Next Article in Journal
Enhanced Oil Recovery: An Update Review
Previous Article in Journal
Securing Fluid Resources for Geothermal Projects in a World of Water Scarcity
Previous Article in Special Issue
A Transient Model for Fuel Cell Cathode-Water Propagation Behavior inside a Cathode after a Step Potential
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Principles and Materials Aspects of Direct Alkaline Alcohol Fuel Cells

1
School of Chemical Engineering and Advanced Materials, Merz Court, Newcastle University, Newcastle upon Tyne, NE1 7RU, UK
2
Max Planck Institute for Dynamics of Complex Technical Systems, Sandtorstrasse 1, 39106 Magdeburg, Germany
3
Portable Energy Systems, Chair for Process Systems Engineering, Otto von Guericke University, Universitaetsplatz 2, 39106 Magdeburg, Germany
*
Author to whom correspondence should be addressed.
Energies 2010, 3(8), 1499-1528; https://doi.org/10.3390/en3081499
Submission received: 21 July 2010 / Revised: 6 August 2010 / Accepted: 16 August 2010 / Published: 24 August 2010
(This article belongs to the Special Issue Fuel Cells)

Abstract

:
Direct alkaline alcohol fuel cells (DAAFCs) have attracted increasing interest over the past decade because of their favourable reaction kinetics in alkaline media, higher energy densities achievable and the easy handling of the liquid fuels. In this review, principles and mechanisms of DAAFCs in alcohol oxidation and oxygen reduction are discussed. Despite the high energy densities available during the oxidation of polycarbon alcohols they are difficult to oxidise. Apart from methanol, the complete oxidation of other polycarbon alcohols to CO2 has not been achieved with current catalysts. Different types of catalysts, from conventional precious metal catalyst of Pt and Pt alloys to other lower cost Pd, Au and Ag metal catalysts are compared. Non precious metal catalysts, and lanthanum, strontium oxides and perovskite-type oxides are also discussed. Membranes like the ones used as polymer electrolytes and developed for DAAFCs are reviewed. Unlike conventional proton exchange membrane fuel cells, anion exchange membranes are used in present DAAFCs. Fuel cell performance with DAAFCs using different alcohols, catalysts and membranes, as well as operating parameters are summarised. In order to improve the power output of the DAAFCs, further developments in catalysts, membrane materials and fuel cell systems are essential.

1. Introduction

Easy storage and handling, high energy density and wide availability are features that make alcohols attractive liquid fuels for the most promising alternative power sources for transportation, portable electronics and stationary applications. One of the first investigations of alcohol in fuel cells was conducted by Palve in 1954. He demonstrated that methanol acted as a fuel in aqueous electrolytes [1]. Since the development of perfluorinated cation exchange membranes, such as Nafion®(DuPont), proton exchange membranes (PEM) have dominated as the electrolyte for fuel cells. Ethanol [2,3,4,5,6,7], 2-propanol [8,9], ethylene glycol [10,11,12,13] and glycerol [14] have all been considered for fuel cell applications. In direct alcohol fuel cells (DAFCs), alcohol is directly oxidised at the anode in an acidic medium using a proton exchange membrane (PEM). Direct methanol fuel cells are the most advanced and studied systems since methanol, with only one carbon in its molecule and no C-C bonds, is the simplest alcohol. Electrooxidation of methanol is relatively simple compared to other alcohols. However, two major obstacles have restrained the more rapid development and applications of DAFCs, even for methanol systems:
Alcohol crossover from the anode to the cathode
The relatively low activity and complex reaction mechanism of most alcohols
Apart from methanol, it is particularly difficult to break the C-C bond in alcohols during electrochemical oxidation. In addition to these two major issues, the high costs of proton exchange membranes and precious metal catalysts (Pt and Pt/Ru based catalysts), and CO poisoning of Pt catalysts at lower temperature in acidic media have further hampered development of DAFCs.
It is known that for many reactions, electrocatalysts perform better in alkaline electrolytes. In fact, one of the first direct methanol fuel cells (DMFCs) developed in 1955 by Justi and Winsel were operated in alkaline media and used porous nickel for the anode and porous nickel-silver for the cathode [1]. A problem with alkaline fuel cells is the carbonation of the solution due to CO2 production of the fuel oxidation and from air:
2OH + CO2→CO32− + H2O
This can cause solid precipitation of carbonate salts on the porous electrode and a pH decrease in the alkaline electrolyte solution [15]. Consequently, it leads to a reduction in reactivity for fuel oxidation in the system [16]. The application of anion exchange membranes (AEMs) as the solid polymer electrolyte can reduce formation of carbonate from carbon dioxide, compared to aqueous electrolyte. It was also discovered that the electro-oxidation of methanol in alkaline electrolytes is structure insensitive [17], which opens up the opportunity for use of non-precious metals, such as Pd, Ag, Ni and perovskite-type oxides [18,19,20], which are significantly cheaper than Pt based catalysts, in alkaline fuel cells. Furthermore, the ionic current in the alkaline fuel cell is due to conduction of hydroxide ions. This ionic flow is in the reverse direction to that in proton conducting systems. As such, the direction of the electro-osmotic drag is reversed, reducing alcohol crossover [21].
In this study, we review recent developments in the field of direct alkaline alcohol fuel cells. We discuss progress in understanding the reaction mechanisms of fuel oxidation and oxygen reduction in alkaline media, and the recent catalyst development for alcohol oxidation and oxygen reduction. Moreover, the development of anion exchange membranes for alkaline fuel cell applications is presented. Finally, fuel cell performance using different alcohols is compared.

