such as the one shown in
Fig. 37), and those droplets intersecting the solid material and not
directly connecting to the liquid water inlet are removed. The rest
of the liquid water are then dilated to fill the void space between
the connected droplets, and the volume fraction of the dilated
liquid water in the void space of the micro-structure is obtained
as the liquid water volume fraction corresponding to the capillary
pressure and contact angle. By using the FM model, the relationship between the capillary pressure and liquid water volume
fraction was studied by using different GDL samples, and
reasonable results were obtained [274,275,278]. Fig. 44 shows
the visualization of liquid water distribution in GDL obtained
from FM model [278]. Fig. 44a presents the micro-structure of
GDL without any liquid water, and Fig. 44bef shows that the
liquid water volume fraction increases with the pressure difference between the two phases. Generally, the FM model presented
here is more computationally efficient than the first-principlebased models. However, it can only be applied to certain flow
conditions, such as the capillary dominated flow discussed in this
subsection. Since only the steady state liquid water distribution
can be obtained from the FM model, the liquid water dynamics
cannot be investigated.
8.5.2. Pore-network model
In the PN models [274,279,361] developed for GDL, the GDL
micro-structures were assumed to consist of three components,
which are pores, throats and solid (such as the one shown in
Fig. 39). The shapes of the pores and throats are assumed to be
simple so that the flow resistance through the pores and throats
can be simply calculated. The pressure drop across the pores and
throats can be calculated based on the HagenePoiseulle law [144]
The different models for water transport in GDL are reviewed in
Section 8. This review shows that such models can be classified into
two groups, one assumes homogeneous material of GDL, and
another utilizes real or simplified micro-structure of GDL for
simulations. The homogenous approach is often adopted in full cell
modeling (Section 6). The second approach requires digital microstructure of GDL as the computational domain. Such computational
domains consist of small pores therefore result in much more
computational grids than the homogeneous approach. Therefore,
usually only small samples of GDL are considered in this approach.
The second approach involves both the first-principle-based and
rule-based models. The VOF and LB models are the two most
popular first-principle-based models for simulating two-phase
flow in GDL, and the FM and PN models are the two mostly widely
used rule-based models.
9. Modeling water transport in catalyst layer
Since the detailed formulations for the top-down models
considering homogeneous materials are described in Section 6, and
the various models with porous micro-structures are presented in
Section 8, they are not repeated in this section. It should be noticed
that the pores in CL are much smaller than in GDL (refer to the
analysis in Section 3.3), therefore the VOF model presented in
Section 8.3 cannot be applied to CL micro-structures, and the
bottom-up approach such as the LB model is needed to simulate
liquid water transport in CL pores. Because the calculations for the
rule-based models (the FM and PN models) are based on top-down
formulation, they may not be able to be directly applied to CL
micro-structures as well. Rather than just accounting for the StefaneMaxwell diffusion in the pores of GDL, the much smaller pores
in CL also results in Knudsen diffusion (as described in Section 3.3),
therefore the gas diffusion function given in Equation (94) for GDL
needs to be modified as
VYi ¼
X Yi Jj
jsi
Yj Ji
rg Deff
i;j
Ji
rg Deff
K;i
(103)
K. Jiao, X. Li / Progress in Energy and Combustion Science 37 (2011) 221e291
269
Fig. 44. Visualization of liquid water distribution in GDL obtained from FM model (a: GDL micro-structure without any liquid water; bef: steady state liquid water distribution in
GDL corresponding to different capillary pressures) [278].
2
1
where Deff
K,i (m s ) is the effective Knudsen diffusion coefficient of
gas species i. In this section, only the models not previously
included are described. The agglomerate model, with simplified CL
micro-structure and based on top-down formulation, is presented
first, followed by the MD and off-lattice pseudo particle simulations
for investigating the reaction mechanisms and material microstructures.
9.1. Agglomerate model
As mentioned previously, the top-down approach cannot be
applied to the small pores of CL. However, by assuming that the
catalyst, ionomer and part of void space are homogeneously mixed
and form micrometre agglomerates, the top-down approach can
still be used. Due to the fact that the sluggish electrochemical
reaction in cathode CL is the limiting factor of cell performance,
such agglomerate models usually focused on cathode CL. The
agglomerate model has been extensively studied in the past decade
[214e218,269e271,379e381], and different simplified microstructures were proposed such as spherical agglomerate [271],
cylindrical agglomerate [379], ordered CL [380], and non-uniform
CL [269]. Most of the previous agglomerate models are one- or twodimensional, and a three-dimensional agglomerate model was
developed recently to fully account for the effect of agglomerate
arrangement on cell performance [271].
By assuming spherical agglomerates with a diameter of 5 mm,
Fig. 46 shows the schematics of agglomerate arrangements and the
corresponding computational domains in cathode CL for the
agglomerate model developed in [271]. The in-line, uni-directional
staggered and bi-directional staggered arrangements were
considered. Since the top-down formulation is described in Section
6, the top-down formulation of the agglomerate model is not
repeated here. Corresponding to the three agglomerate arrangements, Fig. 47 shows the reaction rates at a current density of
0.6 A cm 2 in cathode CL [271]. To achieve the same current
density, it can be observed that the reaction rate is the lowest with
in-line agglomerate arrangement, and it is the highest with unidirectional agglomerate arrangement. High reaction rate indicates
that the activation loss is high. With such significant differences
between the three arrangements, it was concluded that the in-line
arrangement features much better cell performance than the other
two arrangements. Generally, the agglomerate models are useful in
optimizing the CL micro-structures and catalyst distribution.
However, for more detailed investigation on the reaction mechanisms and material micro-structures in CL, MD and office-lattice
pseudo particle simulations are needed.
9.2. Molecular dynamics and office-lattice pseudo particle
simulations
Fig. 45. Liquid water distribution in GDL obtained from PN model when a steady state
is reached [279].
The micro-structure of CL depends on particle size of catalyst,
catalyst loading, ionomer fraction, addition of other materials (such
as PTFE to increase hydrophobicity), temperature, and so on. Due to
the fact that the electrochemical reactions only occur at the threephase contact interface, the CL micro-structure becomes critically
important in optimizing cell performance. Investigation of reaction
mechanisms and morphology in CL is therefore demanded, which
270
K. Jiao, X. Li / Progress in Energy and Combustion Science 37 (2011) 221e291
Fig. 46. Schematics of agglomerate arrangements and corresponding computational domains of cathode CL [271].
can be achieved by MD and office-lattice pseudo particle simulations. Quantum mechanical MD and MC simulations have been
performed to study the elementary reaction processes on catalyst
surfaces [293e298] by focusing on the understanding of carbon
monoxide (CO) adsorption processes [293,294] and the reaction
kinetics of on different metal surfaces [295e297]. CGMD simulations have also been carried out to evaluate the key factors for CL
fabrication [299]. The simulations were performed with the presence of carbon/platinum particle, water, ionomer and solvent (used
for CL fabrication) [299]. Fig. 48 shows the equilibrium structure of
a catalyst blend composed of carbon/platinum, ionomer, water and
implicit solvent [299,346]. It was found that the hydrophobic
backbones are attached to the catalyst surfaces, and the polymer
side chains tend to leave from the catalyst surfaces. Water and
ionomer are clustered together. Generally, the CGMD simulation
was demonstrated as a powerful tool to provide valuable microstructural information of CL to help design optimization.
9.3. Summary
The different models for water transport in CL are reviewed in
Section 9. Water transports in both the ionomer and pores in CL, and
the models for water transport in ionomer and pores are introduced in
Sections 7 (in ionomer) and 8 (in pores). However, the top-down VOF
K. Jiao, X. Li / Progress in Energy and Combustion Science 37 (2011) 221e291
271
model cannot be applied with CL micro-structures due to the small
length scale, and the rule-based models (the FM and PN models) based
on top-down formulation may not be able to be directly applied to CL
micro-structures as well. Knudsen diffusion also needs to be considered in the small pores of CL. Top-down formulation can still be used in
agglomerate models with simplified CL structures that neglect the
nanometre holes in CL. The MD and off-lattice pseudo particle simulations have been conducted for investigating the reaction mechanisms and material micro-structures.
10. Modeling water transport in flow channel
The multi-component gas transport in flow channel can be
modeled by using the top-down approach described in Section 6 by
fully accounting for the convective and diffusive mass transfer. The
multi-component two-phase transport involving gas and liquid
water in flow channel can be simulated by using the two-fluid and
mixture models presented in Section 6, or by using the VOF and LB
models described in Section 8. The two-fluid and mixture models
are able to predict liquid water fraction, however, the detailed
liquid water transport behaviours cannot be simulated because the
interface between gas and liquid cannot be tracked. Better correlations between gas and liquid velocities are also needed to predict
liquid water transport more accurately. Due to the numerical
instability, the LB model often assumed unit density ratio between
gas and liquid. This assumption may be reasonable for flow in GDL
and CL when the capillary effect dominates. However, for the
convection dominated flow in flow channel, the inertia force
becomes significant, and the unit (or low) density ratio assumption
fails. The most suitable model for simulating liquid water transport
so far is the VOF model, which can account for all the major flow
effects such as inertia force, viscous force, surface tension and
gravity and track the interface between gas and liquid. By using the
VOF model, liquid water transport behaviour in flow channel has
been extensively investigated in single serpentine [262,265e267],
serpentine-parallel [11], and straight-parallel [12] flow channels.
The effect of surface wettability of flow channel on liquid water
transport was also studied by using VOF model [263,264].
10.1. Volume-of-fluid model
Corresponding to the liquid water injection condition for
membrane hydration, Figs. 49 and 50 show the liquid water
transport in a three-cell stack with serpentine flow channel design
600
Uni-directional staggered
agglomerate arrangement
Reaction Rate, A cm
-3
500
400
300
Bi-directional staggered
agglomerate arrangement
200
In-line agglomerate
arrangement
100
0
0
2
4
6
8
10
Distance in CL from Membrane to GDL, μm
Fig. 47. Reaction rates at a current density of 0.6 A cm
to the agglomerate arrangements in Fig. 46 [271].
2
in cathode CL corresponding
Fig. 48. Equilibrium structure of a catalyst blend composed of carbon/platinum
(black), ionomer (red), water (green) and implicit solvent [299,346]. (For interpretation
of the references to colour in this figure legend, the reader is referred to the web
version of this article.)
[11] obtained from VOF model. When small liquid water droplets
are supplied (Fig. 49), all the droplets hit the end wall of the inlet
manifold driven by gas flow, and then they all flow into the last cell
(cell 3). When a large amount of liquid water is supplied (Fig. 50),
the supplied liquid water moves slower than the small droplets,
and it is able to be distributed into the three cells, however, most of
the supplied liquid water flows into cell 3. Largely uneven distributions of supplied liquid water among the different cells are
observed for the two liquid water injection modes, suggesting that
liquid water injection may cause water flooding and membrane
dehydration in different single cells of PEMFC stack. Liquid water
sticking at the dead ends of the inlet and outlet manifolds can also
be observed at the end of the water transport processes in Figs. 49
and 50. Fig. 51 presents the liquid water transport in a cross section
of the inlet manifold corresponding to the water transport process
in Fig. 50 [11]. It can be observed that the liquid water is first
pushed to the end wall of the inlet manifold, with a vortex formed
there, the liquid water is pushed back towards to the inlet again,
and then it is moved to the three cells by the inlet gas flow. Such
unordered liquid water movement may affect reactant transport.
Figs. 52 and 53 shows the pressure drops for each single cell and the
whole stack corresponding to the water transport processes in Figs.
49 and 50, respectively. In Figs. 49 and 52, when the liquid water
droplets enter cell 3 (in the first 3 ms), the pressure drop in this cell
increases significantly, and then decreases to its normal state. The
pressure drop variations shown in Fig. 53 with a large amount of
liquid water supply is more significant than in Fig. 52, it can be
noticed from Fig. 53 that the pressure drops in all the three cells all
change dramatically until the liquid water is removed out of the
stack. The results shown in Figs. 52 and 53 indicate that liquid
water movement leads to unordered reactant transport, which may
cause unstable cell performance. Recall the experimental observations shown in Figs. 23 and 24, different cell voltages were
reported at the same current density for the same cell due to the
presence of liquid water. In fact, the variations of pressure drop
caused by the presence of water can also be used as useful indicators of the water content within the cell, as the experimental
272
K. Jiao, X. Li / Progress in Energy and Combustion Science 37 (2011) 221e291
Fig. 49. Liquid water transport in a three-cell stack with serpentine flow channel
design when liquid water droplets are injected [11].
study in [382] showed that maintaining back pressure at the
cathode exit is an indication that the product water is removed
properly.
10.2. Other aspects
Even though the various numerical simulations have been carried
out to investigate liquid water transport in flow channel, it is
worthwhile to be noticed that most of the studies neglected the
effects of GDL. In fact, the flow characteristics in flow channel with
and without GDL can be significantly different. Fig. 54 shows the
pressure drops from inlet to outlet of a serpentine flow channel with
a GDL compressed to different thicknesses and without GDL [189].
The pressure drops with GDL were experimentally measured by
inserting various metal shims parallel to the GDL to control the
thickness, and the pressure drop without GDL was calculated based
on known flow resistances. For the most uncompressed condition,
the porosity and permeability of the GDL are all the highest, allowing
more gas flow through it. By further compressing the GDL, it is
harder for gas flowing through it, and gas can only flow in flow
channel if without GDL. It can be observed in Fig. 54 that the pressure increases with the decrement of GDL thickness for the same
inlet Reynolds number, and it is the highest without GDL. The
differences among the different compressions and without GDL are
Fig. 51. Liquid water transport in a cross section of inlet manifold corresponding to the
water transport process in Fig. 50 [11].
very significant, especially at high inlet Reynolds number. The
observation in Fig. 54 indicates that GDL has to be considered for
studying the flow characteristics in flow channel, especially when
strong flow through GDL is possible. However, as mentioned in
Section 8, the small pores in GDL need very fine grids, and therefore
usually only small samples of GDL were simulated to investigated
liquid water behaviours. For the computational domain of real GDL
micro-structure with a dimension corroding to the length of flow
channel (e.g. on the level of 10 or 102 mm), extremely large number
of grid is needed, which is not affordable with most of the current
computational power. However, with the porous formulation by
assuming homogenous GDL (as described in Section 6), it is still
possible to simulate flow channel and GDL simultaneously, which
may result in unrealistic liquid water behaviour in GDL. It is also
worthwhile to mention that a new concept of “porous media flow
field” has been proposed for fuel cell water management, and
promising potential in performance improvement has also been
demonstrated through numerical modeling [383,384]. This concept
is to fill porous media in the flow channel region, allowing a simultaneous transport of gas, liquid, heat and electron through the
porous media flow channel. With such porous flow channel,
the transport in flow channel can be modeled in a similar way as in
GDL [383,384].
10.3. Summary
Fig. 50. Liquid water transport in a three-cell stack with serpentine flow channel
design when a large amount of liquid water is supplied [11].
The different models for water transport in flow channel are
reviewed in Section 10. The previously described full cell models
273
K. Jiao, X. Li / Progress in Energy and Combustion Science 37 (2011) 221e291
1.4
1.2
Pres s ure D rop, kPa
a
Cell 1
Cell 2
Cell 3
Overall
30
Cell 1
Cell 2
Cell 3
Overall
25
Pres s ure D rop, kPa
a
1
0.8
0.6
20
15
10
5
0.4
0
0.2
0
0.5
1
1.5
2
2.5
0
3
0.5
1
0.6
Pressure D rop, kPa
0.5
Cell 1
Cell 2
Cell 3
Overall
0.4
b
1
Pres s ure D rop, kPa
b
1.5
2
2.5
3
Time, ms
Time, ms
0.8
0.3
Cell 1
Cell 2
Cell 3
Overall
0.6
0.4
0.2
0.2
0
10
20
30
40
50
0
60
10
20
11. Cold start
In winter conditions, it is unavoidable for vehicles driving below
the freezing point of water (0 C), therefore, for successful
commercialization of PEMFC in automotive applications, rapid
startup from subzero temperatures must be achieved, which is
referred to as “cold start” of PEMFC. The major problem of PEMFC
cold start is that the product water freezes when the temperature
inside PEMFC is lower than the freezing point of water. If the CL is
fully covered by ice before cell temperature rises above freezing
point, the electrochemical reaction may be stopped due to the
blockage of the reaction site (as shown in Fig. 3a). In addition, ice
formation may also result in serious damage to the structure of the
MEA. Therefore, for PEMFC in automotive applications, successful
cold start is of paramount importance. PEMFC cold start capability
50
60
70
Fig. 53. Pressure drops for each single cell and the whole stack corresponding to the
water transport process in Fig. 50 (a: for the first 3 ms; b: for the whole process) [11].
still needs significant improvement, especially for unassisted cold
start, this is because assisted cold start might increase the volume
and weight of the system, as well as the operation complexity and
installation costs.
160
140
Pressure D rop, kPa
(Section 6), and the VOF and LB models (Section 8) can all be used
for modeling water transport in flow channel. Due to the numerical
instability, the LB model often assumes unit density ratio between
gas and liquid. This assumption may be reasonable for flow in GDL
and CL when the capillary effect dominates. However, for the
convection dominated flow in flow channel, the inertia force
becomes significant, and the unit (or low) density ratio assumption
fails. The most suitable model for simulating liquid water transport
in flow channel so far is therefore the VOF model.
40
Time, ms
Time, ms
Fig. 52. Pressure drops for each single cell and the whole stack corresponding to the
water transport process in Fig. 49 (a: for the first 3 ms; b: for the whole process) [11].
30
120
100
80
60
8 μm
10 μm
12 μm
No GDL
40
20
0
0
1000
2000
3000
4000
5000
6000
Inlet Reynolds Number
Fig. 54. Pressure drops from inlet to outlet of a serpentine flow channel with a GDL
compressed to different thicknesses and without GDL [189].