2. Principles and Mechanisms of a Direct Alkaline Alcohol Fuel Cell (DAAFC)

Compared to the use of hydrogen in fuel cells, liquid alcohol fuels are easy to store and transport. The direct oxidation of alcohol will also eliminate a heavy reformer needed to produce hydrogen from liquid or gaseous carbonaceous fuels. Figure 1 shows a schematic principle of a direct methanol alkaline fuel cell with a mixture of methanol, water and sodium hydroxide circulating in the anode, and oxygen in the cathode.
Figure 1. Schematic diagram of direct methanol alkaline fuel cell.
Figure 1. Schematic diagram of direct methanol alkaline fuel cell.
Energies 03 01499 g001
On the anode, methanol is oxidised to carbon dioxide and electrons are produced, while on the cathode side, oxygen accepts electrons and is reduced to produce hydroxide ions. Hydroxide ions then transfer through the electrolyte to the anode. The reactions and thermodynamic potentials [vs. standard hydrogen electrode (SHE)] on each electrode are shown below:
Anode oxidation:
CH3OH + 6OH→6e + CO2 + 5H2O   E° = −0.81 V
Cathode reduction:
3/2O2 + 3H2O + 6e →6OHE° = 0.40 V
Overall reaction:
CH3OH + 3/2O2→CO2 + 2H2O   E° = 1.21 V

2.1. Alcohol Oxidation

The mechanism of electrochemical oxidation of methanol on platinum in alkaline systems involves formation of adsorbed methanolic species and OH on the Pt surface. The oxidation takes place through a series of reaction steps involving successive electron transfer, i.e. partial oxidation, to form adsorbed species. These react with adsorbed OH to potentially form carbon dioxide. The reaction mechanism has been written as follows [22]:
Pt + OH→Pt-(OH)ads + e
Pt + (CH3OH)sol→Pt-(CH3OH)ads
Pt-(CH3OH)ads + OH→Pt-(CH3O)ads + H2O + e
Pt-(CH3O)ads + OH→Pt-(CH2O)ads + H2O + e
Pt-(CH2O)ads + OH→Pt-(CHO)ads + H2O + e
Pt-(CHO)ads + OH→Pt-(CO)ads + H2O + e
Pt-(CHO)ads + Pt-(OH)ads + 2OH→2Pt + CO2 + 2H2O + 2e
Pt-(CHO)ads + Pt-(OH)ads + OH→Pt + Pt-(COOH)ads + H2O + e
Pt-(CO)ads + Pt-(OH)ads + OH→2Pt + CO2 + H2O + e
Pt-(CO)ads + Pt-(OH)ads↔Pt + Pt-(COOH)ads
Pt-(COOH)ads + OH→Pt-(OH)ads + HCOO
Pt-(COOH)ads + Pt-(OH)ads→2Pt +CO2 + H2O
The rate-determining step is most likely the reaction with the oxidation of the active intermediate—CHO [23,24,25,26,27,28,29]. For other alcohols, the higher energy density and lower toxicity of polyhydric alcohols still attract lots of interest in applying these alcohols in fuel cells, but the electrochemical oxidation is more difficult and the reaction mechanism is more complicated, since it involves breaking C-C bond(s). The reaction at low temperature leads to a low Faradic efficiency. The mechanism and kinetics of oxidation of different alcohols using Pt-based and non-Pt based metal catalysts have been investigated.
Tremiliosi-Filho et al. [30] and de Lima and Varela [31] investigated ethanol oxidation on polycrystalline gold electrodes. Liang et al. [32] and Nguyen et al. [33] have both studied ethanol oxidation on palladium catalysts. For ethanol oxidation in alkaline media, the formation of acetaldehyde or acetic acid was identified [34]. The steps involved can be presented as:
1. M + OH→M-OHads + e
2. M + CH3CH2OH→M-(CH3 CH2OH)ads
3. M-(CH3CH2OH)ads + 3OH→M-(CH3CO)ads + 3H2O + 3e
4. M-(CH3CO)ads + M-OHads→M-CH3COOH + M
5. M-CH3COOH + OH→M + CH3COO + H2O
M = Au or Pd.
The rate-determining step is step 4, in which the adsorbed ethoxy intermediate is removed by adsorbed hydroxyl ions to form acetate. Fang et al. studied the mechanism of ethanol electrooxidation on a palladium electrode in alkaline solution with various concentrations of NaOH (0.01 to 5 M) [35]. Cyclic voltammetry and in situ Fourier Transform Infrared (FTIR) spectroelectrochemistry were used to investigate oxidation products at different pH values. They found that the ethanol oxidation activity on Pd was largely affected by pH. The highest activity was obtained in 1 M NaOH. This was supported by their previous study [36], which was attributed to OHads species on Pd. The density functional theory (DFT) calculations show that in acidic media continued dehydrogenation of ethanol is difficult due to the lack of OH species to instantly remove hydrogen, while in alkaline media, ethanol and sufficient OH can be adsorbed on Pd leading to continuous oxidation. The main oxidation product was sodium acetate with NaOH concentrations higher than 0.5 M. No CO formation was detected by FTIR suggesting low poisoning effect with ethanol oxidation on Pd in alkaline media.
The mechanism for electrochemical oxidation of ethylene glycol in alkaline media was proposed by Matsuoka et al. [37]. They stated that ethylene glycol can be oxidised to oxalate (COO--COO-) via a non-poisoning path, and to formate via a poisoning path (Figure 2). Further oxidation of formate will produce CO poisoning species, while oxalate is very stable in alkaline medium and cannot be further oxidised, and as a result, no CO poisoning species are produced. In this case, the poisoning effect on the platinum is less significant compared to methanol.
Glycerol has attracted interest in its use as fuel for DAFCs recently. It is produced in large quantities as a byproduct of biodiesel production. DAFCs provide the possibility of generating energy from the waste. The mechanism of electrochemical oxidation of glycerol in alkaline solution was investigated by Roquet et al. and Schell et al. [38,39]. Glyceraldehyde was found to be the main reaction product, but formic, glycolic, tartronic and glyceric acids were also detected [38].
Figure 2. Reaction pathways of ethylene glycol electrochemical oxidation in alkaline media [37].
Figure 2. Reaction pathways of ethylene glycol electrochemical oxidation in alkaline media [37].
Energies 03 01499 g002
In Table 1, the electrochemical oxidation of various alcohols studied in alkaline media has been summarised.
Table 1. Electrochemical oxidation of various alcohols in alkaline medium.
Table 1. Electrochemical oxidation of various alcohols in alkaline medium.
FuelAnode ReactionsE° (V/SHE)Energy density (Wh/kg)
Methanol CH3OHCH3OH + 6OH→CO2 + 5H2O + 6e−0.816100
Ethanol CH3CH2OHCH3CH2OH +2OH→CH3CHO + 2H2O + 2e
CH3CH2OH + 4OH→CH3COOH + 3H2O + 4e
CH3CH2OH + 12 OH→2CO2 +9 H2O+12e
−0.778030
iso-Propanol CH3CHOHCH3CH3CHOHCH3 + 2OH→CH3COCH3 + 2H2O + 2e
CH3COCH3 + 16OH→3CO2 + 11H2O + 16e
−0.678600
Ethylene glycol (CH2OH)2(CH2OH)2 + 14OH→2CO32− + 10H2O + 10e or
(CH2OH)2 + 10OH→(CO2)22− + 8H2O + 8e
−0.725200
Glycerol HOCH2CHOHCH2OHHOCH2CHOHCH2OH+ 20OH→3CO32− + 14H2O + 14e or HOCH2CHOHCH2OH + 12OH→(COO-COH-COO) + 10H2O + 10e−0.695000

2.2. Oxygen Reduction

The oxygen reduction reaction (ORR) has been studied extensively. Effects of different pH, electrolytes, temperature and pre-conditioning of electrodes on ORR have been investigated. The most accepted mechanism of ORR was first proposed by Damjanovic et al. [40,41]. They suggested that the ORR proceeds along two parallel reaction paths with comparable rates. One is the direct four-electron transfer reduction from oxygen to water; the other is the formation of hydrogen peroxide as an intermediate in a two-electron transfer reaction. The reaction scheme is shown as follows [42]:
Energies 03 01499 i001
where a indicates an adsorbed state and b is the bulk solution.
The reactions in alkaline media can be presented as:
O2 + 2H2O + 4e→4OH
and the path involving hydrogen peroxide suggested in [38] is:
O2 + e ↔ O2
O2 + H2O→HO2 + OH
OH + e ↔ OH
In the low current density region, the Tafel slope is −60 mV dec−1 and the reaction order is 0.5 with respect to pH in alkaline solutions. The fractional reaction order was interpreted in terms of the first electrochemical step as rate determining under a Temkin isotherm, i.e. the adsorption of reaction intermediates Oads, OHads and HO2ads [43].
The rate expression under Temkin conditions of adsorption is:
j = k [ O 2 ] [ H + ] 1 2 exp ( η F RT ) = 3.74 k [ O 2 ] [ OH ] 1 2 exp ( η F RT )
where k is the rate constant, η is the overpotential.
In the high current density region, the Tafel slope is −120 mV dec−1 and the reaction order is 0 with respect to pH in alkaline solutions, with H2O as a reacting species. The adsorption of intermediate species changes to a Langmuir isotherm and Temkin conditions no longer hold. The rate expression under Langmuir conditions is:
j = k [ O 2 ] exp ( β η F RT )
where β is the symmetry factor.
The reaction is first order with respect to O2 in solutions. It was found that the amount of H2O2 formed is greater in alkaline than in acid media [40,44]. In alkaline solutions, about 80% of the reduction current is via the direct reduction and the remainder of the current goes to the formation of H2O2, which suggests a more complicated mechanism.

3. Catalysts for DAAFCs

As in proton exchange membrane fuel cells, platinum and other precious metals are still the most common catalysts used in DAAFCs. Low cost platinum-free and non-precious metal catalysts have been investigated for use in alkaline environment.

3.1. Alcohol Oxidation Catalysts

For alcohol oxidation under alkaline solutions, the reaction current densities obtained during alcohol oxidation at certain potentials are at least an order of magnitude greater than in acidic electrolytes. The application of alkali electrolytes could lead to a reduction in catalyst loadings, as well as allowing less expensive, non-precious metal catalysts. In this case, the cost of electrocatalysts and thus fuel cells may be reduced. Methanol oxidation in alkaline media on platinum single crystal and polycrystalline electrodes has been investigated by several research groups [27,28,29,45]. Unlike in acid, the electrode does not suffer severely from poisoning in alkali because bonding of the chemisorbed intermediates on platinum is weak and the amount of the suggested poisoning species COads, is smaller than in an acidic medium [22].