274
K. Jiao, X. Li / Progress in Energy and Combustion Science 37 (2011) 221e291
11.1. Experimental work
Experimental studies for PEMFC cold start have been carried out
by focusing on the effects of cell design and operating conditions on
cold start performance, material degradations, property measurement at subzero temperatures, and so on. Tajiri et al. [385,386]
designed experimental procedures to mainly investigate the
effects of MEA characteristics on PEMFC cold start, and the initial
water content in Nafion membrane was controlled by different
purging methods. They found that the best cold start capability is
achieved with the lowest initial water content, because more
product water can be taken by the membrane, resulting in less
water freezing in the CL. Hou et al. [387,388] investigated the
performance degradations after freezeethaw cycles, and they
indicated that membrane degradation is not a major issue for
PEMFC cold start. Cho et al. [389,390] studied the freezeethaw
cycles with a subfreezing temperature of 10 C, they reported that
the unpurged initial condition resulted in severe performance
degradation, and the solution purge (purged with anti-freezes such
as methanol) might be an effective way to achieve fast startup.
Similarly, other experimental studies that focused on the effects of
purging methods [391], startup temperatures [392], and purging
temperatures [393] on cold start performance were also carried
out. Thompson et al. [394] experimentally determined the ORR
kinetics for PEMFC operating at subzero temperatures, and no
significant change in the ORR mechanism was found. Ge and Wang
[395,396] conducted the first two in situ visualizations of ice
formation on the cathode CL surface by cooling the transparent
cells with coolant recirculation, and they have reported that the
freezing point of water in the CL is around 1 C. Ishikawa et al.
[397] also reported that the freezing temperature of liquid water in
the CL changes between 0.1 and 2.2 C. All these studies
[395e397] showed that liquid water freezes at the temperatures
lower than 0 C so-called the freezing point depression, which is
mainly due to the wettability and extremely small pore size in CL,
known as the GibbseThomson undercooling [398]. Sun et al. [399]
described a new method called catalytic hydrogen/oxygen reaction
assisted cold start by supplying mixed hydrogen and air (or oxygen)
into the fuel cell, however, further investigations are needed for this
method. McDonald et al. [400] conducted an ex situ freezeethaw
study on MEA, and they reported that no degradation was observed
for dry MEA. Cappadonia et al. [65,66] performed experimental
measurements of Nafion membrane conductance at various
subzero temperatures, and Thompson et al. [64] reported the
conductivities of Nafion membrane at different subzero temperatures based on the measured membrane conductance and sample
size. Chacko et al. [401] measured the HFR of a PEMFC before,
during and after cold start. The measurements in [64e66,401] all
observed decrements of membrane conductance/conductivity at
subzero temperatures. In [64], it was also found that such decrements occurred when the water phase changes were observed, and
the amount of non-frozen membrane water content at different
subzero temperatures were also estimated.
11.2. Numerical model
Numerical models for PEMFC cold start have also been developed. Sundaresan and Moore [402] conducted an analytical model
for cold start of PEMFC stacks. This one-dimensional model could
predict the temperature for each of the single cells by performing
energy balance and heat transfer analysis, and this model could also
reveal the effects of the endplate thermal mass and the internal
heating method for PEMFC cold start. Khandelwal et al. [403] also
conducted a one-dimensional thermal model, and similar to Sundaresan and Moore [402], the cold start capability of PEMFC stacks
can be evaluated. They reported that adjusting the startup current
density, coolant heating, isolation of stack endplates are all effective
ways to optimize the cold start performance. Mao and Wang [404]
developed an analytical model, not only for temperature, this onedimensional model can also predict the amount of ice formation in
CL, water transport, changes of cell voltage and current density etc.
Wang [405,406] also conducted analytical studies and defined
some important parameters affecting PEMFC cold start performance, and a three-step electrode process was also defined. The
analytical models in [402e406] can only roughly predict the PEMFC
cold start performance, and in order to investigate the fundamental
physics of PEMFC cold start, multi-dimensional and multiphase
models are needed. However, not many literatures are related to
this field [258e260,407e409]. Ahluwalia and Wang [407] conducted a simple two-dimensional cold start model for single
PEMFCs, and they reported that high startup current density is
favourable for rapid cold start, and they also investigated the effects
of feed gas temperatures, operating pressure, and electrical heating
on cold start performance. Mao et al. [258] developed a threedimensional, multiphase model, Meng [408] conducted a twodimensional multiphase model for PEMFC cold start simulations,
and based on the work in Mao et al. [258,404], Jiang et al.
[259,260,409] further conducted non-isothermal cold start simulations for PEMFC, and evaluated the effects different design and
operating parameters on cold start performance. However, all the
numerical models in [258e260,407e409] assumed instantaneous
desublimation of water vapour to ice, and the reaction product is
water vapour. Since no liquid water is considered, these models can
only provide reasonable results when the temperature is lower
than the freezing point of water. It is worthwhile to mention that
several correlations for proton conductivity of Nafion membrane at
subzero temperatures have been developed from available experimental data. These correlations can be found in [62,259,406].
Development of multi-dimensional, multiphase PEMFC cold
start model with the capability to consider the detailed phase
changes of water in both the membrane electrolyte and the pore
volumes of the GDL and CL needs to be carried out for better
predictions of cold start performance. Recently, a three-dimensional multiphase cold start model has been developed with the
capability to account for all kinds of water (vapour, liquid water, ice,
non-frozen water in ionomer and frozen water in ionomer) [62].
This model was then used to evaluate different design and operating parameters on cold start performance [63,261]. This model
can be briefly represented by the following conservation equations.
Mass of gas mixture (solved in flow channel, GDL and CL):
v
3 1
vt
slq
!
sice rg þ V$ rg u g ¼ Sm
(104)
Momentum of gas mixture (solved in flow channel, GDL and CL):
!
!
!
rg !
rg !
ug
ug ug
v
þ V$
vt 3 1 slq sice
32 1 slq sice 2
!
!
ug
¼ Vpg þ mg V$ V
þV
3 1 slq sice
3 1
!!
!
ug
2
m V V$
þ Su
3 g
3 1 slq sice
!T
ug
slq
sice
!!
ð105Þ
Gas species (solved in flow channel, GDL and CL, i represents
hydrogen, oxygen or vapour):
v
3 1
vt
slq
!
sice rg Yi þ V$ rg u g Yi ¼ V$ rg Deff
i VYi þ Si
(106)
275
K. Jiao, X. Li / Progress in Energy and Combustion Science 37 (2011) 221e291
Liquid water (solved in GDL and CL): can be represented by
Equation (64)
Ice (solved in GDL and CL):
vð3sice rice Þ
¼ Sice
vt
rmem
EW
vt
In Membrane
Non-frozen
Membrane Water
In CL
Non-frozen
Membrane Water
(107)
Non-frozen membrane water (solved in membrane and CL): can
be represented by Equation (65)
Frozen membrane water (solved in membrane):
v ulf
Normal Operating Condition:
In GDL
Vapour
Vapour
Liquid
Liquid
Cold Start:
¼ Sfmw
In Membrane
Non-frozen
Membrane Water
Frozen
Membrane Water
In CL
Non-frozen
Membrane Water
Vapour
(108)
Energy (solved in whole computational domain): can be represented by Equation (66)
Electronic potential (solved in CL, GDL, BP): can be represented
by Equation (67)
Ionic potential (solved in CL, membrane): can be represented by
Equation (68)
Apparently, this cold start model is essentially based on the
framework of the top-down two-fluid model described in Section
6, by adding two more conservation equations for ice in pores of
CL and GDL and frozen water in membrane (Equations (107) and
(108)). The phase changes between all different water are
accounted for in the source terms, in a similar approach to the
two-fluid model. It is also worthwhile to be mentioned that
another three-dimensional multiphase cold start model was
developed based on the mixture model framework described in
Section 6 by replacing liquid water by ice [258e260] (liquid water
and frozen water in ionomer were neglected in this model). So far
the models based on the two-fluid model framework [62,63,261]
and based on the mixture model framework [258e260] are the
only two three-dimensional multiphase cold start models, and the
cold start model based on the two-fluid model framework
[62,63,261] is the only model that fully accounts for all kinds of
water (vapour, liquid water, ice, non-frozen water in ionomer and
frozen water in ionomer) with non-equilibrium phase change
processes.
Since the two-fluid model is described in Section 6, therefore
the details of the cold start model based on the two-fluid model
framework are not presented here. As shown in Fig. 4, the phase
change of water is very complicated and therefore hard to be
implemented. Due to the fact that non-humidified inlet gases are
often supplied during PEMFC cold start to avoid ice formation.
Therefore the non-equilibrium water transfer into and out of ionomer can be assumed only between non-frozen water in ionomer
and vapour in pores. In CL, for simplicity, it was also assumed that
the non-frozen water in ionomer freezes to ice in pores [62,63,261].
Based on these simplifications, the water phase change processes in
different components of PEMFC for both the normal operating
condition and cold start with partially or non-humidified inlet
gases are shown in Fig. 55 [62], which was implemented in the cold
start model in [62,63,261]. Similar to the two-fluid model described
in Section 6, this cold start model also assumes that the non-frozen
water in ionomer is equivalent to liquid water [137], and therefore
the latent heats of the non-frozen water in ionomer and liquid
water were assumed to be the same, as it was shown that the
difference is very small [410]. For modeling water freezing and
melting in ionomer, Equation (9) can be used to calculate the
maximum allowed non-frozen water content in ionomer, and the
phase change function can be implemented in the source terms of
the water conservation equations as described in Section 6. The
experiments in [395e397] all observed that the fusion and desublimation of water in CL take place at the temperatures below the
In GDL
Vapour
Liquid
Liquid
Ice
Ice
Fig. 55. Schematics of water phase change in different components of PEMFC for both
normal operating condition and cold start (with partially or non-humidified gases
supplied) [62].
normal freezing point of water (273.15 K), which is mainly due to
the wettability and extremely small pore size in CL, known as the
GibbseThomson undercooling [398]. The difference between the
freezing point in small pores and normal freezing point of water
(TN ¼ 273.15 K) is defined as the freezing point depression, TFPD (K),
and can be calculated as
TFPD ¼
TN sj273:15K cosq
rice hfusn rCL;GDL
(109)
where hfusn (J kg 1) is the latent heat of fusion for water, rCL,GDL (m)
the pore radius of CL or GDL, s (N m 1) the surface tension coefficient between liquid water and gas, and rice (kg m3) the density of
ice. By using typical values of these parameters, it was calculated
that TFPD is about 1 K in CL and about 0 K in GDL [62]. Also note
that the ice formation in flow channel was neglected in all the
models to the best of the authors’ knowledge, because ice is formed
in CL and GDL first, and sticks on the solid materials so that hard to
move. Details of the cold start model can be found in [62,63].
11.3. Cold start characteristics
Based on the cold start model developed in [62], extensive
analyses of different colds start processes were carried out
[62,63,261]. Fig. 56 shows the evolutions of current densities, ice
volume fraction in cathode CL and cell temperature for both the
3 C) cold start
failed (from
10 C) and successful (from
processes [62]. For the failed cold start process shown in Fig. 56a,
the current density increases quickly at the beginning due to the
fast electrochemical double layer charging and discharging process.
Then the variation of the current density becomes the minimum for
a period of time, owing to the combined effects of the membrane
hydration, temperature increment and ice blockage. Finally the
current density drops fast indicating that the cold start process is
failed. The fast drop of current density is caused by the ice blockage
on the active catalyst surface, and it occurs when the ice volume
fraction in the cathode CL reaches unity (higher than 0.9). The ice
volume fraction in cathode CL increases almost linearly in Fig. 56a.
During a cold start process, the product water is absorbed by ionomer, taken by gas streams and freezes to ice. Before ice formation
can take place in CL pores, the water content in the ionomer of CL
must reach the saturation value to freeze (Equation (9)). Therefore
the ice volume fraction in cathode CL in Fig. 6 remains unchanged
in the first second of the cold start process because the product
276
Current D ens ity, A cm
-2
a
0.2
1
0.18
0.8
0.16
0.6
0.14
0.4
0.12
0.2
Ice V olume Fraction in Catho de CL
K. Jiao, X. Li / Progress in Energy and Combustion Science 37 (2011) 221e291
0.1
0
2
4
6
8
10
12
14
10
9
8
7
6
5
4
3
2
1
0
-1
-2
-3
0.25
0.2
0.15
0.1
0.05
0
0
5
10
15
20
25
30
o
Current D ens ity, A cm
-2
b
Cell Temperature, C
Time, s
35
Time, s
Fig. 56. Evolutions of current densities, ice volume fraction in cathode CL and cell
temperature for a failed (a: from 10 C) and a successful (b: from 3 C) cold start
processes [62].
water is absorbed by ionomer. This indicates that purging the cell to
ensure dry ionomer (more capacity to store product water) is
critical to reduce ice formation. For the successful cold start process
shown in Fig. 56b, the current density increases with the increment
of the cell temperature, and the increment of the cell temperature
becomes the lowest at around 0 C, because heat is needed for ice
melting.
Corresponding to the failed cold start process shown in Fig. 56a,
the transient ice volume fraction in a cross section of cathode CL,
non-frozen water content in a cross section of membrane and CLs,
and temperature in a cross section of all cell components are presented in Figs. 57e59, respectively [62]. In Fig. 57, the ice first
appears under the land because the temperature is low in those
areas (close to the surrounding walls), which result in low saturation water content. In the areas close to the membrane, more water
in the ionomer of CL could diffuse into the membrane. Therefore,
the ice is first generated at the locations away from the membrane
interface. As time passes, more ice is formed and the cathode CL is
fully blocked (12 s), indicating that no further electrochemical
reaction can take place. In Fig. 58, it can be observed that the nonfrozen water content in the cathode CL increases fast and the
changes of the water content in the membrane and anode CL are
relatively slow. The reason is that the water production rate is
higher than the water diffusion rate in the ionomer (low diffusivity
at subzero temperatures). Fig. 58 even shows that the water
content difference across the membrane is still very significant
even at 35 s (23 s after the electrochemical reaction stops),
suggesting that increasing the ionomer fraction in the cathode CL
may have more significant effects than increasing the thickness of
the membrane layer in reducing the amount of ice formation. In
Fig. 59, the highest temperature is very close to the membrane due
to the ohmic heating, a slight shift to the cathode CL due to the
other heating sources (activational heat, reversible heat and latent
heat). The temperature is also the highest under the flow channel
rather than under the land because the heat is lost at the
surrounding walls. It is also observed that the temperature in the
anode is slightly higher than in the cathode, and the reasons are: 1)
the ohmic heat is the highest in the anode CL (a low water content
results in a low membrane conductivity); 2) the heat transfer rate is
higher in the anode, due to the higher thermal conductivity of
hydrogen by comparing with oxygen and the blockage of ice in the
cathode CL.
Corresponding to the successful cold start process shown in
Fig. 56b, the transient ice and liquid water volume fractions in
a cross section of CLs and GDLs are shown in Figs. 60 and 61,
respectively [62]. In Fig. 60, the ice first melts in the CLs under
the land (t ¼ 1 s), then the whole CL and the GDL under the land,
and finally the whole area. It is also observed that the ice melting
in the anode is slightly faster than in the cathode, due to the
faster heat transfer in the anode, as mentioned earlier. In Fig. 61,
the locations of the liquid water formation matches the locations
of the ice melting in Fig. 60. It should be noticed that the liquid
water in the cathode CL becomes the maximum at 3 s, and at
30 s, the liquid water in the anode is still decreasing and it
remains almost unchanged in the cathode, because water is
produced in cathode.
The transport phenomena during PEMFC cold start are investigated in Figs. 56e61. The key parameters affecting PEMFC cold start
performance also need to be identified. The effects of membrane
thickness on cold start performance for both the potentiostatic and
galvanostatic cold start processes from 20 C are shown in Fig. 62
[261]. During a potentiostatic cold start process, the cell voltage is
controlled while the current density varies and mainly depends on
the cell voltage, temperature increment/decrement, ionomer
hydration/dehydration and water freezing/melting. On the other
hand, the current density is controlled during a galvanostatic cold
start process and therefore the cell voltage changes. For the
potentiostatic cold start processes in Fig. 62a, it can be noticed that
the highest current densities achieved during the cold start
processes are about 0.19, 0.25 and 0.49 A cm 2 with Nafion 117, 115
and 112, respectively. Such significant difference among the three
membranes is mainly attributed to the different ohmic losses,
which are mainly caused by the difference in the membrane
thickness. Since the membrane conductivities at subzero temperatures are much lower than at normal operating temperatures (e.g.
the membrane conductivity with a water content of 15 at 20 C is
only about 10% of at 80 C), the effects of membrane thickness on
the cell performance are therefore more significant at subzero
temperatures. Due to the fact that the water production rate is
proportional to the current density, the ice formation process is the
slowest with Nafion 117 and the quickest with Nafion 112, and the
failure times of the cold start processes are about 8, 7, and 4 s with
Nafion 117, 115 and 112, respectively. For the galvanostatic cold start
processes in Fig. 62b (the water production rates are same),
different from the potentiostatic condition, the thinnest membrane
results in the slowest ice formation process for the galvanostatic
condition, the reason is that the Nafion 112 membrane can absorb
water more quickly than the other two membranes (to be explained
with Fig. 64).
The evolutions of various heat generation rates, heat loss rates
through BPs, and cell temperatures when Nafion 112 is used for
both the potentiostatic and galvanostatic cold start processes from
K. Jiao, X. Li / Progress in Energy and Combustion Science 37 (2011) 221e291
277
Fig. 57. Transient ice volume fraction in a cross section of cathode CL corresponding to the failed cold start process in Fig. 56a [62].
20 C are presented in Fig. 63 [261]. During a cold start process,
heat is generated from the electrochemical reactions (activational
heat and reversible heat), electron and ion transport (ohmic heat),
and water phase change (latent heat). For the potentiostatic cold
start process in Fig. 63a, it can be observed that the largest heating
source is the ohmic heat, mainly due to the low membrane
conductivity at subzero temperatures. The activational heat is the
second largest heating source, followed by the reversible heat and
latent heat. The heat loss is caused by the heat transfer at the outer
surfaces of the BPs and by the outflow of the gas streams, and it has
been found that the heat loss due to the outflow of the gas streams
can be neglected by comparing with the heat loss from the BPs [62].
For the galvanostatic cold start process in Fig. 63b, the activational
heat generation rates increase when the ice volume fractions are
high, due to the fact that higher activational energy is needed to
maintain the current density when the reaction area becomes
smaller due to the ice blockage. The largest heating source is the
activational heat, since the current density is lower than the
potentiostatic condition in Fig. 63a. The latent heat is the lowest.