3.1.1. Precious Metal Catalysts

Uniform platinum nanotubes have been synthesised by directly mixing Ag nanowires and H2PtCl6 in saturated NaI solutions. The Pt nanotubes were tested for methanol oxidation in alkaline. They demonstrated high active areas and higher activity compared to conventional Pt/C catalysts [46]. In acid solution, Pt/Ru is the most effective bimetallic catalyst used for methanol oxidation. It has been applied for more than 30 years and no significantly better system has yet been found. Ru dissolved in a Pt matrix reduces the degree of anode poisoning. The Ru-containing catalysts shifted the onset potential for methanol oxidation to more negative values [47]. The study of Tripkovic et al. [48] investigated methanol oxidation on a Pt/C and a Pt2Ru3/C electrode and showed that in alkaline solutions the difference in activity between the catalysts was smaller than in acid solutions. This proved the above study that the poisoning effect from fuel oxidation to catalyst is less severe.
Matsuoka et al. [13] used an OH-form anion exchange membrane and polyhydric alcohols in DAAFCs with Pt-Ru/C as the anode catalysts and Pt/C and Ag for the cathode catalysts. The fuel cells showed good performance with the maximum power densities obtained in the order of ethylene glycol > glycerol > methanol > erythritol > xylitol.
For alcohol oxidation, the combination of (alloying) elements such as Ru, Mo, Sn, Re, Os, Rh, Pb and Bi with platinum give tolerance to the poisoning species compared to platinum alone.
Verma et al. [49] studied three different catalysts: Pt-black, Pt-Ru (40 wt%:20 wt%)/C and Pt/C (40 wt%), with varying loads at the anode against a MnO2 cathode for alkaline fuel cells using methanol, ethanol or sodium borohydride as fuel. Liquid KOH solution was used as the electrolyte. The Pt-Ru/C produces the best cell performance for all fuels. Ethanol produced a single peak at 0.03 V (vs. NHE) and no characteristics of C-C bond breaking were observed. Chemical analysis confirmed production of acetaldehyde and acetic acid. For methanol, a broad plateau was produced from –0.4 to 0.6 V and indicated that Pt black was more active for ethanol oxidation than for methanol oxidation.
Electrooxidation of ethanol on Pt-ZrO2/C catalyst was studied by cyclic voltammetry (CV) and electrochemical impedance spectroscopy (EIS) in alkaline solution. Pt-ZrO2/C showed higher catalytic activity than Pt/C (20 wt% E-TEK). The optimum ratio for Pt:ZrO2 was 1:4 [50].
Multi-metallic Pt based catalysts were studied on the electrooxidation of ethylene glycol (EG) in alkaline medium [51]. The addition of Bi to platinum caused a positive shift of the onset potential for EG oxidation by 70 mV but an increase of current density. The ternary catalyst PtPdBi/C did not change the EG oxidation onset potential, and a higher current density was obtained compared to PtBi/C. The performance of these catalysts was evaluated with an alkaline membrane fuel cell. OCV and power density are summarised in Table 2.
Table 2. Fuel cell performance using ethylene glycol in alkaline medium at multi-metallic Pt based catalysts [51].
Table 2. Fuel cell performance using ethylene glycol in alkaline medium at multi-metallic Pt based catalysts [51].
CatalystOCV/VPower density/mWcm−2
Pt/C0.6619
PtBi/C0.8322
PtPdBi/C0.8128
Improved OCV and power density were obtained from binary and ternary catalysts. The catalysts containing Bi seem to activate the oxidation of EG in oxalate compared to Pt. It was proposed that in binary and ternary catalysts with Bi and Pd, Bi mainly favours the adsorption of OH but affects the product distribution by changing the composition of chemisorbed species; Pd can reduce the poisoning on Pt sites by changing the composition of chemisorbed species.
Palladium is a very good electrocatalyst for organic fuel electro-oxidation and is a potential alternative to platinum when alloyed with non-noble metals for use in DAFCs. Although Pd does not show comparable activity to Pt on methanol oxidation, it demonstrated good activity in several studies on 1-propanol, 2-propanol, ethanol, ethylene glycol and glycerol oxidation [32,52,53,54,55].
Pd binary catalysts with addition of Ag, Au or Ru showed improved activity compared to Pd catalyst alone. Pd/C promoted by Au, with the optimum ratio of Pd:Au = 4:1, demonstrated a significant increase of catalytic activity and stability for 2-propanol oxidation [56]. Bunazawa and Yamazaki also prepared PdAu/C catalyst for methanol fuel cells using ultrasonic synthesis and found that EG was the optimum solvent resulting in highly dispersed nanoparticles on carbon support [57]. Lamy et al. [58] also used carbon supported Au/Pd for a fuel cell with anion exchange membrane and glycerol as the fuel. The performance with Au/Pd was lower than using Au or Pd alone, and the highest power output was obtained with Pd/C catalysts. An enhancement was also observed from Pd-Ag/C catalyst towards ethanol electrooxidation in alkaline media [33]. The catalyst was prepared by co-reduction method and Pd-Ag alloy nanoparticles were formed. The Pd-Ag/C catalyst exhibited high activity, enhanced adsorbed CO tolerance and stability for ethanol oxidation. Addition of Ru has the same enhancement on Pd catalyst for ethanol oxidation [18]. With Pd-Ru catalyst, the ethanol oxidation current density was four times of that from Pt-Ru within the potential 0.3–0.4V (RHE). An et al. used Pd-Ni as the anode catalyst for a direct EG fuel cell with Tokuyama AEM and a non-Pt cathode catalyst [59]. A power density of 67 mWcm2 was achieved, which is the highest reported for EG fuel cells.
Ternary and quaternary Pt catalysts were studied for ethanol oxidation in alkaline by Bambagioni et al. [60]. Pd was spontaneously deposited on Vulcan XC-72 supported Ni-Zn and Ni-Zn-P alloys using PdIV salts. The Pd-(Ni-Zn)/C and Pd-(Ni-Zn-P)/C catalysts exhibited promising activity for ethanol oxidation. The onset potential was −0.6 V vs. Ag/AgCl/KClsat, i.e. −0.4 V vs. NHE, and a specific current of 3600 A (g Pd)1 was obtained on both catalysts. Higher activity and stability of the catalysts were due to interaction with oxygen atoms from Ni-O moieties.
Multi-walled carbon nanotubes (MWCNT) were used as the substrate for Pd and Pt-Ru catalysts in fuel cells with anion exchange membranes (AEM) using various alcohols as the fuel. The catalysts were very active for the oxidation of all alcohols [52,61]. The effect of a small addition of MWCNT (1–5%) and Ni (1%) into Pd catalyst was studied for methanol oxidation in 1M KOH [62]. Addition of 1% MWCNT in Pd catalyst led to a 5–8 fold increase of catalytic activity; while further addition of 1% Ni into Pd-MWCNT further increased activity by 1.4–1.7 fold. The increase of activity can be ascribed to improved geometrical and electronic properties of the catalyst.
Metal oxide promoted Pd/C catalysts were also investigated by Xu et al. for alcohol oxidation in alkaline media [63]. The addition of oxides like CeO2, NiO, Co3O4 and Mn3O4 significantly promoted the activity and stability of Pd/C catalysts. Pd-Co3O4 showed the highest activity for methanol, EG and glycerol oxidation; Pd-NiO was the most active for ethanol oxidation, and Pd-Mn3O4/C improved stability of catalysts.
Another important precious metal for alcohol oxidation in alkaline media is Au. It has been studied for methanol oxidation [64,65]. The mechanism and kinetic studies of ethanol oxidation on gold in alkaline media have been carried out by different groups [30,31]. Au was also used in methanol, ethanol and EG oxidation. Au:Ru (1:1) exhibited higher activity than Au [18].