The heat losses to surroundings all increase with the increment of
cell temperature for both the potentiostatic and galvanostatic cold
start processes in Fig. 63, and the cell temperature increment
becomes slower during the cold start process.
To further examine the effects of the membranes on the ice
formation processes, the evolutions of amounts of ice formation,
amounts of water absorbed by ionomer, and amounts of water
taken by the gas streams for the potentiostatic (cell voltage is
0.3 V) and galvanostatic (current densities are 0.15 and
0.05 A cm 2) cold start processes from 20 C are shown in Fig. 64
[261]. All the water amounts shown in this figure are normalized
by the amounts of the water production. It should be noticed that
the summation of the amount of water absorbed by the ionomer,
amount of ice formation and amount of water taken by the gas
streams is equal to the amount of the water production. Therefore,
the normalized amounts of water in Fig. 64 represent the
percentages of the product water in different forms. In Fig. 64a for
the potentiostatic cold start process, it can be observed that the
amount of water taken by the gas streams can be neglected, due to
the fact that the saturation pressures of water vapour are very low
at subzero temperatures. Initially most of the product water is
absorbed by the ionomer, and most of the absorbed water is in the
cathode CLs (shown in Fig. 58). With the increments of the nonfrozen water content in the ionomer of the cathode CLs, the
saturation levels are reached, and then most of the product water
becomes ice. Fig. 64a shows that the Nafion 117 membrane is able
to absorb the largest portions of the product water among the
278
K. Jiao, X. Li / Progress in Energy and Combustion Science 37 (2011) 221e291
Fig. 58. Transient non-frozen water content in a cross section of membrane and CLs corresponding to the failed cold start process in Fig. 56a [62].
three membranes before the CLs are severely blocked by ice.
However, for the galvanostatic (0.15 A cm 2) cold start process in
Fig. 64b, the water production rates are the same for the different
membranes (same current density), therefore it is more meaningful to compare the effects of membrane thickness on ice
formation. It can be noticed that by using Nafion 112 the ionomer
can absorb about 10% more of the product water than by using
Nafion 115 and 117, and the reason is that a thinner membrane
results in a larger water gradient across the membrane, hence
faster water absorption. Advantage of Nafion 112 in absorbing
product water is more apparent in Fig. 64c when the current
density is lower (0.05 A cm 2). In this case, the water production is
also lower, allowing more water being absorbed by ionomer
before ice fully plugging CL. Even the thinnest membrane (Nafion
112) shows the greatest effect on reducing ice formation, it should
be mentioned that the thickness of membrane still needs to be
sufficiently high to maintain enough water capacity to store
product water. With the same membrane (Nafion 117), the effects
of ionomer volume fraction on ice formation are investigated in
Fig. 65 [261]. It can be observed that increasing the ionomer
volume fraction in the CLs from 0.2 to 0.4 can absorb 20% more of
the product water, which is a significant improvement in reducing
the ice formation. One of the reasons is that the higher ionomer
volume fraction increases the water capacity of the ionomer in the
CLs, and another is that it also provides wider paths for the water
transport in the CLs, therefore the membrane can absorb the
product water in the cathode CLs more effectively.
Fig. 66 presents the evolutions of current densities and cell
temperatures for the self and assisted potentiostatic (cell voltage is
0.3 V) cold start processes from 20 C when Nafion 117 is used
[261]. It can be noticed that cell insulation, heating the outer
surface and inlet air can all accelerate the temperature increment,
however, since all the cold start processes failed at about the same
time (about 9 s), the ice formation rate is not effectively reduced. It
was reported that a heating power of 0.04 W cm 2 is needed to
increase the cell temperature for 1 C in 9 s when heating the outer
surface, however, a heating power of only 0.01 W cm 2 is needed
for the same effect when heating the inlet air [261]. This indicates
that heating the inlet air is more efficient than heating the outer
surface, because heating the inlet air results in higher temperature
increment in CL, which increases the reaction rate and therefore the
heat generation. However, heating on the outer surface may be
easier than heating the inlet air, because the inlet air must be
heated to very high temperatures to carry enough amount of heat,
or the inlet air flow rate needs to be very high. Another effective
method to accelerate the heating-up process is reducing the total
thermal mass of the cell [261].
Generally, based on the present understanding of PEMFC cold
start, a thinner membrane is more favourable in reducing the ice
formation since it can result in a larger water gradient across the
membrane thus accelerating the water absorption from the cathode
CL into the membrane. However, the membrane thickness still needs
to be kept sufficiently high to ensure a sufficient amount of water
capacity to store the product water. The potentiostatic condition is
K. Jiao, X. Li / Progress in Energy and Combustion Science 37 (2011) 221e291
279
Fig. 60. Transient ice volume fraction in a cross section of CLs and GDLs corresponding
to the successful cold start process in Fig. 56b [62].
Fig. 59. Transient temperature in a cross section of all cell components corresponding
to the failed cold start process in Fig. 56a [62].
more favourable than the galvanostatic condition for PEMFC cold
start. Even though larger portions of the product water can be
absorbed by the ionomer at lower current densities to reduce the ice
formation rate, operating at high current densities is still favourable
due to the faster heating-up. The gas streams can only take negligible
portion of the product water due to the low saturation pressures of
water at subzero temperatures, and therefore has negligible
improvement in reducing the ice formation. Optimizing the ionomer
volume fraction in the CLs is an effective way to decelerate the ice
formation by increasing the water capacity of the ionomer in the CLs
and accelerating the water diffusion into the membrane. The
external heating on the outer surfaces of PEMFCs results in direct
improvements in raising the cell temperature, however, with
negligible improvement in reducing the ice formation. Heating up
the inlet air can increase the cell temperature more effectively than
applying heat on the outer surfaces of PEMFCs. However, heating on
the surfaces of PEMFCs may be easier to implement than heating the
inlet air, because the inlet air must be heated to very high temperatures to carry enough amount of heat, otherwise the inlet air flow
rate must be very high. Heating up the inlet air also has negligible
improvement in reducing the amount of ice formation. Reducing the
total thermal mass of the cell has significant improvements in
accelerating the heating up of the cell.
11.4. Summary
The previous studies on PEMFC cold start are reviewed in
Section 11. It has been identified that ice formation that hinders
Fig. 61. Transient liquid water volume fraction in a cross section of CLs and GDLs
corresponding to the successful cold start process in Fig. 56b [62].
280
K. Jiao, X. Li / Progress in Energy and Combustion Science 37 (2011) 221e291
0.4
0.7
0.35
0.6
0.5
Nafion 117
Nafion 115
Nafion 112
0.25
0.4
0.3
0.2
0.2
0.15
0.1
Current
1
2
3
4
5
6
7
8
-15
Activational heat
0.6
0.3
Reversible heat
Cell V oltage, V
-19
Heat loss
Latent heat
0.1
-20
1
2
3
4
Time, s
b
0.9
0.8
0.7
0.65
0.6
0.6
0.5
0.55
0.4
0.5
0.3
0.45
0.2
0.4
Voltage
0.3
0.1
0
0 1 2 3 4 5 6 7 8 9 10 11 12 13
-15
-16
0.2
o
1
Ice
0.7
0.35
-18
0.2
0
Heat Generation and Los s Rates , W
0.8
0.75
Nafion 117
Nafion 115
Nafion 112
-17
Cell temperature
0.4
9
Ice V olume Fractio n in Catho de CL
0.85
-16
0.5
Time, s
b
-14
0.7
0
0
0
Ohmic heat
0.8
Activational heat
Cell temperature
Fig. 62. Effects of membrane thickness on cold start performance for potentiostatic
(a: cell voltage is 0.3 V) and galvanostatic (b: current density is 0.15 A cm 2) cold start
processes from 20 C [261].
reactant delivery and damages cell materials is the major issue for
PEMFC cold start. Enhancing water absorption by membrane
electrolyte is the most effective way to reduce the ice formation for
self cold start. Cell purging, increasing ionomer fraction in CL,
proper control of membrane thickness and startup current (or
voltage), and differential pressurization are all proper ways to
enhance membrane electrolyte water absorption. Reducing cell
thermal mass is useful for accelerating temperature increment.
Assisted cold start methods mainly involving external heating have
also been approved for fast heating-up.
12. High temperature polymer electrolyte membrane fuel cell
HT-PEMFCs with operating temperatures higher than 100 C
have attracted growing interests in the past decade. By comparing
with conventional PEMFCs operating at around 80 C, HT-PEMFCs
with elevated operating temperatures feature faster electrochemical kinetics, simpler water management (presence of liquid
water can be neglected, and membrane hydration may not be
needed), higher CO tolerance (e.g. >1% CO at 150 C [411]), and
easier cell cooling and waster heat recovery.
Although HT-PEMFCs have many attractive features, technical
challenges still remain and are mostly related to the membrane. PFSA
polymer membranes (e.g. Nation membranes) widely used in
-17
Ohmic heat
0.1
-18
Reversible heat
Heat loss
-19
Latent heat
0
-20
0
1
2
3
4
5
6
7
8
9 10 11 12 13
Time, s
Time, s
b
Cell Temperature, C
0.3
0.9
o
-2
0.8
-13
1
Cell Temperature, C
Ice
Heat Generation and Los s Rates , W
0.9
0.45
Current D ensity , A cm
a
1
0.5
Ice V olume Fractio n in Catho de CL
a
Fig. 63. Evolutions of various heat generation rates, heat loss rates through BPs, and cell
temperatures when Nafion 112 is used for potentiostatic (a: cell voltage is 0.3 V) and
galvanostatic (b: current density is 0.15 A cm 2) cold start processes from 20 C [261].
conventional PEMFCs suffer significant decrement in mechanical
strength at the high operating temperature of HT-PEMFCs, and the
much lower relative humidity in HT-PEMFCs than in conventional
PEMFCs due to the significantly increased vapour saturation pressure
with temperature also results in severe reduction of proton conductivity of the PFSA polymer membranes. Therefore, developing
membranes with high mechanical strength at the temperatures
higher than 100 C and with high proton conductivity in anhydrous
environments becomes the major challenge, and most of the previous
HT-PEMFC related researches focused on this important issue [412].
PBI membranes first proposed by Aharoni and Litt [413] have been
investigated in the previous studies and recognized as a promising
membrane when doped with a strong oxo-acid (e.g. phosphoric acid
or sulphuric acid) for HT-PEMFCs [414e416]. Moreover, phosphoric
acid doped PBI membrane first suggested for fuel cell applications by
Wainright et al. [417] has attracted most of the attentions due to its
relatively higher proton conductivity and mechanical strength by
comparing with the other types of acid doped PBI membranes.
Therefore, in this section, only the HT-PEMFCs with PBI membranes
are focused on, and excellent discussions on the development of
different high temperature membranes can be found in [412].
12.1. Experiential work
The proton conductivity measurements of phosphoric acid
doped PBI membranes have been carried out [411,417e420] and it
281
K. Jiao, X. Li / Progress in Energy and Combustion Science 37 (2011) 221e291
a
100
110
90
100
90
A mount of Water, %
80
A mount of Water, %
Ionomer volume fraction in CL = 0.2
Ionomer volume fraction in CL = 0.4
70
Ice formation
60
Nafion 117
Nafion 115
Nafion 112
50
40
Ionomer absorption
30
20
10
Ice formation
80
70
60
50
40
30
20
Ionomer absorption
10
0
0
Taken by gas streams
0
1
2
3
4
5
6
7
8
Taken by gas streams
0
9
1
2
3
4
b
6
7
8
9
Fig. 65. Evolutions of amounts of ice formation, amounts of water absorbed by ionomer, and amounts of water taken by the gas streams when Nafion 117 is used for
potentiostatic (cell voltage is 0.3 V) cold start processes from 20 C with different
ionomer volume fractions in CLs (water amounts are all normalized by amounts of the
water production) [261].
100
90
80
A mount of Water, %
5
Time, s
Time, s
70
Ice formation
60
Nafion 117
Nafion 115
Nafion 112
50
40
Ionomer absorption
30
20
10
0
Taken by gas streams
0
1
2
3
4
5
6
7
8
9
10 11 12 13
conductivity. The experimental study in [421] reported that the
thermal stability of a PBI membrane with a doping level of 4.8 (4.8
phosphoric acid molecules per PBI repeat unit) is more than
enough for use as a membrane in HT-PEMFCs. Weng [422] et al.
concluded that the EOD effect is negligible in PBI membranes,
which could further simplify the water management of HT-PEMFCs.
In situ tests of HT-PEMFCs were also conducted and promising cell
performances were obtained under various operating conditions
with good CO tolerance and acceptable performance degradation
[419,423e425].
Time, s
12.2. Numerical model
c 100
Nafion 117
Nafion 115
Nafion 112
Ionomer absorption
90
70
60
50
40
0.2
30
0.19
Taken by gas streams
0
10
20
30
40
50
60
70
Time, s
Fig. 64. Evolutions of amounts of ice formation, amounts of water absorbed by ionomer, and amounts of water taken by the gas streams for potentiostatic (a: cell voltage
is 0.3 V) and galvanostatic (b: current density is 0.15 A cm 2; c: current density is
0.05 A cm 2) cold start processes from 20 C (water amounts are all normalized by
amounts of the water production) [261].
-8
0.18
Current
0.17
0.16
-12
0.15
-14
0.14
0.13
a
b
c
d
e
0.12
0.11
has been found that the temperature, phosphoric acid doping level
and surrounding relative humidity all have significant effects on the
proton conductivity. It was shown that the proton conductivity of
phosphoric acid doped PBI membranes increases with temperature
by following the Arrhenius Law [417], and the experimental
measurements in [411,418e420] also observed that increasing both
the phosphoric acid doping level and surrounding relative
humidity all have significant improvements on the proton
-10
Temperature
0.1
0
1
2
3
4
5
6
7
-16
o
Ice formation
0
Current D ensity , A cm
10
-6
Cell Temperature, C
20
-2
A mount of Water, %
80
Numerical models for HT-PEMFCs with phosphoric acid doped
PBI membranes have also been developed in the previous studies
[425e430]. Cheddie and Munroe developed a one-dimensional
model [426] and then further extended their model to threedimensional [427]. A three-dimensional model similar to [427] was
introduced in [425] as well. Both the steady and unsteady three-
-18
-20
8
9
Time, s
Fig. 66. Evolutions of current densities and cell temperatures for self and assisted
potentiostatic (cell voltage is 0.3 V) cold start processes from 20 C when Nafion 117
is used (a: self cold start without cell insulation; b: cold start with cell insulation; c:
cold start with 0.045 W cm 2 heating on outer surface and with cell insulation; d: cold
start with 0.18 W cm 2 heating on outer surface and with cell insulation; e: cold start
with inlet air heated to 80 C without cell insulation) [261].
282
K. Jiao, X. Li / Progress in Energy and Combustion Science 37 (2011) 221e291
f1 f2
exp
T
Eact
RT
(110)
where f1 and f2 are the two different pre-exponential factors, T
(K) the temperature, R (8314 J kmol 1 K 1) the universal gas
constant, and Eact the activation energy (J mol 1) can be calculated as [430]
Eact ¼
619:6DL þ 21; 750
(111)
where DL is the phosphoric acid doping level of PBI membranes,
which is defined as the number of phosphoric acid molecules per
PBI repeat unit. The pre-exponential factors f1 and f2 in Equation
(110) further accounts for the effect of phosphoric acid doping level
and humidification on the membrane proton conductivity, and can
be calculated as [430]
f1 ¼ 168DL3
6324DL2 þ 65; 760DL þ 8460
(112)
8
< 1 þ ð0:01704T 4:767Þa; if 373:15 K T 413:15 K
f2 ¼
1 þ ð0:1432T 56:89Þa; if 413:15 K < T 453:15 K
:
1 þ ð0:7T 309:2Þa; if 453:15 K < T 473:15 K
1.2
1
(113)
psat ¼ 0:68737T 3
732:39T 2 þ 263; 390T
31; 919; 000
(114)
0.3
Cell V oltage, V
where the unit of T is K, a is the corresponding surrounding water
activity. In addition, for temperatures higher than 100 C, Equation
(11) cannot accurately predict the water saturation pressure, and the
following correlation was developed to calculate the water saturation
pressure (psat, Pa) in the temperature range of 100 to 200 C [430].
0.4
0.8
0.6
0.2
0.4
where the unit of T is K.
0
12.3. Performance
The effects of operating temperature, membrane doping level,
inlet relative humidity and feed gas characteristics on HT-PEMFC
0.1
Cell temperature = 190 oC
Cell temperature = 150 oC
Cell temperature = 110 oC
0.2
0
0.2
0.4
0.6
0.8
1
-2
kion ¼
performance with PBI membrane are shown in Figs. 67e70,
respectively [430]. In Fig. 67, the cell operates with hydrogen and
air without humidification at atmospheric pressure, the stoichiometry ratio is 2 for both the anode and cathode for a reference
current density of 1.5 A cm 2, and the phosphoric acid doping level
for the PBI membrane is 6. No apparent concentration loss is
observed for all the operating temperatures due to the high stoichiometry ratios and the avoided liquid water formation. The peak
power densities are obtained at a cell voltage of 0.4 V for all the
operating temperatures. An increment of the peak power density of
0.065 W cm 2 (from 0.213 to 0.278 W cm 2) is obtained by
increasing the operating temperature from 110 to 150 C, and the
increment is 0.062 W cm 2 (from 0.278 to 0.34 W cm 2) from 150
to 190 C. The almost linear and significant increment of the peak
power density with temperature indicates that operating the cell at
high temperatures is favourable, and the main reasons are the
enhanced electrochemical kinetics and the membrane proton
conductivity at high operating temperatures. In Fig. 68, the operating condition is similar to in Fig. 67 and the operating temperature is fixed at 190 C. The results in this figure indicate that
increasing the phosphoric acid doping level of PBI membrane have
significant improvement on the cell performance. However, it
should be noticed that the phosphoric acid doping level of 9 is still
not feasible with an operating temperature of 190 C [430], and
therefore further development to increase the thermal stability
while keeping the phosphoric acid doping level for PBI membranes
is needed. In Fig. 69, the operating condition is similar to in Fig. 67
and the operating temperature is fixed at 190 C. It should be
noticed that the relative humidities of 0.25% and 3.8% at 190 C are
equivalent to 100% relative humidities at 25 C and 80 C, respectively, meaning that the feed gases are fully humidified at room
temperature and at 80 C. It can be noticed that humidifying the
feed gases at room temperature has almost negligible improvement
on the peak power density (2%, from 0.34 to 0.346 W cm 2), and the
peak power density is increased by about 14% (from 0.34 to
0.386 W cm 2) by increasing the relative humidity from 0 to 100%
at 80 C. However, obtaining a relative humidity of 100% at room
temperature is much easier than achieving a 100% relative
humidity at 80 C. To obtain a 100% relative humidity at 80 C,
liquid water injection is needed if humidified at room temperature,
otherwise the temperature of the humidifier needs to be increased
to at least 80 C. Both of the humidification methods require more
complex system design as well as extra power consumption.
Therefore, humidifying the feeds gases may not be a very maneuverable way to improve the cell performance. Despite that,
Pow er D ens ity, W cm
dimensional models were presented by Peng et al. [428,429].
However, the numerical models in [425e427] assumed constant
proton conductivities of the membranes, and only the temperature
dependence of the membrane proton conductivity was considered
in [428,429]. As mentioned earlier, temperature, phosphoric acid
doping level and surrounding relative humidity all have significant
effects on the membrane proton conductivity, therefore, these
effects need to be fully accounted for in numerical models.
A three-dimensional non-isothermal model of HT-PEMFCs with
phosphoric acid doped PBI membranes has been developed with
a semi-empirical correlation based on the Arrhenius Law and previously reported experimental data to fully account for the effects of
temperature, phosphoric acid doping level and surrounding relative
humidity on the membrane proton conductivity, and therefore these
effects on the cell performance are all considered in this model [430].
For modeling HT-PEMFC, since the operating temperature is higher
than conventional PEMFCs (negligible liquid water formation), and
the relatively humidities of inlet gases are low (negligible membrane
hydration), the liquid water transport and water transport in
membrane can be safely neglected. Therefore, HT-PEMFC models are
much simpler than conventional PEMFC models. Since the full cell
models for conventional PEMFC are described in Section 6, therefore
the simpler HT-PEMFC model by neglecting the transport of liquid
water and water in membrane is not presented here. It is worthwhile
to be mentioned that a semi-empirical correlation was first formulated based on the Arrhenius Law and previously reported experimental data to fully account for the effects of temperature,
phosphoric acid doping level and surrounding relative humidity on
the membrane proton conductivity (kion, S m 1) in [430]:
0
1.2
Current Density, A cm-2
Fig. 67. Effects of operating temperature on HT-PEMFC performance with PBI
membrane [430].
283
K. Jiao, X. Li / Progress in Energy and Combustion Science 37 (2011) 221e291
0.3
0.6
0.2
0.4
Doping level = 9
Doping level = 6
Doping level = 3
0.2
0
0
0.2
0.4
0.6
0.8
1
1.2
0.1
0.3
0.8
0.6
0.2
0.4
0.1
Air, St = 2, 1 atm
Air, St = 4, 1 atm
Air, St = 2, 2 atm
Oxygen, St = 2, 1 atm
0.2
0
0
1.4
0
-2
0.2
0.4
0.6
0.8
1
1.2
-2
1
Cell V oltag e, V
0.8
0.4
-2
0.4
Pow er D ens ity, W cm
1
Cell V oltag e, V
1.2
0.5
Pow er D ens ity, W cm
1.2
0
1.4
-2
Current Density, A cm
Current Density, A cm
Fig. 68. Effects of PBI membrane doping level on HT-PEMFC performance [430].
Fig. 70. Effects of stoichiometry ratios of feed gases, operating pressure, and air/
oxygen on HT-PEMFC performance with PBI membrane [430].
promising cell performance can be obtained without humidification, as shown in this Figure. Fig. 70 compares the effects of stoichiometry ratios of the feed gases, operating pressure, and air/
oxygen on the cell performance. In this figure, the operating
condition is similar to in Fig. 67 and the operating temperature is
fixed at 190 C. It can be noticed that increasing the stoichiometry
ratio from 2 to 4 for both the anode and cathode has almost
negligible improvement on the cell performance, and pressurizing
the cell from 1 to 2 atm results in an increment in the peak power
density of 11%, and the increment is 15% by replacing the supplied
air with oxygen at 1 atm. Since the other design and operating
parameters are kept the same, the changes of the cell performance
are only attributed to the changes of the concentrations of the
reactants.
Generally, the cell performances by running with hydrogen and
air shown in Figs. 67e70 are very promising even without any
external humidification and pressurization. The negligible liquid
water formation and water transport in membrane greatly simplify
the water management. The easy thermal management and high CO
tolerance are also very attractive features of HT-PEMFCs. Nevertheless, the high operating temperature results in more difficulty for
startup from both the normal and subzero temperatures. Therefore,
so far HT-PEMFCs are mostly likely considered to be suitable for
stationary applications. Further developments of high temperature
1.2
0.4
0.2
0.6
0.4
0.1
Inlet RH = 0%
Inlet RH = 0.25%
Inlet RH = 3.8%
0.2
0
0
0.2
0.4
0.6
0.8
1
1.2
Pow er D ens ity, W cm
Cell V oltag e, V
0.3
0.8
-2
1
0
1.4
-2
Current Density, A cm
Fig. 69. Effects of inlet relative humidity on HT-PEMFC performance with PBI
membrane [430].
membranes as well as the materials for CL and GDL are critically
important for future practical applications of HT-PEMFC.
12.4. Summary
The previous studies on HT-PEMFC operating at between 100 C
and 200 C are reviewed in Section 12. With simplified water and
thermal management, HT-PEMFCs with acid doped PBI membranes
have attracted many attentions in the past decade. Promising
performances of HT-PEMFCs have already been demonstrated. The
main drawback of HT-PEMFC by comparing with conventional
PEMFC is the more difficult startup from both the normal and
subzero temperatures. Development of membranes those are able
to offer excellent proton conductivity and stability in dry and hot
environment is the key for the succession of HT-PEMFC. Measurements of water and proton transport properties in the different
high temperature membranes are necessary, and the measured
properties are pivotal for HT-PEMFC modeling.
13. Summary and outlook
Water management to ensure both effective membrane hydration and fast reactant delivery has become one of the most
important issues for polymer electrolyte membrane fuel cell
(PEMFC). Even though many water management strategies have
been developed in the past twenty years, understanding water
transport in PEMFC is of paramount importance to implement
these strategies properly according to design and operating
conditions. In this article, the previous studies related to water
transport in PEMFC have been comprehensively reviewed. In
PEMFC, water exists in forms of vapour, liquid and ice (during cold
start) in pores of catalyst layer (CL) and gas diffusion layer (GDL)
and in flow channel, and water in ionomer of membrane and CL can
be categorized into non-freezable, freezable and free types based
on how tight water molecules are bound to proton exchange sites.
Understanding water transport in PEMFC is therefore an intricate
study, and it requires multi-discipline knowledge that involves
fluid mechanics, thermodynamics, materials science, and so on.
Presently available experimental techniques are excellent tools
that can predict water, various gas species, temperature and current
distributions in different layers of PEMFC. Difficulties for simultaneous measurements of the various parameters, modifications of
cell and system design required by measurements, and cost for
materials and building testing apparatus are the disadvantages of
experimental observations when compared with numerical
284
K. Jiao, X. Li / Progress in Energy and Combustion Science 37 (2011) 221e291
models. Simultaneously measuring more parameters with
minimum modification of cell and system design is the primary
target of future experimental observations.
Modeling water transport in PEMFC involves developing rulebased and first-principle-based models. Rule-based models relying
on physical rules for certain flow types have been mainly used for
investigating liquid water transport in porous media and
membrane electrolyte. Development of first-principle-based
models relying on solving a set of governing equations has become
a multi-scale work. The top-down models with homogeneous
material assumption have been extensively developed for full cell
modeling, which are excellent tools for investigating the various
transport phenomena in different PEMFC components simultaneously, and therefore are useful for design optimization. Other
top-down models have also been developed to investigate the
transport phenomena in pores of GDL and CL (CL structure has to be
simplified) and in flow channel and membrane electrolyte with
fewer assumptions and more comprehensive treatments than the
full cell models. The bottom-up models from atomistic to nanoscales involving molecular dynamics (MD) and off-lattice pseudo
particle methods have been developed to study the mechanisms of
water/proton transport in membrane electrolyte, self-organization
of membrane electrolyte at different hydration levels, mechanisms
of elementary electrochemical reaction processes in CL, material
morphology in CL, and so on. The bottom-up models based on
lattice pseudo particle methods, often referred to as the lattice
Boltzmann (LB) model, have also been adopted to simulate gas and
liquid water transport in GDL and CL as well as in simplified
structures of membrane electrolyte. The different models have
been adopted for different purposes to provide a comprehensive
view of water transport in PEMFC. With the improvement of
computational power, development of full cell models involving
the volume-of-fluid (VOF) method for simulating multiphase
dynamics and the chemical potential method for water/proton
transport in membrane electrolyte and with GDL and CL (simplified
to avoid nanometre pores) micro-structures is the ultimate goal for
top-down modeling of PEMFC in the future. Increasing both the size
of computational system and time scale to account for more
complex transport phenomena is the trend of developing bottomup models in the future for PEMFC.
An important water management related issue, cold start, has
attracted many attentions in recent years. It has been identified that
ice formation that hinders reactant delivery and damages cell
materials is the major issue for PEMFC cold start, and enhancing the
water absorption by membrane electrolyte has been identified to
be the most effective to reduce the ice formation for self cold start.
Cell purging, increasing ionomer fraction in CL, proper control of
membrane thickness and startup current (or voltage), and differential pressurization have all been identified as effective ways to
enhance membrane electrolyte water absorption. Reducing cell
thermal mass has been confirmed to be useful for accelerating
temperature increment. Assisted cold start methods mainly
involving external heating have also been approved for fast heating-up, other methods such as supplying anti-freezes have also
been investigated. Measurements of material and transport properties at subzero temperatures, and experimental and bottom-up
modeling studies for understanding the water phase change
processes all need to be carried out in the future researches. The
information obtained from the experimental and bottom-up
modeling work is critically important for developing more accurate
and comprehensive top-down models for PEMFC cold start.
With simplified water and thermal management, high temperature PEMFCs (HT-PEMFCs) operating at temperatures higher than
100 C have attracted many attentions in the past decade. Promising performances of HT-PEMFCs with polybenzimidazole (PBI)
membranes have already been demonstrated. The main drawback
of HT-PEMFC by comparing with conventional PEMFC is the more
difficult startup from both the normal and subzero temperatures.
Development of membranes those are able to offer excellent proton
conductivity and stability in dry and hot environment is the key for
the succession of HT-PEMFC. Measurements of water and proton
transport properties in the different high temperature membranes
are necessary, and the measured properties are pivotal for HTPEMFC modeling.
Acknowledgments
The financial support by the Natural Sciences and Engineering
Research Council of Canada (NSERC) via a strategic Project Grant
(Grant No. 350662-07) and by Auto21 is greatly appreciated. The
authors also thank our colleagues Ibrahim Alaefour, Prodip Das,
Hao Wu and Nada Zamel and Drs. Jaewan Park and Yun Wang at the
University of California Davis and University of California Irvine for
their useful discussions and providing original artworks.
References
[1] Li X. Principles of fuel cells. New York: Taylor & Francis; 2006.
[2] Larminie J, Dicks A. Fuel cell systems explained. Chichester: John Wiley &
Sons; 2003.
[3] Grubb WT. Ion-exchange batteries. In: 11th battery research and development conference, Atlantic City, New Jersey, USA, May 22e23, 1957.
[4] Grubb WT. Fuel cell. US Patent No. 2913511; 1959.
[5] Grubb WT. Batteries with solid ion exchange electrolytes. Journal of Electrochemical Society 1959;106:275e8.
[6] Cohen R. Gemini fuel cell system. In: 20th power sources conference, Fort
Monmouth, New Jersey, USA, May 24e26, 1966.
[7] Buchi FN, Srinivasan S. Operating proton exchange membrane fuel cells
without external humidification of the reactant gases. Journal of Electrochemical Society 1997;144:2767e72.
[8] Watanabe M, Satoh Y, Shimura C. Management of the water content in
polymer electrolyte membranes with porous fiber wicks. Journal of Electrochemical Society 1993;140:3190e3.
[9] Watanabe M. Solid polymer electrolyte fuel cell. US Patent No. 5472799;
1995.
[10] Watanabe M, Uchida H, Seki Y, Emori M. Self-humidifying polymer electrolyte membranes for fuel cells. Journal of Electrochemical Society
1996;143:3847e52.
[11] Jiao K, Zhou B, Quan P. Liquid water transport in parallel serpentine channels
with manifolds on cathode side of a PEM fuel cell stack. Journal of Power
Sources 2006;154:124e37.
[12] Jiao K, Zhou B, Quan P. Liquid water transport in straight micro-parallelchannels with manifolds for PEM fuel cell cathode. Journal of Power Sources
2007;157:226e43.
[13] Reiser CA, Sawyer RD. Solid polymer electrolyte fuel cell stack water
management system. US Patent No. 4769297; 1988.
[14] Reiser CA. Water and heat management in solid polymer fuel cell stack. US
Patent No. 4826742; 1989.
[15] Pollegri A, Spaziante PM. Bipolar separator for electrochemical cells and
method of preparation thereof. US Patent No. 4197178; 1980.
[16] Voss HH, Chow CY. Coolant flow field plate for electrochemical fuel cells. US
Patent No. 5230966; 1993.
[17] Spurrier FR, Pierce BE, Wright MK. Fuel cell plates with improved arrangement of process channels for enhanced pressure drop across the plates. US
Patent No. 4631239; 1986.
[18] Rock JA. Mirrored serpentine flow channels for fuel cell. US Patent No.
6099984; 1998.
[19] Chow CY, Wozniczka B, Chan JKK. Integrated reactant and coolant fluid flow
field layer for an electrochemical fuel cell. US Patent No. 5804326; 1998.
[20] Chow CY, Wozniczka B, Chan JKK. Integrated reactant and coolant fluid flow
field layer for a fuel cell with membrane electrode assembly. Canadian
Patent No. 2274974; 1999.
[21] Wilson MS. Fuel cell with interdigitated porous flow-field. US Patent No.
5641586; 1997.
[22] Kazim A, Liu HT, Forges P. Modelling of performance of PEM fuel cell with
conventional and interdigitated flow fields. Journal of Applied Electrochemistry 1999;29:1409e16.
[23] Murphy OJ, Cisar A, Clarke E. Low cost light weight high power density PEM
fuel cell stack. Electrochimica Acta 1998;43:3829e40.
[24] Goo Y, Kim J. New concept of bipolar plate and humidity control in PEMFC.
In: 1st European PEFC forum, Lucerne, Switzerland, July 2e6, 2001.
[25] Watkins DS, Dircks KW, Epp DG. Novel fuel cell fluid flow field plate. US
Patent No. 4988583; 1991.
K. Jiao, X. Li / Progress in Energy and Combustion Science 37 (2011) 221e291
[26] Watkins DS, Dircks KW, Epp DG. Fuel cell fluid flow field plate. US Patent No.
5108849; 1992.
[27] Wood DL, Yi JS, Nguyen TV. Effect of direct liquid water injection and
interdigitated flow field on the performance of proton exchange membrane
fuel cells. Electrochimica Acta 1998;43:3795e809.
[28] Li X, Sabir I. Review of bipolar plates in PEM fuel cells: flow-field designs.
International Journal of Hydrogen Energy 2005;30:359e71.
[29] Li X, Sabir I, Park J. A flow channel design procedure for PEM fuel cells with
effective water removal. Journal of Power Sources 2007;163:933e42.
[30] Lee HK, Park JH, Kim DY, Lee TH. A study on the characteristics of the
diffusion layer thickness and porosity of the PEMFC. Journal of Power
Sources 2004;131:200e6.
[31] Jordan LR, Shukla AK, Behrsing T, Avery NR, Muddle BC, Forsyth M. Diffusion
layer parameters influencing optimal fuel cell performance. Journal of Power
Sources 2000;86:250e4.
[32] Park GG, Sohn YJ, Yang TH, Yoon YG, Lee WY, Kim CS. Effect of PTFE contents
in the gas diffusion media on the performance of PEMFC. Journal of Power
Sources 2004;131:182e7.
[33] Lin G, Nguyen TV. Effect of thickness and hydrophobic polymer content of
the gas diffusion layer on electrode flooding level in a PEMFC. Journal of
Electrochemical Society 2005;152:A1942e8.
[34] Qi Z, Kaufman A. Improvement of water management by a microporous
sublayer for PEM fuel cells. Journal of Power Sources 2002;109:38e46.
[35] Williams MV, Kunz HR, Fenton JM. Influence of convection through gas
diffusion layers on limiting current in PEMFCs using a serpentine flow field.
Journal of Electrochemical Society 2004;151:A1617e27.
[36] Gostick JT, Fowler MW, Pritzker MD, Ioannidis MA, Behra LM. In-plane and
through-plane gas permeability of carbon fiber electrode backing layers.
Journal of Power Sources 2006;162:228e38.
[37] Benziger J, Nehlsen J, Blackwell D, Brennan T, Itescu J. Water flow in the gas
diffusion layer of PEM fuel cells. Journal of Membrane Science
2005;261:98e106.
[38] Chen J, Matsuura T, Hori M. Novel gas diffusion layer with water management function of PEMFC. Journal of Power Sources 2004;131:155e61.
[39] Jordan LR, Shukla AK, Behrsing T, Avery NR, Muddle BC, Forsyth M. Effect of
diffusion-layer morphology on the performance of polymer electrolyte fuel
cells operating at atmospheric pressure. Journal of Applied Electrochemistry
2000;30:641e6.
[40] Lister S, McLean G. PEM fuel cell electrodes. Journal of Power Sources
2004;130:61e76.
[41] Park S, Lee JW, Popov BN. Effect of carbon loading in microporous layer on
PEM fuel cell performance. Journal of Power Sources 2006;163:357e63.
[42] Atiyeh HK, Karan K, Peppley B, Phoenix A, Halliop E, Pharoah J. Experimental
investigation of the role of a microporous layer on the water transport and
performance of a PEM fuel cell. Journal of Power Sources 2007;170:111e21.
[43] Park S, Lee JW, Popov BN. Effect of PTFE content in microporous layer on
water management in PEM fuel cells. Journal of Power Sources
2008;177:457e63.