3.1.2. Non-Precious Metal Catalysts

One of the most attractive advantages of alkaline fuel cells is the possibility of using non-precious metal catalysts. Nickel has been used as an electro-catalyst for both anodic and cathodic reactions in organic synthesis and water electrolysis. It has been studied for the electro-oxidation of alcohols for decades [66]. Van Effen and Evans [67] studied the oxidation of ethanol in KOH solution on nickel. They discovered that the reaction involved the formation of a higher valence nickel oxide, which acted as a chemical oxidizing agent. This fact was confirmed [68] by cyclic voltammetry, which proved the presence of a mediation process involving the higher oxides and the organic molecules, and by alternating current impedance measurements.
Taraszewska and Roslonek [19] found that a glassy carbon/Ni(OH)2 modified electrode was an effective catalyst for the oxidation of methanol. The use of Ni as a catalyst for the electro-oxidation of methanol in alkaline medium was studied using cyclic voltammetry by Abdel Rahim et al. [69]. Ni dispersed on graphite showed catalytic activity towards methanol oxidation but massive Ni did not. Ni was dispersed on graphite by the electro-deposition from acidic Ni2SO4 solution. The catalytic activity of the C/Ni electrodes towards methanol oxidation was found to vary with the amount of electro-deposited Ni. Methanol oxidation starts as Ni-oxide is formed on the electrode surface and indicates that the electro-oxidation of methanol is activation controlled, proceeding by direct chemical reaction with NiO(OH) for thin nickel oxides and by charge transfer with the electrode for thick oxides. The accumulation of NiO(OH) had an inhibiting effect on activity which can be counteracted by a period of re-activation in the hydrogen evolution potential region.
A Ni zeolite has been reported with a superior methanol oxidation activity to Ni and Pt supported on graphite [70]. This performance was assigned to a lower blocking of active sites than Pt due to repeated utilization.
Synthesis and characterisation of electrodeposited Ni–Pd alloy electrodes for methanol oxidation in alkaline electrolyte was studied by Kumar et al. [71]. Structural characterisation of the electrocatalysts showed that the Ni–Pd catalysts formed a complete solid solution in the alloy. The percentage shift in the d-spacing, calculated from X-ray diffraction, was in good agreement with the palladium percentages in the alloy. The catalysts were active for methanol oxidation in alkaline medium.
A Co-W alloy prepared by electroplating was evaluated by Shobba et al. for methanol oxidation in alkaline media [72]. Heat treatment was shown to improve oxidation performance with open circuit potentials of 0.918 V obtained.
P-M-C complex materials were prepared by polymer template coordination with metal-ions loaded with a carbon-reduction method. P stands for polymer; M = Fe, Co and Ni; C is carbon. The catalyst showed very high electrocatalytic activity and stability for ethanol oxidation in alkaline media [73].
Lanthanum, strontium oxides have been studied by several research groups: Raghuveer and Vishwanathan [74] investigated the electro-oxidation of methanol on bulk and nano crystalline La1.8Sr0.2CuO4 in 1 M KOH. Raghuveer et al. [75] investigated methanol oxidation on La1.8Sr0.2CuO4, La1.6Sr0.4CuO4 and La1.9Sr0.1Cu0.9Sb0.1O4 in 3.0 M KOH. Yu et al. [76] studied La0.75Sr0.25CuO3δ and La0.75Sr0.25CoO3δ in 1 M NaOH and 1 M CH3OH. All oxides gave reasonable oxidation currents in potential ranges appropriate to fuel cell operation.
Perovskite-type oxides have been investigated for methanol and EG alkaline fuel cells. Perovskite-type ternary oxides with molecular formulae of La2-xSrxNiO4 (0 ≤ x ≤ 1) were prepared by a modified citric acid sol-gel route for a direct methanol alkaline fuel cell (DMAFC). There was no poisoning of the catalyst detected by the methanol oxidation [77].
Platinum free nanosize RuM (M = Ni, Co and Fe) catalysts for glycerol oxidation in alkaline media were prepared on Vulcan XC72 from Ru(OH)3Cl3 and metal sulfates in a hydrogen atmosphere at 400 °C. Glycerol oxidation was tested in 8M KOH and 4M glycerol at 60 °C. With a catalyst loading of 30–40 mg cm−2, good activity for glycerol oxidation was observed [78].

3.2. Catalysts for Oxygen Reduction

For fuel cell performance, oxygen reduction is a kinetically limiting factor, particularly with hydrogen fuel cells. In direct alcohol fuel cells, fuel crossover to the cathode causes a mixed potential on the cathode and further decreases the oxygen reduction performance.