[44] Gostick JT, Ioannidis MA, Fowler MW, Pritzker MD. On the role of the
microporous layer in PEMFC operation. Electrochemistry Communications
2009;11:576e9.
[45] Yan WM, Wu DK, Wang XD, Ong AL, Lee DJ, Su A. Optimal microporous layer
for proton exchange membrane fuel cell. Journal of Power Sources
2010;195:5731e4.
[46] Koncar GJ, Marianowski LG. Proton exchange membrane fuel cell separator
plate. US Patent No. 5942347; 1999.
[47] Reiser C. Ion exchange membrane fuel cell power plant with water
management pressure differentials. US Patent No. 5700595; 1997.
[48] Meyer AP, Scheffler GW, Margiott PR. Water management system for solid
polymer electrolyte fuel cell power plants. US Patent No. 5503944; 1996.
[49] Aldhart OJ, Rosso MJ. Ion exchange fuel cell assembly with improved water
and thermal management. US Patent No. 4826741; 1987.
[50] Buie CR, Posner JD, Fabian T, Cha S, Kim D, Prinz FB, et al. Water management
in proton exchange membrane fuel cells using integrated electroosmotic
pumping. Journal of Power Sources 2006;161:191e202.
[51] Nguyen TV, Knobbe MW. A liquid water management strategy for PEM fuel
cell stacks. Journal of Power Sources 2003;114:70e9.
[52] Landau MB. Fuel cell product-water management system. US Patent No.
4037024; 1977.
[53] Wilkinson DP, Voss HH, Prater K. Water management and stack design for
solid polymer fuel cells. Journal of Power Sources 1994;49:117e27.
[54] Voss HH, Wilkinson DP, Pickup PG, Johnson MC, Basura V. Anode water
removal: a water management and diagnostic technique for solid polymer
fuel cells. Electrochimica Acta 1995;40:321e8.
[55] Picot D, Metkemeijer R, Bezian JJ, Rouveyre L. Impact of the water symmetry
factor on humidification and cooling strategies for PEM fuel cell stacks.
Journal of Power Sources 1998;75:251e60.
[56] Bazylak A. Liquid water visualization in PEM fuel cells: a review. International Journal of Hydrogen Energy 2009;34:3845e57.
[57] Li H, Tang Y, Wang Z, Shi Z, Wu S, Song D, et al. A review of water flooding
issues in the proton exchange membrane fuel cell. Journal of Power Sources
2008;178:103e17.
[58] Dai W, Wang H, Yuan XZ, Martin JJ, Yang D, Qiao J, et al. A review on water
balance in the membrane electrode assembly of proton exchange membrane
fuel cells. International Journal of Hydrogen Energy 2009;34:9461e78.
285
[59] Yousfi-Steiner N, Moçotéguy P, Candusso D, Hissel D, Hernandez A,
Aslanides A. A review on PEM voltage degradation associated with water
management: impacts, influent factors and characterization. Journal of
Power Sources 2008;183:260e74.
[60] Schmittinger W, Vahidi A. A review of the main parameters influencing longterm performance and durability of PEM fuel cells. Journal of Power Sources
2008;180:1e14.
[61] Ji M, Wei Z. A review of water management in polymer electrolyte
membrane fuel cells. Energies 2009;2:1057e106.
[62] Jiao K, Li X. Three-dimensional multiphase modeling of cold start processes
in polymer electrolyte membrane fuel cells. Electrochimica Acta 2009;
54:6876e91.
[63] Jiao K, Li X. Effects of various operating and initial conditions on cold start
performance of polymer electrolyte membrane fuel cells. International
Journal of Hydrogen Energy 2009;34:8171e84.
[64] Thompson EL, Capehart TW, Fuller TJ, Jorne J. Investigation of low-temperature proton transport in Nafion using direct current conductivity and
differential scanning calorimetry. Journal of Electrochemical Society
2006;153:A2351e62.
[65] Cappadonia M, Erning JW, Stimming U. Proton conduction of Nafion 117
membrane between 140 K and room temperature. Journal of Electroanalytical Chemistry 1994;376:189e93.
[66] Cappadonia M, Erning JW, Saberi Niaki SM, Stimming U. Conductance of
Nafion 117 membranes as a function of temperature and water content.
Solid State Ionics 1995;77:65e9.
[67] Yoshida H, Miura Y. Behavior of water in perfluorinated ionomer membranes
containing various monovalent cations. Journal of Membrane Science
1992;68:1e10.
[68] Pineri M, Volino F, Escoubes M. Evidence for sorptionedesorption phenomena
during thermal cycling in highly hydrated perfluorinated membranes. Journal
of Polymer Science Part B: Polymer Physics 1985;23:2009e20.
[69] Nakamura K, Hatekeyama T, Hatekeyama H. Relationship between hydrogen
bonding and bound water in polyhydroxystyrene derivatives. Polymer
1983;24:871e6.
[70] Kim YS, Dong L, Hickner MA, Glass TE, Webb V, McGrath JE. State of water in
disulfonated poly(arylene ether sulfone) copolymers and a perfluorosulfonic
acid copolymer (Nafion) and its effect on physical and electrochemical
properties. Macromolecules 2003;36:6281e5.
[71] Siu A, Schmeisser J, Holdcroft S. Effect of water on the low temperature
conductivity of polymer electrolytes. Journal of Physical Chemistry B 2006;
110:6072e80.
[72] Springer TE, Zawodzinski TA, Gottesfeld S. Polymer electrolyte fuel cell
model. Journal of Electrochemical Society 1991;138:2334e42.
[73] Eikerling M, Kornyshev AA, Stimming U. Electrophysical properties of
polymer electrolyte membranes: a random network model. Journal of
Physical Chemistry B 1997;101:10807e20.
[74] Eikerling M, Kharkats Yu I, Kornyshev AA, Volfkovich Yu M. Phenomenological theory of electro-osmotic effect and water management in polymer
electrolyte proton-conducting membranes. Journal of Electrochemical
Society 1999;145:2684e99.
[75] Weber AZ, Newman J. Transport in polymer-electrolyte membranes.
I. Physical Model. Journal of Electrochemical Society 2003;150:A1008e15.
[76] Choi P, Jalani NH, Datta R. Thermodynamics and proton transport in Nafion.
Journal of Electrochemical Society 2005;152:E84e9.
[77] Elfring GJ, Struchtrup H. Thermodynamics of pore wetting and swelling in
Nafion. Journal of Membrane Science 2008;315:125e32.
[78] Roudgar A, Narasimachary SP, Eikerling M. Hydrated arrays of acidic surface
groups as model systems for interfacial structure and mechanisms in PEMs.
Journal of Physical Chemistry B 2006;110:20469e77.
[79] Lide DR, editor. CRC Handbook of Chemistry and Physics. 88th ed. Boca
Raton: CRC Press/Taylor and Francis; 2007.
[80] Ye Q, Nguyen TV. Three-dimensional simulation of liquid water distribution
in a PEMFC with experimentally measured capillary functions. Journal of
Electrochemical Society 2007;154:B1242e51.
[81] Zawodzinski TA, Springer TE, Davey J, Jestel R, Lopez C, Valerio J, et al. A
comparative study of water uptake by and transport through ionomeric fuel
cell membranes. Journal of Electrochemical Society 1993;140:1981e5.
[82] Zawodzinski TA, Derouin C, Radzinski S, Sherman RJ, Smith VT, Springer TE,
et al. Water uptake by and transport through Nafion 117 membranes. Journal
of Electrochemical Society 1993;140:1041e7.
[83] Connors KA. Chemical kinetics. New York: VCH Publishers; 1990.
[84] Kreuer KD, Rabenau A, Weppner W. Vehicle mechanism, a new model for the
interpretation of the conductivity of fast proton conductors. Angewandte
Chemie International Edition in English 1982;21:208e9.
[85] Agmon N. Hydrogen bonds, water rotation and proton mobility. Journal de
Chimie Physique et de Phyico-Chimie Biologique 1996;93:1714e36.
[86] Kornyshev AA, Kuznetsov AM, Spohr E, Ulstrup J. Kinetics of proton transport
in water. Journal of Physical Chemistry B 2003;107:3351e66.
[87] Agmon N. The Grotthuss mechanism. Chemical Physics Letters 1995;
244:456e62.
[88] Marx D, Tuckerman ME, Hutter J, Parinello M. The nature of the hydrated
excess proton in water. Nature 1999;397:601e4.
[89] Tuckerman ME, Laasonen K, Sprik M, Parrinello M. Ab initio molecular
dynamics simulation of the solvation and transport of H3Oþ and OH ions in
water. Journal of Physical Chemistry 1995;99:5749e52.
286
K. Jiao, X. Li / Progress in Energy and Combustion Science 37 (2011) 221e291
[90] Tuckerman ME, Laasonen K, Sprik M, Parrinello M. Ab initio molecular
dynamics simulation of the solvation and transport of hydronium and
hydroxyl ions in water. Journal of Chemical Physicals 1995;103:150e61.
[91] Essam JW. Percolation theory. Reports on Progress in Physics 1980;43:
833e912.
[92] Garboczi EJ, Snyder KA, Douglas JF, Thorpe MF. Geometrical percolation
threshold of overlapping ellipsoids. Physical Review E 1995;52:819e28.
[93] O’Hayre R, Cha SW, Collela W, Prinz FB. Fuel cell fundamentals. New York:
John Wiley & Sons; 2006.
[94] Edmondson CA, Fontanella JJ. Free volume and percolation in S-SEBS and
fluorocarbon proton conducting membranes. Solid State Ionics 2002;
152e153:355e61.
[95] Gottesfeld S, Zawodzinski TA. Polymer electrolyte fuel cells. In: Alkire RC,
Gerischer H, Kolb DM, Tobias CW, editors. Advances in electrochemical
science and engineering. New York: Wiley-VCH; 1999. p. 195e301.
[96] Rieke PC, Vanderborgh NE. Temperature dependence of water content and
proton conductivity in polyperfluorosulfonic acid membranes. Journal of
Membrane Science 1987;32:313e28.
[97] Paddison SJ. The modeling of molecular structure and ion transport in
sulfonic acid based ionomer membranes. Journal of New Materials for
Materials for Electrochemical Systems 2001;4:197e207.
[98] Paddison SJ. Fuel cell technology and applications: Part 1. In: Vielstich W,
Lamm A, Gasteiger HA, editors. Handbook of fuel cells: fundamentals,
technology, applications, vol. 3. Chichester: Wiley; 2003. p. 396e411.
[99] Paddison SJ. Proton conduction mechanisms at low degrees of hydration in
sulfonic acid-based polymer electrolyte membranes. Annual Review of
Materials Research 2003;33:289e319.
[100] Kreuer KD, Paddison SJ, Spohr E, Schuster M. Transport in proton conductors
for fuel-cell applications: simulations, elementary reactions, and phenomenology. Chemical Reviews 2004;104:4637e78.
[101] Zawodzinski TA, Neeman M, Sillerud LO, Gettesfeld S. Determination of
water diffusion coefficients in perfluorosulfonate ionomeric membranes.
Journal of Physical Chemistry 1991;95:6040e4.
[102] Weppner W, Huggins RA. Determination of the kinetic parameters of mixedconducting electrodes and application to the system Li3Sb. Journal of Electrochemical Society 1977;124:1569e78.
[103] Yeo SC, Eisenberg A. Physical properties and supermolecular structure of
perfluorinated ion-containing (Nafion) polymers. Journal of Applied Polymer
Science 1977;21:875e98.
[104] Ju H, Wang CY, Cleghorn S, Beuscher U. Nonisothermal modeling of polymer
electrolyte fuel cells. Journal of Electrochemical Society 2005;152:A1645e53.
[105] Motupally S, Becker AJ, Weidner JW. Diffusion of water in Nafion 115
membranes. Journal of Electrochemical Society 2000;147:3171e7.
[106] Fuller TF, Newman J. Experimental determination of the transport number of
water in Nafion 117 membrane. Journal of Electrochemical Society 1992;
139:1332e7.
[107] Zawodzinski TA, Springer TE, Uribe F, Gottesfeld S. Characterization of
polymer electrolytes for fuel cell applications. Solid State Ionics 1993;
60:199e211.
[108] Zawodzinski TA, Davey J, Valerio J, Geottesfeld S. The water content
dependence of electro-osmotic drag in proton-conducting polymer electrolytes. Electrochimica Acta 1995;40:297e302.
[109] Bernardi DM, Verbrugge MW. A mathematical model of the solid-polymerelectrolyte fuel cell. Journal of Electrochemical Society 1992;139:2477e91.
[110] Buchi FN, Wakizoe M, Srinivasan S. Microelectrode investigation of oxygen
permeation in perfluorinated proton exchange membranes with different
equivalent weights. Journal of Electrochemical Society 1996;143:927e32.
[111] Srinivasan S, Parthasarathy A, Appleby AJ, Martin CR. Temperature dependence of the electrode kinetics of oxygen reduction at the platinum/Nafion
interface e a microelectrode investigation. Journal of Electrochemical Society
1992;139:2530e7.
[112] Denny VE, Edwards DK, Mills AP. Transfer process. New York: Holt, Rinehart
& Winston; 1973.
[113] Prausnitz JM, Reid RC, Sherwood TK. The properties of gases and liquids. 3rd
ed. New York: McGraw-Hill; 1977.
[114] Takamura Y, Nakashima E, Yamada H, Tasaka A, Inaba M. Effects of
temperature and relative humidity on oxygen permeation in Nafion and
sulfonated poly(arylene ether sulfone). ECS Transactions 2008;16:881e9.
[115] Yoshida N, Ishisaki T, Watakabe A, Yoshitake M. Characterization of Flemion
membranes for PEFC. Electrochimica Acta 1998;43:3749e54.
[116] Cussler EL. Diffusion: mass transfer in fluid systems. Cambridge: Cambridge
University Press; 1995.
[117] Bruggeman DAG. The calculation of various physical constants of heterogeneous substances. I. The dielectric constants and conductivities of mixtures
composed of isotopic substances. Annalen der Physik 1935;24:636e64.
[118] Tomadakis MM, Sotirchos SV. Ordinary and transition regime diffusion in
random fiber structures. AIChE Journal 1993;39:397e412.
[119] Mezedur MM, Kaviany M, Moore W. Effect of pore structure, randomness
and size on effective mass diffusivity. AIChE Journal 2002;48:15e24.
[120] Neale GH, Nader WK. Prediction of transport processes within porous media:
diffusive flow processes within a homogeneous swarm of spherical particles.
AIChE Journal 1973;19:112e9.
[121] Nam JH, Kaviany M. Three-dimensional effects of liquid water flooding in the
cathode of a PEM fuel cell. International Journal of Heat and Mass Transfer
2003;46:4595e611.
[122] Das PK, Li X, Liu ZS. Effective transport coefficients in PEM fuel cell catalyst
and gas diffusion layers: beyond Bruggeman approximation. Applied Energy
2010;87:2785e96.
[123] Gostick JT, Fowler MW, Ioannidis MA, Pritzker MD, Volfkovich YM, Sakars A.
Capillary pressure and hydrophilic porosity in gas diffusion layers for polymer electrolyte fuel cells. Journal of Power Sources 2006;156:375e87.
[124] Bear J. Dynamics of fluids in porous media. New York: Elsevier; 1972.
[125] Sinha PK, Wang CY. Probing effects of GDL microstructure on liquid water
transport by pore network modeling. ECS Transactions 2006;3:387e96.
[126] Brackbill JU, Kothe DB, Zemach C. A continuum method for modeling surface
tension. Journal of Computational Physics 1992;100:335e54.
[127] Dullien FAL. Porous media: fluid transport and pore structure. 2nd ed. San
Diego: Academic Press; 1992.
[128] Udell KS. Heat transfer in porous media considering phase change and
capillarity e the heat pipe effect. International Journal of Heat and Mass
Transfer 1985;28:485e95.
[129] Nguyen TV, Lin G, Ohn H, Hussey D, Jacobson D, Arif M. Measurements of
two-phase flow properties of the porous media used in PEM fuel cells. ECS
Transactions 2006;3:415e23.
[130] Kumbur EC, Sharp KV, Mench MM. Validated Leverett approach for multiphase flow in PEFC diffusion media: I. hydrophobicity effect. Journal of
Electrochemical Society 2007;154:B1295e304.
[131] Kumbur EC, Sharp KV, Mench MM. Validated Leverett approach for multiphase flow in PEFC diffusion media: II. compression effect. Journal of Electrochemical Society 2007;154:B1305e14.
[132] Kumbur EC, Sharp KV, Mench MM. Validated Leverett approach for multiphase flow in PEFC diffusion media: III. temperature effect and unified
approach. Journal of Electrochemical Society 2007;154:B1315e24.
[133] Fairweather JD, Cheung P, St-Pierre J, Schwartz DT. A microfluidic approach
for measuring capillary pressure in PEMFC gas diffusion layers. Electrochemistry Communications 2007;9:2340e5.
[134] He W, Yi JS, Nguyen TV. Two-phase flow model of the cathode of PEM fuel
cells using interdigitated flow fields. AIChE Journal 2000;46:2053e64.
[135] Cammenga HK. Evaporation mechanisms of liquids. In: Kaldis E, editor. Current
topics in materials science, vol. 5. Amsterdam: Elsevier; 1980. p. 335e446.
[136] Collier JG, Thome JR. Convective boiling and condensation. 3rd ed. Oxford:
Clarendon Press; 1994.
[137] Wu H, Li X, Berg P. On the modeling of water transport in polymer electrolyte membrane fuel cells. Electrochimica Acta 2009;54:6913e27.
[138] Natarajan D, Nguyen TV. A two-dimensional, two-phase, multicomponent,
transient model for the cathode of a proton exchange membrane fuel cell
using conventional gas distributors. Journal of Electrochemical Society
2001;148:A1324e35.
[139] Kim J, Rao PSC, Annable MD. Determination of effective airewater interfacial
areas in partially saturated porous media using surfactant adsorption. Water
Resources Research 1997;33:2705e12.
[140] Meng H. A two-phase non-isothermal mixed-domain PEM fuel cell model
and its application to two-dimensional simulations. Journal of Power Sources
2007;168:218e28.
[141] Uchida M, Fukuoka Y, Sugawara Y, Eda N, Ohta A. Effects of microstructure of
carbon support in the catalyst layer on the performance of polymer-electrolyte fuel cells. Journal of Electrochemical Society 1996;143:2245e52.
[142] Renksizbulut M. Lecture notes for convective heat transfer. Waterloo:
University of Waterloo; 2007.
[143] Wang CY. Fundamental models for fuel cell engineering. Chemical Reviews
2004;104:4727e66.
[144] Bird RB, Stewart WE, Lightfoot EN. Transport phenomena. New York: John
Wiley & Sons; 1960.
[145] Schroeder P. Uber erstarrungseund quellungserscheinungen von gelatin.
Zeitschrift fur Physikalische Chemie 1903;45:75e117.
[146] Opekar F, Svozil D. Electric resistance in a Nafion membrane exposed to air
after a step change in the relative humidity. Journal of Electroanalytical
Chemistry 1995;385:269e71.
[147] Satterfield MB, Benziger JB. Non-Fickian water vapor sorption dynamics by
Nafion membranes. Journal of Physical Chemistry B 2008;112:3693e704.
[148] Onishi LM, Prausnitz JM, Newman J. Journal of Physical Chemistry B
2007;111:10166e73.
[149] Berg P, Promislow K, St. Pierre J, Stumper J, Wetton B. Water management in
PEM fuel cells. Journal of Electrochemical Society 2004;151:A341e53.
[150] Shah AA, Kim GS, Sui PC, Harvey D. Transient non-isothermal model of
a polymer electrolyte fuel cell. Journal of Power Sources 2007;163:793e806.
[151] Vorobev A, Zikanov O, Shamim T. A computational model of a PEM fuel cell
with finite vapor absorption rate. Journal of Power Sources 2007;166:
92e103.
[152] Focke WW. On the mechanism of transfer enhancement by eddy promoters.
Electrochimica Acta 1983;28:1137e46.
[153] Brown CJ, Pletcher D, Walsh FC, Hammond JK, Robinson D. Local mass
transport effects in the FM01 laboratory electrolyser. Journal of Applied
Electrochemistry 1992;22:613e9.
[154] Cleghorn SJC, Derouin CR, Wilson MS, Gottesfeld S. A printed circuit board
approach to measuring current distribution in a fuel cell. Journal of Applied
Electrochemistry 1998;28:663e72.
[155] Stumper J, Campell S, Wilkinson D, Johnson M, Davis M. In-situ methods for
the determination of current distributions in PEM fuel cells. Electrochimica
Acta 1998;43:3773e83.
K. Jiao, X. Li / Progress in Energy and Combustion Science 37 (2011) 221e291
[156] Rajalakshmi N, Raja M, Dhathathreyan KS. Evaluation of current distribution
in a proton exchange membrane fuel cell by segmented cell approach.
Journal of Power Sources 2002;112:331e6.
[157] Bender G, Wilson MS, Zawodzinski TA. Further refinements in the
segmented cell approach to diagnosing performance in polymer electrolyte
fuel cells. Journal of Power Sources 2003;123:163e71.
[158] Hakenjos A, Muenter H, Wittstadt U, Hebling C. A PEM fuel cell for combined
measurement of current and temperature distribution, and flow field
flooding. Journal of Power Sources 2004;131:213e6.
[159] Natarajan D, Nguyen TV. Current distribution in PEM fuel cells. Part 1:
oxygen and fuel flow rate effects. AIChE Journal 2005;51:2587e98.
[160] Natarajan D, Nguyen TV. Current distribution in PEM fuel cells. Part 2: air
operation and temperature effect. AIChE Journal 2005;51:2599e608.
[161] Wieser Ch, Helmbold A, Gülzow E. A new technique for two-dimensional
current distribution measurements in electrochemical cells. Journal of
Applied Electrochemistry 2000;30:803e7.
[162] Noponen M, Mennola T, Mikkola M, Hottinen T, Lund P. Measurement of
current distribution in a free-breathing PEMFC. Journal of Power Sources
2002;106:304e12.
[163] Noponen M, Hottinen T, Mennola T, Mikkola M, Lund P. Determination of
mass diffusion overpotential distribution with flow pulse method from
current distribution measurements in a PEMFC. Journal of Applied Electrochemistry 2002;32:1081e9.
[164] Brett DJL, Atkins S, Brandon NP, Vesovic V, Vasileiadis N, Kucernak AR.
Measurement of the current distribution along a single flow channel of
a solid polymer fuel cell. Electrochemistry Communications 2001;3:628e32.
[165] Brett DJL, Atkins S, Brandon NP, Vesovic V, Vasileiadis N, Kucernaka AR.
Localized impedance measurements along a single channel of a solid polymer fuel cell. Electrochemical and Solid-State Letters 2003;6:A63e6.
[166] Mench MM, Wang CY. An in situ method for determination of current
distribution in PEM fuel cells applied to a direct methanol fuel cell. Journal of
Electrochemical Society 2003;150:A79e85.
[167] Mench MM, Wang CY, Ishikawa M. In situ current distribution measurements in polymer electrolyte fuel cells. Journal of Electrochemical Society
2003;150:A1052e9.
[168] Dong QL, Kull J, Mench MM. Real-time water distribution in a polymer
electrolyte fuel cell. Journal of Power Sources 2005;139:106e14.
[169] Ren X, Gottesfeld S. Electro-osmotic drag of water in poly (perfluorosulfonic
acid) membranes. Journal of Electrochemical Society 2001;148:A87e93.
[170] Janssen GJM, Overvelde MLJ. Water transport in the proton-exchangemembrane fuel cell: measurements of the effective drag coefficient. Journal
of Power Sources 2001;101:117e25.
[171] Mench MM, Dong QL, Wang CY. In situ water distribution measurements in
a polymer electrolyte fuel cell. Journal of Power Sources 2003;124:90e8.
[172] Yang XG, Burke N, Wang CY, Tajiri K, Shinohara K. Simultaneous measurements of species and current distributions in a PEFC under low-humidity
operation. Journal of Electrochemical Society 2005;152:A759e66.
[173] Dong Q, Mench MM, Cleghorn S, Beuscher U. Distributed performance of
polymer electrolyte fuel cells under low-humidity conditions. Journal of
Electrochemical Society 2005;152:A2114e22.
[174] Wang M, Guo H, Ma Ch. Temperature distribution on the MEA surface of
a PEMFC with serpentine channel flow bed. Journal of Power Sources
2006;157:181e7.
[175] Shimoi R. Visualization of the membrane temperature field of a polymer
electrolyte fuel cell. Journal of Energy Resources Technology 2004;
126:258e62.
[176] Mench MM, Buford DJ, Davis TW. In situ temperature distribution
measurement in an operating polymer electrolyte fuel cell. In: 2003 internal
mechanical engineering congress & exposition, Washington, DC, USA,
November 15e21, 2003.
[177] Vie PJS, Kjelstrup S. Thermal conductivities from temperature profiles in the
polymer electrolyte fuel cell. Electrochimica Acta 2004;49:1069e77.
[178] Wilkinson M, Blanco M, Gu E, Martin J, Wilkinson D, Zhan J, et al. In situ
experimental technique for measurement of temperature and current
distribution in proton exchange membrane fuel cells. Electrochemical and
Solid State Letters 2006;9:A507e11.
[179] Lee CY, Hsieh CL, Wu GW. Novel method for measuring temperature
distribution within fuel cell using micro-sensors. Japanese Journal of Applied
Physics 2007;46:3155e8.
[180] Fabian T, O’Hayre R, Prinz FB, Santiago JG. Measurement of temperature and
reaction species in the cathode diffusion layer of a free-convection fuel cell.
Journal of Electrochemical Society 2007;154:B910e8.
[181] Alaefour IE. Temperature distribution, heat and water transport in PEMFC.
PhD proposal, University of Waterloo, Waterloo; 2008.
[182] Alaefour IE, Qi L, Park J, Li X. Temperature distribution in a PEM fuel cell with
multiple straight flow channels. In: Hydrogen and fuel cells 2009, Vancouver,
British Columbia, Canada, May 31eJune 3, 2009.
[183] Tuber K, Pocza D, Hebling C. Visualization of water building in the
cathode of a transparent PEM fuel cell. Journal of Power Sources
2003;124:403e14.
[184] Yang XG, Zhang FY, Lubawy AL, Wang CY. Visualization of liquid water
transport in a PEFC. Electrochemical and Solid-State Letters
2004;7:A408e11.
[185] Zhang FY, Yang XG, Wang CY. Liquid water removal from a polymer electrolyte fuel cell. Journal of Electrochemical Society 2006;153:A225e32.
287
[186] Spernjak D, Prasad AK, Advani SG. Experimental investigation of liquid water
formation and transport in a transparent single-serpentine PEM fuel cell.
Journal of Power Sources 2007;170:334e44.
[187] Ge S, Wang CY. Liquid water formation and transport in the PEFC anode.
Journal of Electrochemical Society 2007;154:B998e1005.
[188] Hussaini IS, Wang CY. Visualization and quantification of cathode channel
flooding in PEM fuel cells. Journal of Power Sources 2009;187:444e51.
[189] Jiao K, Park J, Li X. Experimental investigations on liquid water removal from
the gas diffusion layer by reactant flow in a PEM fuel cell. Applied Energy
2010;87:2770e7.
[190] Geiger AB, Tsukada A, Lehmann E, Vontobel P, Wokaun A, Scherer GG. In situ
investigation of two-phase flow patterns in flow fields of PEFC’s using
neutron radiography. Fuel Cells 2002;2:92e8.
[191] Pekula N, Heller K, Chuang PA, Turhan A, Mench MM, Brenizer JS, et al. Study
of water distribution and transport in a polymer electrolyte fuel cell using
neutron imaging. Nuclear Instruments & Methods in Physics Research
2005;A542:134e41.
[192] Hickner MA, Siegel NP, Chen KS, McBrayer DN, Hussey DS, Jacobson DL, et al.
Real-time imaging of liquid water in an operating proton exchange
membrane fuel cell. Journal of Electrochemical Society 2006;153:A902e8.
[193] Owejan JP, Trabold TA, Jacobson DL, Baker DR, Hussey DS, Arif M. In situ
investigation of water transport in an operating PEM fuel cell using neutron
radiography: Part 2 e transient water accumulation in an interdigitated
cathode flow field. International Journal of Heat and Mass Transfer
2006;49:4721e31.
[194] Zhang J, Kramer D, Shimoi R, Ono Y, Lehmann E, Wokaun A, et al. In situ
diagnostic of two-phase flow phenomena in polymer electrolyte fuel cells by
neutron imaging part B: material variations. Electrochimica Acta
2006;51:2715e27.
[195] Park J, Li X, Tran D, Abdel-Baset T, Hussey DS, Jacobson DL, et al. Neutron
imaging investigation of liquid water distribution in and the performance of
a PEM fuel cell. International Journal of Hydrogen Energy 2008;33:3373e84.
[196] Satija R, Jacobson DL, Arif M, Werner SA. In situ neutron imaging technique
for evaluation of water management systems in operating PEM fuel cells.
Journal of Power Sources 2004;129:238e45.
[197] Hickner MA, Siegel NP, Chen KS, Hussey DS, Jacobson DL, Arif M. In situ highresolution neutron radiography of cross-sectional liquid water profiles in
proton exchange membrane fuel cells. Journal of Electrochemical Society
2008;155:B427e34.
[198] Turhan A, Kim S, Hatzell M, Mench MM. Impact of channel wall hydrophobicity on through-plane water distribution and flooding behavior in a polymer electrolyte fuel cell. Electrochimica Acta 2010;55:2734e45.
[199] Hartnig C, Manke I, Kuhn R, Kleinau S, Goebbels J, Banhart J. High-resolution
in-plane investigation of the water evolution and transport in PEM fuel cells.
Journal of Power Sources 2009;188:468e74.
[200] Sinha PK, Halleck P, Wang CY. Quantification of liquid water saturation in
a PEM fuel cell diffusion medium using X-ray microtomography. Electrochemical and Solid-State Letters 2006;9:A344e8.
[201] Park J, Li X. An experimental and numerical investigation on the cross flow
through gas diffusion layer in a PEM fuel cell with a serpentine flow channel.
Journal of Power Sources 2007;163:853e63.
[202] Weber AZ, Newman J. Modeling transport in polymer-electrolyte fuel cells.
Chemical Reviews 2004;104:4679e726.
[203] Mauritz KA, Moore RB. State of understanding of Nafion. Chemical Reviews
2004;104:4535e85.
[204] Springer TE, Wilson MS, Gottesfeld S. Modeling and experimental diagnostics in polymer electrolyte fuel cells. Journal of Electrochemical Society
1993;140:3513e26.
[205] Bernardi DM, Verbrugge MW. Mathematical model of a gas diffusion electrode bonded to a polymer electrolyte. AIChE Journal 1991;37:1151e63.
[206] Nguyen TV, White RE. A water and heat management model for protonexchange-membrane fuel cells. Journal of Electrochemical Society
1993;140:2178e86.
[207] Fuller TF, Newman J. Water and thermal management in solid-polymerelectrolyte fuel cells. Journal of Electrochemical Society 1993;140:1218e25.
[208] Amphlett JC, Baumert RM, Mann RF, Peppley BA, Roberge PR, Harris TJ.
Performance modeling of the Ballard Mark IV solid polymer electrolyte fuel
cell: I. mechanistic model development. Journal of Electrochemical Society
1995;142:1e8.
[209] Amphlett JC, Baumert RM, Mann RF, Peppley BA, Roberge PR, Harris TJ.
Performance modeling of the Ballard Mark IV solid polymer electrolyte fuel
cell: II. empirical model development. Journal of Electrochemical Society
1995;142:9e15.
[210] Ticianelli EA. A modelling approach to the characterization of the limiting
polarization behaviour of gas diffusion electrodes. Journal of Electroanalytical Chemistry 1995;387:1e10.
[211] Yi JS, Nguyen TV. An along-the-channel model for proton exchange
membrane fuel cells. Journal of Electrochemical Society 1998;145:1149e59.
[212] Yi JS, Nguyen TV. Multicomponent transport in porous electrodes of proton
exchange membrane fuel cells using the interdigitated gas distributors.
Journal of Electrochemical Society 1999;146:38e45.
[213] Gurau V, Liu H, Kakac S. Two-dimensional model for proton exchange
membrane fuel cells. AIChE Journal 1998;44:2410e22.
[214] Gloaguen F, Durand R. Simulations of PEFC cathodes: an effectiveness factor
approach. Journal of Applied Electrochemistry 1997;27:1029e35.
288
K. Jiao, X. Li / Progress in Energy and Combustion Science 37 (2011) 221e291
[215] Bultel Y, Ozil P, Durand R. Modelling the mode of operation of PEMFC
electrodes at the particle level: influence of ohmic drop within the active
layer on electrode performance. Journal of Applied Electrochemistry
1998;28:269e76.
[216] Bultel Y, Ozil P, Durand R. Modified thin film and agglomerate models for
active layers of P.E. fuel cells. Electrochimica Acta 1998;43:1077e87.
[217] Bultel Y, Ozil P, Durand R. Modelling of mass transfer within the PEM fuel cell
active layer: limitations at the particle level. Journal of Applied Electrochemistry 1999;29:1025e33.
[218] Marr C, Li X. Composition and performance modelling of catalyst layer in
a proton exchange membrane fuel cell. Journal of Power Sources
1999;77:17e27.
[219] Okada T, Xie G, Meeg M. Simulation for water management in
membranes for polymer electrolyte fuel cells. Electrochimica Acta
1998;43:2141e55.
[220] Eikerling M, Kharkats YI, Kornyshev AA, Volfkovich YM. Phenomenological
theory of electro-osmotic effect and water management in polymer electrolyte proton-conducting membranes. Journal of Electrochemical Society
1998;145:2684e99.
[221] Futerko P, Hsing IM. Thermodynamics of water vapor uptake in perfluorosulfonic acid membranes. Journal of Electrochemical Society
1999;146:2049e53.
[222] Okada T. Theory for water management in membranes for polymer electrolyte fuel cells: part 1. The effect of impurity ions at the anode side on the
membrane performances. Journal of Electroanalytical Chemistry 1999;
465:1e17.
[223] Okada T. Theory for water management in membranes for polymer electrolyte fuel cells: part 2. The effect of impurity ions at the cathode side on the
membrane performances. Journal of Electroanalytical Chemistry
1999;465:18e29.
[224] Elliot JA, Hanna S, Elliot AMS, Cooley GE. Atomistic simulation and molecular
dynamics of model systems for perfluorinated ionomer membranes. Physical
Chemistry Chemical Physics 1999;1:4855e63.
[225] Paddison SJ, Zawodzinski TA. Molecular modeling of the pendant chain in
Nafion. Solid State Ionics 1998;113e115:333e40.
[226] Dutta S, Shimpalee S, Van Zee JW. Three-dimensional numerical simulation
of straight channel PEM fuel cells. Journal of Applied Electrochemistry
2000;30:135e46.
[227] Dutta S, Shimpalee S, Van Zee JW. Numerical prediction of mass-exchange
between cathode and anode channels in a PEM fuel cell. International
Journal of Heat and Mass Transfer 2001;44:2029e42.
[228] Zhou T, Liu H. A general three-dimensional model for proton exchange
membrane fuel cells. International Journal of Transport Phenomena
2001;3:177e98.
[229] Berning T, Lu DM, Djilali N. Three-dimensional computational analysis of
transport phenomena in a PEM fuel cell. Journal of Power Sources 2002;
106:284e94.
[230] Mazumder S, Cole JV. Rigorous 3-D mathematical modeling of PEM fuel cells:
I model predictions without liquid water transport. Journal of Electrochemical Society 2003;150:A1503e9.
[231] Lee WK, Shimpalee S, Van Zee JW. Verifying predictions of water and current
distributions in a serpentine flow field polymer electrolyte membrane fuel
cell. Journal of Electrochemical Society 2003;150:A341e8.
[232] Um S, Wang CY. Three-dimensional analysis of transport and electrochemical reactions in polymer electrolyte fuel cells. Journal of Power Sources
2004;125:40e51.
[233] Wang Y, Wang CY. Transient analysis of polymer electrolyte fuel cells.
Electrochimica Acta 2005;50:1307e15.
[234] Shimpalee S, Greenway S, Spuckler D, Van Zee JW. Predicting water and
current distributions in a commercial-size PEMFC. Journal of Power Sources
2004;135:79e87.