3.2.1. Precious Metal ORR Catalysts

Platinum, silver and palladium compounds have been used as ORR catalysts in DAAFCs. Platinum and Pt/Ru catalysts were used as the cathode catalysts for direct methanol fuel cells by Yu and Scott [79]. It was found that ORR activity was lower on Pt/Ru than on Pt due to the low activity of Ru in alkaline solution. The methanol tolerance of the Pt/C electrode with a Pt loading of 1.2 mg cm−2 was 0.2 M methanol in 1 M NaOH suggesting crossover may have less significant effect in alkaline.
A Ag/C catalyst was prepared by simple thermal decomposition of AgNO3 in argon at 700 °C by Gamburzev et al. [80] to use in AFCs. Kostowskyj et al. [81] produced Ag nanowires by coating Ag on carbon nanotubes. The electrochemical test results showed that the electrode activity increased with increasing Ag concentration when mixed with CNTs.
Highly selective Ag-W2C/C and W2C/C electrocatalysts for oxygen reduction were developed for potential application in mixed-reactant alcohol fuel cells by Meng et al. [82]. The catalysts were prepared by an intermittent microwave heating (IMH) method. Both the W2C/C and Ag-W2C/C showed activity for ORR in alkaline media. The introduction of W2C into the Ag/C enhanced its activity. The Ag-W2C/C catalyst was highly selective for oxygen reduction in the presence of alcohol.
Pd was also studied as cathode catalyst for DAAFC [57,83]. Although the activity of Pd is lower than Pt, the Pd/C catalyst is less sensitive to alcohol contamination as compared to Pt/C, indicating better alcohol tolerance. Kim et al. [84] used Pd-Sn/C as a cathode catalyst on DMAFC; this cathode exhibited higher performance than commercial Pt/C catalyst in high methanol concentration. It suggested good methanol tolerance with Pd-Sn/C catalyst indicating promising application on DAAFCs.

3.2.2. Non-Precious ORR Catalysts

One of the major advantages of using alkaline media for fuel cells is the potential of applying low cost non noble metals as the catalysts. A wide range of non-noble metal catalysts have been employed in DAAFCs.
Manganese oxides have attracted increasing interest as oxygen reduction catalysts in alkaline media in recent years. The activity of MnOx depends upon the type of oxide and the carbon support. The most active were the MnO2, Mn2O3 and MnOOH (manganese oxy-hydroxy) oxides [85,86,87]. The ORR mechanism on these catalysts involves the formation of the HO2 intermediate via the transfer of two electrons followed by a dismutation reaction of this species into OH and O2.
Metal macrocyles have been used as oxygen reduction catalysts since the 1960’s. They have shown high activity towards the ORR in alkaline media. Cobalt and iron phthalocyanine and porphyrins are of the most interest. Co phthalocyanine has exhibited similar kinetics on ORR as Pt [88]. Electrodes made of Cobalt/iron tetraphenylporphyrin (CoTPP, FeTPP) demonstrated good performance, even better than Ag catalyst. Heat treatment of carbon supports can result in structural changes and increase the surface area, resulting in enhanced catalytic activity [89,90]. However, the stability of the catalyst was poor due to loss of Co [91]. It has been reported that iron phthalocyanine (FePc) has the ability to promote the direct 4e oxygen reduction to water, while CoPc promotes O2 reduction to H2O2 by a 2e pathway [92,93,94,95]. This 4e pathway is more desirable as it avoids the production of hydrogen peroxide which can damage the electrode structure. One important advantage from MnOx and metal macrocycles to Pt is their insensitivity to fuel and tolerance to alcohol crossover.
Carbon supported La1-xSrxMnO3 (LSM/C) was prepared for oxygen reduction with the presence of EG [20]. LSM/C exhibited high activity for O2 reduction regardless of EG. This indicates LSM is a good cathode catalyst in DAAFC with no crossover problem. A Fe-Co nanoparticle based ORR catalyst, Hypermec TM K14 developed by Acta S.P.A, Cascina (Pisa), Italy, was reportedly applied by few groups as the cathode for their alkaline alcohol fuel cells [59,96,97,98]. Power densities ranging from 55–170 mW cm2 with different fuel cell configurations and fuels were achieved indicating a very promising catalyst for DAAFCs.

4. Electrolyte and Membrane Electrolyte for DAAFC

Most studies of direct alcohol fuel cells have used aqueous alkaline electrolyte or added alkaline in the fuel solutions, as an anolyte [21,49,99,100]. To date the performance of these fuel cells has not matched that achieved with the fuel cells using proton exchange membranes (PEMs) in acidic environment.

4.1. Cation Exchange Membranes

Polymer electrolyte materials such as Nafion® are used commercially both for proton transfer and for sodium ion transfer. An example is the electrolysis of sodium chloride to produce chlorine and sodium hydroxide. The conduction process in this system is realised by the transport of Na+ ions from anode to cathode. A practical consequence of DMAFCs operating in such a manner is that hydroxide ions produced by oxygen reduction react rapidly with Na+ ions leading to the formation of sodium hydroxide in the cathode side of the cell. In addition, water is transported across the cell to the cathode. This being the case, it is necessary to recycle this component to the anode to maintain an overall sodium ion balance in the system. A schematic diagram of the electrochemical reactions and transport processes in such a system is shown in Figure 3. The performance of a Na+ conducting Nafion based DMFC has been studied [101]. Although the open circuit voltage obtained was favourable (ca. 0.8 V), in comparison to DMFCs operating in acidic media, the polarisation of the DMAFC was greater [102]. An attraction of these solid polymer electrolytes (SPEs) is that the Na+ ion form of Nafion offers much greater stability than currently available OH ion (anion) conducting membranes. However, the disadvantage of the Na+ ion form of Nafion is that the ionic conductivity is much lower than that of the H+ ion form of Nafion, and anion conducting membranes for OH. This is determined by mobility of ions. The mobility of Na+ in water solution is seven times lower than that of H+ and four times lower than that of OH-. The internal resistance of the cell as measured by high frequency AC impedance was 0.47 Ω [101]. This value is much higher than that with H+ form Nafion, whose value is around or lower than 0.05 Ω [100]. Thus there will be a compromise between ionic conductivity and electrocatalyst activity at both anode and cathode. These factors affect the membrane, membrane-catalyst layer interface and the bounded electrocatalyst layers.
Figure 3. Schematic diagram of DMAFC with cation exchange membrane and incorporating an electrolyte circulating layer in the cathode compartment [101].
Figure 3. Schematic diagram of DMAFC with cation exchange membrane and incorporating an electrolyte circulating layer in the cathode compartment [101].
Energies 03 01499 g003