[235] Meng H, Wang CY. Electron transport in PEFCs. Journal of Electrochemical
Society 2004;151:A358e67.
[236] Meng H, Wang CY. Large-scale simulation of polymer electrolyte fuel cells by
parallel computing. Chemical Engineering Science 2004;59:3331e43.
[237] Ju H, Meng H, Wang CY. A single-phase, non-isothermal model for PEM fuel
cells. International Journal of Heat and Mass Transfer 2005;48:1303e15.
[238] Wang Y, Wang CY. Ultra large-scale simulation of polymer electrolyte fuel
cells. Journal of Power Sources 2006;153:130e5.
[239] Liu Z, Mao Z, Wang C, Zhuge W, Zhang Y. Numerical simulation of a mini
PEMFC stack. Journal of Power Sources 2006;160:1111e21.
[240] Natarajan D, Nguyen TV. Three-dimensional effects of liquid water flooding
in the cathode of a PEM fuel cell. Journal of Power Sources 2003;115:
66e88.
[241] Lin G, Nguyen TV. A two-dimensional, two-phase model of a PEMFC. Journal
of Electrochemical Society 2006;153:A372e82.
[242] Berning T, Djilali N. A 3D, multiphase, multicomponent model of the cathode
and anode of a PEM fuel cell. Journal of Electrochemical Society
2003;150:A1589e98.
[243] Sui PC, Kumar S, Djilali N. Advanced computational tools for PEM fuel cell
design. Part 1: development and base case simulations. Journal of Power
Sources 2008;180:410e22.
[244] Sui PC, Kumar S, Djilali N. Advanced computational tools for PEM fuel cell
design. Part 2: detailed experimental validation and parametric study.
Journal of Power Sources 2008;180:423e32.
[245] Mazumder S, Cole JV. Rigorous 3-D mathematical modeling of PEM fuel cells:
II. model predictions with liquid water transport. Journal of Electrochemical
Society 2003;150:A1510e7.
[246] Wu H, Berg P, Li X. Steady and unsteady 3D non-isothermal modeling of PEM
fuel cells with the effect of non-equilibrium phase transfer. Applied Energy
2010;87:2778e84.
[247] Wang ZH, Wang CY, Chen KS. Two-phase flow and transport in the air
cathode of proton exchange membrane fuel cells. Journal of Power Sources
2001;94:40e50.
[248] Wang CY, Cheng P. Multiphase flow and heat transfer in porous media.
Advances in Heat Transfer 1997;30:93e196.
[249] Meng H, Wang CY. New model of two-phase flow and flooding dynamics in
polymer electrolyte fuel cells. Journal of Electrochemical Society
2005;152:A1733e41.
[250] Pasaogullari U, Wang CY. Two-phase modeling and flooding prediction of
polymer electrolyte fuel cells. Journal of Electrochemical Society
2005;152:A380e90.
[251] Wang Y, Wang CY. A non-isothermal, two-phase model for polymer electrolyte fuel cells. Journal of Electrochemical Society 2006;153:A1193e200.
[252] Wang Y, Wang CY. Two-phase transients of polymer electrolyte fuel cells.
Journal of Electrochemical Society 2007;154:B636e43.
[253] Wang Y, Basu S, Wang CY. Modeling two-phase flow in PEM fuel cell
channels. Journal of Power Sources 2008;179:603e17.
[254] Basu S, Li J, Wang CY. Two-phase flow and maldistribution in gas channels of
a polymer electrolyte fuel cell. Journal of Power Sources 2009;187:431e43.
[255] Basu S, Wang CY, Chen KS. Two-phase flow maldistribution and mitigation in
polymer electrolyte fuel cells. Journal of Fuel Cell Science and Technology
2009;6:031007-1e031007-11.
[256] You L, Liu H. A two-phase flow and transport model for the cathode of PEM
fuel cells. International Journal of Heat and Mass Transfer 2002;45:2277e87.
[257] You L, Liu H. A two-phase flow and transport model for PEM fuel cells.
Journal of Power Sources 2006;155:219e30.
[258] Mao L, Wang CY, Tabuchi Y. A multiphase model for cold start of polymer
electrolyte fuel cells. Journal of Electrochemical Society 2007;154:B341e51.
[259] Jiang F, Fang W, Wang CY. Non-isothermal cold start of polymer electrolyte
fuel cells. Electrochimica Acta 2007;53:610e21.
[260] Jiang F, Wang CY. Potentiostatic startup of PEMFCs from subzero temperatures. Journal of Electrochemical Society 2008;155:B743e51.
[261] Jiao K, Li X. Cold start analysis of polymer electrolyte membrane fuel cells.
International Journal of Hydrogen Energy 2010;35:5077e94.
[262] Quan P, Zhou B, Sobiesiak A, Liu Z. Water behavior in serpentine microchannel for proton exchange membrane fuel cell cathode. Journal of Power
Sources 2005;152:131e45.
[263] Cai YH, Hu J, Ma HP, Yi BL, Zhang HM. Effects of hydrophilic/hydrophobic
properties on the water behavior in the micro-channels of a proton
exchange membrane fuel cell. Journal of Power Sources 2006;161:843e8.
[264] Mondal B, Jiao K, Li X. Three-dimensional simulation of water droplet
movement in PEM fuel cell flow channels with hydrophillic surfaces. International Journal of Energy Research, in press.
[265] Jiao K, Zhou B. Innovative gas diffusion layers and their water removal
characteristics in PEM fuel cell cathode. Journal of Power Sources
2007;169:296e314.
[266] Jiao K, Zhou B. Effects of electrode wettabilities on liquid water behaviours in
PEM fuel cell cathode. Journal of Power Sources 2008;175:106e19.
[267] Jiao K, Zhou B. Accelerated numerical test of liquid behaviour across gas
diffusion layer in PEM fuel cell cathode. Journal of Fuel Cell Science and
Technology 2009;5:041011-1e041001-10.
[268] Park J, Jiao K, Li X. Numerical investigations on liquid water removal from the
porous gas diffusion layer by reactant flow. Applied Energy 2010;87:2180e6.
[269] Song D, Wang Q, Liu Z, Navessin T, Holdcroft S. Numerical study of PEM fuel
cell cathode with non-uniform catalyst layer. Electrochimica Acta
2004;50:731e7.
[270] Wang Q, Eikerling M, Song D, Liu Z. Structure and performance of different
types of agglomerates in cathode catalyst layers of PEM fuel cells. Journal of
Electroanalytical Chemistry 2004;573:61e9.
[271] Das PK, Li X, Liu ZS. A three-dimensional agglomerate model for the cathode
catalyst layer of PEM fuel cells. Journal of Power Sources 2008;179:186e99.
[272] Park J, Matsubara M, Li X. Application of lattice Boltzmann method to
a micro-scale flow simulation in the porous electrode of a PEM fuel cell.
Journal of Power Sources 2007;173:404e14.
[273] Hao L, Cheng P. Lattice Boltzmann simulations of anisotropic permeabilities
in carbon paper gas diffusion layers. Journal of Power Sources 2009;
186:104e14.
[274] Sinha PK, Mukherjee PP, Wang CY. Impact of GDL structure and wettability
on water management in polymer electrolyte fuel cells. Journal of Materials
Chemistry 2007;17:3089e103.
[275] Mukherjee PP, Wang CY. Polymer electrolyte fuel cell modeling e a pore
scale perspective. In: Li X, editor. Progress in green energy. New York:
Springer; 2009. p. 185e228.
[276] Park J, Li X. Multi-phase micro-scale flow simulation in the electrodes of
a PEM fuel cell by lattice Boltzmann method. Journal of Power Sources
2008;178:248e57.
[277] Hao L, Cheng P. Lattice Boltzmann simulations of liquid droplet dynamic
behavior on a hydrophobic surface of a gas flow channel. Journal of Power
Sources 2009;190:435e46.
K. Jiao, X. Li / Progress in Energy and Combustion Science 37 (2011) 221e291
[278] Schulz VP, Becker J, Wiegmann A, Mukherjee PP, Wang CY. Modeling of twophase behavior in the gas diffusion medium of PEFCs via full morphology
approach. Journal of Electrochemical Society 2007;154:B419e26.
[279] Sinha PK, Wang CY. Pore-network modeling of liquid water transport in gas
diffusion layer of a polymer electrolyte fuel cell. Electrochimica Acta
2007;52:7936e45.
[280] Zhou X, Chen Z, Delgado F, Brenner D, Srivastava R. Atomistic simulation of
conduction and diffusion processes in Nafion polymer electrolyte and experimental validation. Journal of Electrochemical Society 2007;154:B82e7.
[281] Elliott JA, Paddison SJ. Modelling of morphology and proton transport in
PFSA membranes. Physical Chemistry Chemical Physics 2007;9:2602e18.
[282] Vishnyakov A, Neimark AV. Molecular simulation study of Nafion membrane
solvation in water and methanol. Journal of Physical Chemistry 2000;104:
4471e8.
[283] Vishnyakov A, Neimark AV. Molecular dynamics simulation of microstructure and molecular mobilities in swollen Nafion membranes. Journal of
Physical Chemistry 2001;105:9586e94.
[284] Spohr E, Commer P, Kornyshev AA. Enhancing proton mobility in polymer
electrolyte membranes: lessons from molecular dynamics simulations.
Journal of Physical Chemistry 2002;106:10560e9.
[285] Jang SS, Molinero V, Cagin T, Goddard WA. Nanophase-segregation and
transport in Nafion 117 from molecular dynamics simulations: effect of
monomeric sequence. Journal of Physical Chemistry 2004;108:3149e57.
[286] Goddard W, Merinov B, van Duin A, Jacob T, Blanco M, Monirelo V, et al.
Multi-paradigm multi-scale simulations for fuel cell catalysts and
membranes. Molecular Simulation 2006;32:251e68.
[287] Cui S, Liu J, Selvan ME, Keffer DJ, Edwards BJ, Steele WV. A molecular
dynamics study of a Nafion polyelectrolyte membrane and the aqueous
phase structure for proton transport. Journal of Physical Chemistry B
2007;111:2208e18.
[288] Devanathan R, Venkatnathan A, Dupuis M. Atomistic simulation of Nafion
membrane. 2: dynamics of water molecules and hydronium ions. Journal of
Physical Chemistry B 2007;111:13006e13.
[289] Devanathan R, Venkatnathan A, Dupuis M. Atomistic simulation of Nafion
membrane. 1: effect of hydration on membrane nanostructure. Journal of
Physical Chemistry B 2007;111:8069e79.
[290] Venkatnathan A, Devanathan R, Dupuis M. Atomistic simulations of hydrated
Nafion and temperature effects on hydronium ion mobility. Journal of
Physical Chemistry B 2007;111:7234e44.
[291] Wescott JT, Qi Y, Subramanian L, Capehart TW. Mesoscale simulation of
morphology in hydrated perfluorosulfonic acid membranes. Journal of
Chemical Physics 2006;124:134702-1e134702-14.
[292] Galperin DY, Khokhlov AR. Mesoscopic morphology of proton-conducting
polyelectrolyte membranes of Nafion type: a self-consistent mean field
simulation. Macromolecular Theory and Simulations 2006;15:137e46.
[293] Andreaus B, Maillard F, Kocylo J, Savinova E, Eikerling M. Kinetic modeling of
COad monolayer oxidation on carbon-supported platinum nanoparticles.
Journal of Physical Chemistry B 2006;110:21028e40.
[294] Andreaus B, Eikerling M. Active site model for CO adlayer electrooxidation
on nanoparticle catalysts. Journal of Electroanalytical Chemistry
2007;607:121e32.
[295] Norskov JK, Christensen CH. Toward efficient hydrogen production at
surfaces. Science 2006;312:1322e3.
[296] Hammer B, Norskov JK. Theoretical surface science and catalysis e calculations and concepts. In: Gates BC, Knozinger H, editors. Advances in
catalysis. Impact of surface science on catalysis, vol. 45. Amsterdam:
Elsevier; 2000. p. 71e129.
[297] Hammer B, Norskov JK. Why gold is the noblest of all the metals. Nature
1995;376:238e40.
[298] Stamenkovic VR, Mun BS, Mayrhofer KJJ, Ross PN, Markovic NM, Rossmeis J,
et al. Changing the activity of electrocatalysts for oxygen reduction by tuning
the surface electronic structure. Angewandte Chemie International Edition
2006;45:2815e983.
[299] Malek K, Eikerling M, Wang Q, Navessin T, Liu Z. Self-organization in catalyst
layers of polymer electrolyte fuel cells. Journal of Physical Chemistry C
2007;111:13627e34.
[300] Kulikovsky AA. Quasi-3D modeling of water transport in polymer electrolyte
fuel cells. Journal of Electrochemical Society 2003;150:A1432e9.
[301] Meng H. A three-dimensional PEM fuel cell model with consistent treatment
of water transport in MEA. Journal of Power Sources 2006;162:426e35.
[302] Um S, Wang CY. Computational study of water transport in proton exchange
membrane fuel cells. Journal of Power Sources 2006;156:211e23.
[303] Thampan T, Malhotra S, Tang H, Datta R. Modeling of conductive transport in
proton-exchange membranes for fuel cells. Journal of Electrochemical
Society 2000;147:3242e50.
[304] Janssen GJM. A phenomenological model of water transport in a proton
exchange membrane fuel cell. Journal of Electrochemical Society
2001;148:A1313e23.
[305] Baschuk JJ, Li X. A general formulation for a mathematical PEM fuel cell
model. Journal of Power Sources 2004;142:134e53.
[306] Fimrite J, Struchtrup H, Djilali N. Transport phenomena in polymer electrolyte
membranes.
Journal
of
Electrochemical
Society
2005;152:A1804e14.
[307] Baschuk JJ, Li X. A comprehensive, consistent and systematic mathematical
model of PEM fuel cells. Applied Energy 2009;86:181e93.
289
[308] Wang CY. Fuel cell technology and applications: Part 3. In: Vielstich W,
Lamm A, Gasteiger HA, editors. Handbook of fuel cells: fundamentals,
technology, applications, vol. 3. Chichester: Wiley; 2003. p. 337e47.
[309] Le AD, Zhou B. A general model of proton exchange membrane fuel cell.
Journal of Power Sources 2008;182:197e222.
[310] Le AD, Zhou B. Fundamental understanding of liquid water effects on the
performance of a PEMFC with serpentine-parallel channels. Electrochimica
Acta 2009;54:2137e54.
[311] Newman JS. Electrochemical systems. 2nd ed. Englewood Cliffs: Prentice
Hall; 1991.
[312] Nernst W. Zur kinetik der in losung befindlichen korper. Zeitschrift fur
Physikalische Chemie 1888;2:613e37.
[313] Einstein A. On the motion of small particles suspended in liquids at rest
required by the molecular-kinetic theory of heat. Annalen der Physik
1905;17:549e60.
[314] Pintauro PN, Bennion DN. Mass transport of electrolytes in membranes 1:
development of mathematical transport model. Industrial and Engineering
Chemistry Fundamentals 1984;23:230e4.
[315] Wohr M, Bolwin K, Schnurnberger W, Fischer M, Neubrand W,
Eigenberger G. Dynamic modelling and simulation of a polymer membrane
fuel cell including mass transport limitation. International Journal of
Hydrogen Energy 1998;23:213e8.
[316] Mason EA, Malinauskas AP. Gas transport in porous media: the dusty-gas
model. Amsterdam: Elsevier; 1983.
[317] Koter S, Hamann CH. Characteristics of ion-exchange membranes for electrodialysis on the basis of irreversible thermodynamics. Journal of Nonequilibrium Thermodynamics 1990;15:315e33.
[318] Auclair B, Nikonenko V, Larchet C, Metayer M, Dammak L. Correlation
between transport parameters of ion-exchange membranes. Journal of
Membrane Science 2002;195:89e102.
[319] West AC, Fuller TF. Influence of rib spacing in proton-exchange membrane
electrode assemblies. Journal of Applied Electrochemistry 1996;26:557e65.
[320] Hsing IM, Futerko P. Two-dimensional simulation of water transport in polymer electrolyte fuel cells. Chemical Engineering Science 2000;55:4209e18.
[321] Futerko P, Hsing IM. Two-dimensional finite-element method study of the
resistance of membranes in polymer electrolyte fuel cells. Electrochimica
Acta 2000;45:1741e51.
[322] Weber AZ, Newman J. Transport in polymer-electrolyte membranes. II.
Mathematical Model. Journal of Electrochemical Society 2004;151:A311e25.
[323] Verbrugge MW, Hill RF. Analysis of promising perfluorosulfonic acid
membranes for fuel-cell electrolytes. Journal of Electrochemical Society
1990;137:3770e7.
[324] Schlogl R. Zur theorie der anomalen osmose. Zeitschrift fur Physikalische
Chemie 1955;3:73e98.
[325] Singh D, Lu DM, Djilali N. A two-dimensional analysis of mass transport in
proton exchange membrane fuel cells. International Journal of Engineering
Science 1999;37:431e52.
[326] Baschuk JJ, Li X. Modelling of polymer electrolyte membrane fuel cells with
variable degrees of water flooding. Journal of Power Sources 2000;
86:181e96.
[327] Um S, Wang CY, Chen KS. Computational fluid dynamics modeling of proton
exchange membrane fuel cells. Journal of Electrochemical Society 2000;
147:4485e93.
[328] Murgia G, Pisani L, Valentini M, D’Aguanno B. Electrochemistry and mass
transport in polymer electrolyte membrane fuel cells: I. model. Journal of
Electrochemical Society 2002;149:A31e8.
[329] Costamagna P. Transport phenomena in polymeric membrane fuel cells.
Chemical Engineering Science 2001;56:323e32.
[330] Rowe A, Li X. Mathematical modeling of proton exchange membrane fuel
cells. Journal of Power Sources 2001;102:82e96.
[331] Thampan T, Malhotra S, Zhang JX, Datta R. PEM fuel cell as a membrane
reactor. Catalysis Today 2001;67:15e32.
[332] Hertwig K, Martens L, Karwoth R. Mathematical modelling and simulation of
polymer electrolyte membrane fuel cells. Part I: model structures and
solving an isothermal one-cell model. Fuel Cells 2002;2:61e77.
[333] Majsztrik P, Bocarsly A, Benziger J. Water permeation through Nafion
membranes: the role of water activity. Journal of Physical Chemistry B
2008;112:16280e9.