4.2. Anion Exchange Membranes (AEM)

Use of OH- conducting membranes or anion exchange membranes (AEM) can improve ionic conductivity and the fuel cell performance compared to use of CEMs. AEMs are solid polymer electrolyte membranes that contain positive ionic groups (e.g., quaternary ammonium (QA) functional groups such as poly—N+Me3) and mobile negatively charged anions (e.g., usually OH).
Some commercially available AEMs have been used for DAAFC applications. MORGANE®-ADP membranes from Solvay, S.A. were used as the solid polymer electrolyte for DMAFCs in several studies, which were normally used for salt electrodialysis [21,99,100,103]. Encouraging fuel cell performances were obtained, with the highest power density of 20 mW cm−2. The basic properties of the MORGANE®-ADP membrane at 25 °C are shown in Table 3 [99]. Methanol diffusion coefficients were measured by monitoring the transfer of methanol from pure methanol solution across the membrane into a NaOH solution. The methanol diffusion coefficient for Nafion is about 2.8 times higher than that for the ADP membrane. This indicates that the methanol crossover rate for ADP membrane will be potentially lower than Nafion, which may improve fuel cell performance. However, the ADP membrane is not stable in strong alkaline media. The suggested working pH is up to a value of pH 11. In fact, discolouring of the ADP membrane in 1 M NaOH was observed. This could affect the stability and the lifetime of a fuel cell with an ADP membrane.
Table 3. Basic properties of the MORGANE®-ADP membrane [21].
Table 3. Basic properties of the MORGANE®-ADP membrane [21].
MembraneMORGANE®-ADP
MaterialCross-linked fluorinated polymer
Exchange groupQuaternary ammonium
Thickness/μm fully humidified150–160
Resistance (in 0.6 M NaCl)/Ω.cm21.5–4.5
Resistance (in 1 M NaOH)/Ω0.5
Maximum operational temperature/°C55
Working pH0-10
Researches on using few ammonium-type AEM membranes from Tokuyama Co. Japan have been reported. A201 (developing code A-006) and A901 were used for alkaline fuel cells with ethanol [34,96,104] and AHA were used for DAAFC using other alcohols [13]. Basic properties of these membranes are given in Table 4. With A201 at 80 °C, the power density of 55 mW cm−2 supplied by a passive system and 170 mW cm−2 obtained from an active system were reported [96]. The maximum power density of 9 mW cm-2 was obtained from fuel cells using 1 M EG in 1M KOH at 25 °C and AHA [13]. Fumatech has also developed AEMs, but studies on applying these membranes are still undergoing and no report on these studies has been published.
Research on developing anion exchange membranes with high ionic conductivity and good chemical stability in alkaline media as well as low alcohol permeability have been carried out by several groups. Polyvinylidene fluoride (PVDF, -[CH2CF2]n-) and polytetrafluoroetheneco-hexafluoropropylene (FEP, -[CF2CF2]n[CF(CF3)CF2]m-) films grafted with vinylbenzyl chloride using radiation-grafting are compared [105,106]. The FEP membrane showed good properties suitable for the fuel cell application, and it is possible to use these new materials in DAAFC development. However, the cost of FEP is relatively high.
A number of other membrane materials have been reported for application in DMAFCs, such as polysiloxane containing quaternary ammonium groups [107], aminated poly(oxyethylene) methacrylates [108], quaternised polyethersulfone Cardo [109], and quaternised poly(phthalazinon ether sulfone ketone) [110]. The poor stability of these materials in hydroxide electrolytes limited their further application in DMAFC.
Table 4. Basic properties of Tokuyama membranes [13,34,97,98].
Table 4. Basic properties of Tokuyama membranes [13,34,97,98].
MembraneAHAA201A901
TypeStrong basic anion permeable
Electric Resistance (0.5 N NaCl)/Ω.cm24.10.30
Burst strength/MPa≥0.900.40.2
Exchange groupTetraalkyl ammonium groups with polyolefin backbone chainQuaternary ammoniumQuaternary ammonium
Thickness/µm2402810
Ion exchange capacity/mmol g-11.15–1.251.71.7
OH- conductivity/mS cm-2 2911.4
Poly(2,6-dimethyl-1,4-phenylene oxide) (PPO) is a unique material with strong hydrophobicity, a high glass transition temperature (Tg = 212 °C) and hydrolytic stability [111]. PPO based membranes were studied for DMAFC applications. Chloroacetylated poly(2,6-dimethyl-1,4-phenylene oxide) (CPPO) with bromomethylated poly(2,6-dimethyl-1,4-phenylene oxide) (BPPO) blend membranes were characterised [112]. Substituted materials have good miscibility since they all have PPO backbones. A blend membrane with 30–40 wt% CPPO was recommended. The membranes exhibited a high hydroxyl conductivity (up to 0.032 S/cm at 25 °C) and extremely low methanol permeability (1.26–1.04 × 10−7 cm2s−1), which is suitable for application in low temperature DMAFCs [113].
Quaternized polyethersulfone Cardo anion exchange membranes have been prepared [114]. Polyethersulfone Cardo was chloromethylated with the complex solution of chloromethylether and zinc chloride. Subsequent reaction with trimethylamine and ion exchange with sodium hydroxide yielded the alkaline anion exchange membrane. Ionic conductivities of QPES-C membrane in 1 M NaOH solution were 4.1 × 102 S cm−1 to 9.2 × 10−2 S cm−1 and methanol permeability was from 5.72 × 10−8 to 1.23 × 10−7 cm2 s−1 over the temperature range 25–70 °C indicating potential applications in DMAFC.
An alternative approach to the production of AEMs is the use of suitable doping/reaction of polymer films with KOH. Polybenzimidazole (PBI) membranes have attracted interests in recent years for high temperature PEMFC applications. An alternative approach to the production of AEMs is the use of suitable doping/reaction of polymer films with KOH. An alkali-doped PBI membrane showed higher conductivity than H2SO4-doped PBI and H3PO4-doped PBI at 25 degrees C [115].
Hou et al. used KOH-doped PBI as membrane for DAAFCs to be able to work at higher temperatures, i.e. at improved reaction kinetics. The PBI-based MEA yielded power densities of ca. 30 mW cm2 for methanol [116] and 60 mW cm2 for ethanol [117]. They also reported that methanol and ethanol diffusivity through the PBI membrane was significantly below that of Nafion. Modestov et al. reported a direct ethanol fuel cell using alkali doped PBI-membrane with Pt-free catalysts, namely 5,10,15,20-tetrakis(4-methoxyphenyl)-21H,23H-porphine cobalt(II) supported on Vulcan XC72 carbon for the cathode and RuV alloy supported on Vulcan XC72 for the anode. They achieved a maximum power density of 125 mW cm2 [118]. These reports suggest a possibility to develop high temperature DAAFC which could significantly improve the reaction activity on alcohol oxidation.
Non perfluorinated membranes have attracted interests because of their low cost. Yang et al. studied polyvinyl alcohol (PVA) based membranes for use in alkaline alcohol fuel cells. PVA crosslinked with sulfosuccinic acid (SSA) [119], PVA/hydroxyapatite (HAP) composite membrane [120] and PVA/TiO composite membrane [121] were used in methanol, ethanol, isopropanol and borohydride fuel cells with the presence of alkaline (from 2 M KOH up to 8 M KOH) in fuel solutions. The membranes showed good electrochemical performance. A quaternised poly(vinyl alcohol)/alumina (QPVA/Al2O3) nanocomposite polymer membrane was prepared by a solution casting method by Yang et al. [122]. In an alkaline DMFC with the PVA/Al2O3 composite membrane, a peak power density of 36.2 mW cm2 was achieved at room temperature with ambient air.