[334] Ennari J, Elomaa M, Sundholm F. Modelling a polyelectrolyte system in water
to estimate the ion-conductivity. Polymer 1999;40:5035e41.
[335] Urata S, Irisawa J, Takada A, Shinoda W, Tsusuki S, Mikami S. Molecular
dynamics simulation of swollen membrane of perfluorinated ionomer.
Journal of Physical Chemistry B 2005;109:4269e78.
[336] Khalatur PG, Talitskikh SK, Khokhlov AR. Structural organization of watercontaining Nafion: the integral equation theory. Macromolecular Theory and
Simulations 2002;11:566e86.
[337] Groot RD, Warren PB. Dissipative particle dynamics: bridging the gap
between atomistic and mesoscopic simulation. Journal of Chemical Physics
1997;107:4423e35.
[338] Groot RD. Electrostatic interactions in dissipative particle dynamicsdsimulation of polyelectrolytes and anionic surfactants. Journal of
Chemical Physics 2003;118:11265e77.
[339] Yamamoto S, Hyodo SA. A computer simulation study of the mesoscopic
structure of the polyelectrolyte membrane Nafion. Polymer Journal
2003;35:519e27.
290
K. Jiao, X. Li / Progress in Energy and Combustion Science 37 (2011) 221e291
[340] Hayashi H, Yamamoto S, Hyodo SA. Lattice-Boltzmann simulations of flow
through Nafion polymer membranes. International Journal of Modern
Physics B 2003;17:135e8.
[341] Hsu WY, Gierke TD. Elastic theory for ionic clustering in perfluorinated
ionomers. Macromolecules 1982;15:101e5.
[342] Hsu WY, Gierke TD. Ion transport and clustering in Nafion perfluorinated
membranes. Journal of Membrane Science 1983;13:307e26.
[343] Flory PJ. Statistical mechanics of chain molecules. New York: Interscience
Publishers; 1969.
[344] Smitha B, Sridhar S, Khan AA. Solid polymer electrolyte membranes for fuel
cell applications e a review. Journal of Membrane Science 2005;259:10e26.
[345] Haile SM. Fuel cell materials and components. Acta Materialia
2003;51:5981e6000.
[346] Eikerling M, Malek K. Physical modeling of materials for PEFC: structure, properties, and performance. In: Wilkinson PG, editor. Key materials for fuel cells.
[347] Bolhuis PG, Chandler D, Dellago C, Geissler PL. Transition path sampling:
throwing ropes over rough mountain passes, in the dark. Annual Review of
Physical Chemistry 2002;53:291e318.
[348] Dellago C, Bolhuis PG, Chandler D. Efficient transition path sampling:
application to Lennard-Jones cluster rearrangements. Journal of Chemical
Physics 1998;108:9236e45.
[349] Weinan E, Vanden-Eijnden E. Towards a theory of transition paths. Journal of
Statistical Physics 2006;123:503e23.
[350] Torrie G, Valleau J. Monte Carlo free energy estimates using non-Boltzmann
sampling: application to the sub-critical Lennard-Jones fluid. Chemical
Physics Letters 1974;28:578e81.
[351] Mills G, Jonsson H, Schenter GK. Reversible work transition state theory:
application to dissociative adsorption of hydrogen. Surface Science
1995;324:305e37.
[352] Paddison SJ, Elliott JA. On the consequences of side chain flexibility and
backbone conformation on hydration and proton dissociation in perfluorosulfonic acid membranes. Physical Chemistry Chemical Physics 2006;
8:2193e203.
[353] Eikerling M, Paddison SJ, Pratt LR, Zawodzinski TA. Defect structure for
proton transport in a triflic acid monohydrate solid. Chemical Physics Letters
2003;368:108e14.
[354] Yeo RS, Yeager HL. Structural and transport properties of perfluorinated ionexchange membranes. In: Conway BE, White RE, Bockris JOM, editors.
Modern Aspects of Electrochemistry, vol. 16. New York: Plenum Press; 1985.
p. 437e504 [chapter 6].
[355] Litt MH. A reevaluation of Nafion morphology. Polymer Preprint 1997;
38:80e1.
[356] Haubold HG, Vad T, Jungbluth H, Hiller P. Nano structure of Nafion: a SAXS
study. Electrochimica Acta 2001;46:1559e63.
[357] Rubatat L, Gebel G, Diat O. Fibrillar structure of Nafion: matching fourier and
real space studies of corresponding films and solutions. Macromolecules
2004;37:7772e83.
[358] Ohser J, Mucklich F. Statistical analysis of microstructures in materials
science. Chichester: Wiley; 2000.
[359] Schladitz K, Peters S, Reinel-Bitzer D, Wiegmann A, Ohser J. Design of
acoustic trim based on geometric modeling and flow simulation for nonwoven. Computational Materials Science 2006;38:56e66.
[360] Zamel N, Li X, Shen J. Correlation for the effective gas diffusion coefficient in
carbon paper diffusion media. Energy and Fuels 2009;23:6070e8.
[361] Gostick JT, Ioannidis MA, Fowler MW, Pritzker MD. Pore network modeling
of fibrous gas diffusion layers for polymer electrolyte membrane fuel cells.
Journal of Power Sources 2007;173:277e90.
[362] Bhatnagar PL, Gross EP, Krook M. A model for collision processes in gases. I:
small amplitude processes in charged and neutral one-component systems.
Physical Review 1954;94:511e25.
[363] Chen H, Chen S, Matthaeus WH. Recovery of the NaviereStokes equations
using
a
lattice-gas
Boltzmann
method.
Physical
Review
A
1992;45:R5339e42.
[364] Shan X, Chen H. Lattice Boltzmann model for simulating flows with multiple
phases and components. Physical Review E 1993;47:1815e9.
[365] Shan X, Chen H. Simulation of nonideal gases and liquidegas phase transitions by the lattice Boltzmann equation. Physical Review E 1994;49:2941e8.
[366] He X, Luo L. Theory of the lattice Boltzmann method: from the Boltzmann
equation to the lattice Boltzmann equation. Physical Review E 1997;56:6811e7.
[367] Zou Q, He X. On pressure and velocity boundary conditions for the lattice
Boltzmann BGK model. Physics of Fluids 1997;9:1591e8.
[368] Spaid M, Phelan FR. Lattice Boltzmann methods for modeling microscale
flow in fibrous porous media. Physics of Fluids 1997;9:2468e74.
[369] Raabe D. Overview of the lattice Boltzmann method for nano- and microscale fluid dynamics in materials science and engineering. Modelling and
Simulation in Materials Science and Engineering 2004;12:R13e46.
[370] Inamuro T, Ogata T, Tajima S, Konishi N. A lattice Boltzmann method for
incompressible two-phase flows with large density differences. Journal of
Computational Physics 2004;198:628e44.
[371] Park J, Huh KY, Li X. Lattice Boltzmann simulation on the liquid junction
potential in a microchannel. Journal of Electroanalytical Chemistry 2006;
591:141e8.
[372] Joshi AS, Gre KN, Peracchio AA, Chiu WKS. Lattice Boltzmann modeling of 2D
gas transport in a solid oxide fuel cell anode. Journal of Power Sources
2007;164:631e8.
[373] Yan YY, Zu YQ. A lattice Bolzmann method for incompressible two-phase
flows on partial wetting surface with large density ratio. Journal of
Computational Physics 2007;227:763e75.
[374] Hazlett RD. Simulation of capillary-dominated displacements in microtomographic images of reservoir rocks. Transport in Porous Media
1995;20:21e35.
[375] Hilpert M, Miller CT. Pore-morphology-based simulation of drainage in
totally wetting porous media. Advances in Water Resources 2001;
24:243e55.
[376] Patzek TW, Silin DB. Shape factor and hydraulic conductance in noncircular
capillaries: I. one-phase creeping flow. Journal of Colloid and Interface
Science 2001;236:295e304.
[377] Ioannidis MA, Chatzis I. Network modelling of pore structure and transport
properties of porous media. Chemical Engineering Science 1993;48:951e72.
[378] Concus P, Finn R. On the behavior of a capillary surface in a wedge. Applied
Mathematical Sciences 1969;63:292e9.
[379] Lin GY, He WS, Nguyen TV. Modeling liquid water effects in the gas diffusion
and catalyst layers of the cathode of a PEM fuel cell. Journal of Electrochemical Society 2004;151:A1999e2006.
[380] Du CY, Cheng XQ, Yang T, Yin GP, Shi PF. Numerical simulation of the ordered
catalyst layer in cathode of proton exchange membrane fuel cells. Electrochemistry Communications 2005;7:1411e6.
[381] Sun W, Peppley BA, Karan K. An improved two-dimensional agglomerate
cathode model to study the influence of catalyst layer structural parameters.
Electrochimica Acta 2005;50:3359e74.
[382] Park YH, Caton JA. Monitoring an electrode flooding through the back
pressure in a proton exchange membrane (PEM) fuel cell. International
Journal of Green Energy 2008;5:347e59.
[383] Wang Y. Porous-media flow fields for polymer electrolyte fuel cells. I. low
humidity operation. Journal of Electrochemical Society 2009;156:B1124e33.
[384] Wang Y. Porous-media flow fields for polymer electrolyte fuel cells. II.
analysis of channel two-phase flow. Journal of Electrochemical Society
2009;156:B1134e41.
[385] Tajiri K, Tabuchi Y, Wang CY. Isothermal cold start of polymer electrolyte fuel
cells. Journal of Electrochemical Society 2007;154:B147e52.
[386] Tajiri K, Tabuchi Y, Kagami F, Takahashi S, Yoshizawa K, Wang CY. Effects of
operating and design parameters on PEFC cold start. Journal of Power
Sources 2007;165:279e86.
[387] Hou J, Yu H, Zhang S, Sun S, Wang H, Yi B, et al. Analysis of PEMFC freeze
degradation at 20 C after gas purging. Journal of Power Sources 2006;
162:513e20.
[388] Hou J, Yi B, Yu H, Hao L, Song W, Fu Y, et al. Investigation of resided water
effects on PEM fuel cell after cold start. International Journal of Hydrogen
Energy 2007;32:4503e9.
[389] Cho EA, Ko JJ, Ha HY, Hong SA, Lee KY, Lim TW, et al. Characteristics of the
PEMFC repetitively brought to temperatures below 0 C. Journal of Electrochemical Society 2003;150:A1667e70.
[390] Cho EA, Ko JJ, Ha HY, Hong SA, Lee KY, Lim TW, et al. Effects of water removal
on the performance degradation of PEMFCs repetitively brought to <0 C.
Journal of Electrochemical Society 2004;151:A661e5.
[391] Guo Q, Qi Z. Effect of freezeethaw cycles on the properties and performance
of membrane-electrode assemblies. Journal of Power Sources 2006;160:
1269e74.
[392] Yan Q, Toghiani H, Lee YW, Liang K, Causey H. Effect of sub-freezing
temperatures on a PEM fuel cell performance, startup and fuel cell components. Journal of Power Sources 2006;160:1242e50.
[393] St-Pierre J, Roberts J, Colbow K, Campbell S, Nelson A. PEMFC operational and
design strategies for sub zero environments. Journal of New Materials for
Electrochemical Systems 2005;8:163e76.
[394] Thompson EL, Jorne J, Gasteiger HA. Oxygen reduction reaction kinetics in
subfreezing PEM fuel cells. Journal of Electrochemical Society
2007;154:B783e92.
[395] Ge S, Wang CY. In situ imaging of liquid water and ice formation in an
operating PEFC during cold start. Electrochemical and Solid-State Letters
2006;9:A499e503.
[396] Ge S, Wang CY. Characteristics of subzero startup and water/ice formation on
the catalyst layer in a polymer electrolyte fuel cell. Electrochimica Acta
2007;52:4825e35.
[397] Ishikawa Y, Morita T, Nakata K, Yoshida K, Shiozawa M. Behavior of water
below the freezing point in PEFCs. Journal of Power Sources
2007;163:708e12.
[398] Kurz W, Fisher J. Fundamentals of solidification. Aedermannsdorf: Trans
Tech Publications; 1989.
[399] Sun S, Yu H, Hou J, Shao Z, Yi B, Ming P, et al. Catalytic hydrogen/oxygen
reaction assisted the proton exchange membrane fuel cell (PEMFC) startup
at subzero temperature. Journal of Power Sources 2008;177:137e41.
[400] McDonald RC, Mittelsteadt CK, Thompson EL. Effects of deep temperature
cycling on Nafion 112 membranes and membrane electrode assemblies. Fuel
Cells 2004;4:208e13.
[401] Chacko C, Ramasamy R, Kim S, Khandelwal M, Mench M. Characteristic
behavior of polymer electrolyte fuel cell resistance during cold start. Journal
of Electrochemical Society 2008;155:B1145e54.
[402] Sundaresan M, Moore RM. Polymer electrolyte fuel cell stack thermal model
to evaluate sub-freezing startup. Journal of Power Sources 2005;145:
534e45.
K. Jiao, X. Li / Progress in Energy and Combustion Science 37 (2011) 221e291
[403] Khandelwal M, Lee S, Mench MM. One-dimensional thermal model of coldstart in a polymer electrolyte fuel cell stack. Journal of Power Sources
2007;172:816e30.
[404] Mao L, Wang CY. Analysis of cold start in polymer electrolyte fuel cells.
Journal of Electrochemical Society 2007;154:B139e46.
[405] Wang Y. Analysis of the key parameters in the cold start of polymer electrolyte fuel cells. Journal of Electrochemical Society 2007;154:B1041e8.
[406] Wang Y, Mukherjee PP, Mishler J, Mukundan R, Borup RL. Cold start of
polymer electrolyte fuel cells: three-stage startup characterization. Electrochimica Acta 2010;55:2636e44.
[407] Ahluwalia RK, Wang X. Rapid self-start of polymer electrolyte fuel cell stacks
from subfreezing temperatures. Journal of Power Sources 2006;162:502e12.
[408] Meng H. A PEM fuel cell model for cold-start simulations. Journal of Power
Sources 2008;178:141e50.
[409] Jiang F, Wang CY, Chen KS. Current ramping e a strategy for rapid startup of
PEM fuel cells from subfreezing environment. Journal of Electrochemical
Society 2010;157:B342e7.
[410] Ostrovskii VE, Gostev BV. Heat effects and rates and molecular mechanisms
of water sorption by perfluorinated polymer materials bearing functional
groups. Journal of Thermal Analysis and Calorimetry 1996;46:397e416.
[411] Ma YL, Wainright JS, Litt MH, Savinell RF. Conductivity of PBI membranes for
high-temperature polymer electrolyte fuel cells. Journal of Electrochemical
Society 2004;151:A8e16.
[412] Zhang J, Xie Z, Zhang J, Tang Y, Song C, Navessin T, et al. High temperature
PEM fuel cells. Journal of Power Sources 2006;160:872e91.
[413] Aharoni SM, Litt MH. Synthesis and some properties of poly-(2,5-trimethylene benzimidazole) and poly-(2,5-trimethylene benzimidazole
hydrochloride). Journal of Polymer Science: Polymer Chemistry Edition
1974;12:639e50.
[414] Wasmus S, Valeriu A, Mateescu GD, Tryk DA, Savinell RF. Characterization of
H3PO4-equilibrated Nafion 117 membranes using H-1 and P-31 NMR-spectroscopy. Solid State Ionics 1995;80:87e92.
[415] Savinell RF, Litt MH. Proton conducting polymers used as membranes. US
Patent No. 5525436; 1996.
[416] Savadogo O. Emerging membranes for electrochemical systems part II: high
temperature composite membranes for polymer electrolyte fuel cell (PEFC)
applications. Journal of Power Sources 2004;127:135e61.
291
[417] Wainright JS, Wang JT, Savinell RF, Litt MH. Acid-doped polybenzimidazoles:
a new polymer electrolyte. Journal of Electrochemical Society
1995;142:L121e3.
[418] Bouchet R, Siebert E. Proton conduction in acid doped polybenzimidazole.
Solid State Ionics 1999;118:287e99.
[419] Li Q, Hjuler HA, Bjerrum NJ. Phosphoric acid doped polybenzimidazole
membranes: physiochemical characterization and fuel cell applications.
Journal of Applied Electrochemistry 2001;31:773e9.
[420] He R, Li Q, Xiao G, Bjerrum NJ. Proton conductivity of phosphoric acid doped
polybenzimidazole and its composites with inorganic proton conductors.
Journal of Membrane Science 2003;226:169e84.
[421] Samms SR, Wasmus S, Savinell RF. Thermal stability of proton conducting
acid doped polybenzimidazole in simulated fuel cell environments. Journal
of Electrochemical Society 1996;143:1225e32.
[422] Weng D, Wainright JS, Landau U, Savinell RF. Electro-osmotic drag coefficient
of water and methanol in polymer electrolytes at elevated temperatures.
Journal of Electrochemical Society 1996;143:1260e3.
[423] Li Q, He R, Jensen JO, Bjerrum NJ. PBI-based polymer membranes for high
temperature fuel cells e preparation, characterization and fuel cell demonstration. Fuel Cells 2004;4:147e59.
[424] Schmidt TJ, Baurmeister J. Properties of high-temperature PEFC Celtec-P
1000 MEAs in start/stop operation mode. Journal of Power Sources
2008;176:428e34.
[425] Ubong EU, Shi Z, Wang X. A 3D-modeling and experimental validation
of a high temperature PBI based PEMFC. ECS Transactions 2008;16:
79e90.
[426] Cheddie D, Munroe N. Mathematical model of a PEMFC using a PBI
membrane. Energy Conversion and Management 2006;47:1490e504.
[427] Cheddie DF, Munroe NDH. Three dimensional modeling of high temperature
PEM fuel cells. Journal of Power Sources 2006;160:215e23.
[428] Peng J, Lee SJ. Numerical simulation of proton exchange membrane fuel cells
at high operating temperature. Journal of Power Sources 2006;162:1182e91.
[429] Peng J, Shin JY, Song TW. Transient response of high temperature PEM fuel
cell. Journal of Power Sources 2007;179:220e31.
[430] Jiao K, Li X. A three-dimensional non-isothermal model of high temperature
proton exchange membrane fuel cells with phosphoric acid doped polybenzimidazole membranes. Fuel Cells 2010;10:351e62.