5. Performance of DAAFC

The performance of one of the first alkaline alcohol fuel cells was reported by Cairns and Bartosik in 1964 [123]. They operated the cell with vapourised methanol with an invariant alkaline electrolyte operated at 115–130 °C. The conversion of the fuel to electricity in one pass through the cell was 75 to 85%. Power densities of 40–45 mW cm2 were achieved. However, there was a gap on further research on alkaline alcohol fuel cells until this decade.
Most DAAFC performances reported earlier were lower than those of the corresponding the PEM counterpart. The power density of current DMAFCs is typically less than 20% of that obtained from PEM DMFC. The DAAFC studies focused on PtRu and Pt as catalysts; the highest performance reported with such materials was 58 mW cm2 [34]. This has been changed with new advances in AEM and catalyst development in the past few years. Power density of 160 mW cm2 was achieved at 80 °C by Bianchini et al. [96] from an alkaline fuel cell using ethanol. This is comparable to what is achieved from PEM DMFCs. Even with a passive cathode, power density of 55 mW cm2 was obtained using ethanol. Bianchini and coworkers also tested the system with various fuels (methanol, ethanol, glycerol). The wide range of materials used in DAAFC suggest the potential of using low cost materials in DAAFCs. Alkaline alcohol fuel cell performances reported for different alcohol, catalysts, AEMs and operating parameters are summarized in Table 5.
In most studies, alkaline liquid electrolyte was still added to the fuel despite use of AEMs. Only very few studies used liquid fuel without adding alkaline electrolyte. Varcoe and Slade [124] developed a novel quaternary-ammonium-functionalised radiation-grafted ETFE alkaline anion-exchange membrane (AAEM) and applied M+OH free methanol in a methanol/O2 fuel cell. A maximum power density of 8.5 mW cm2 was obtained at 80 °C. Coutanceau et al. also investigated AFCs using Ag/C catalyst for ORR and Pt-Pd anode catalyst with methanol and EG [125]. It was clear that fuel cell performance improved significantly with the addition of NaOH in EG.
Figure 4 compares the DMAFC performance with and without additional alkali in the fuel. With similar catalyst loading and the same type of membrane (ADP), the cell with 1.0 M NaOH in the fuel gave higher current and power density [21,23,96,103]. Without 1 M NaOH, the maximum current density was only around 60 mA cm2, while with alkali, the maximum current density increased to 100 mA cm2; the peak power density also improved from 9 mW cm2 to 18 mW cm2. The OCV without NaOH was about 50 mV (0.65 V) lower than with NaOH in the fuel (0.70 V).
Figure 4. Fuel cell polarisation curves and power density curves for the DMAFC using air at 60 °C, air pressure 1 bar: (a) 1 M methanol with 1 M NaOH [21]; (b) 1 M methanol [103].
Figure 4. Fuel cell polarisation curves and power density curves for the DMAFC using air at 60 °C, air pressure 1 bar: (a) 1 M methanol with 1 M NaOH [21]; (b) 1 M methanol [103].
Energies 03 01499 g004
Although additional liquid electrolyte increases the power density of DAAFCs, the use of alkaline electrolyte in the fuel solution is not favoured from the fuel cell system point of view. The presence of free cations such as K+ and Na+ allows for the unwanted precipitation of salts such as carbonates. This could limit the life time of an alkaline fuel cell. It should be also mentioned that the removal of non-gaseous products such as salts and oxygenates is not trivial in fuel cell systems.
Figure 5 depicts the relationships between reported OCVs and maximum power densities of DAAFCs to alkaline concentration. It can be observed that OCV and power density both increased with alkaline concentration until maximum values were reached. OCV and power density were not affected by further increase in alkaline concentration beyond that indicating a limited effect of high alkaline concentration on fuel cell performance.
Figure 5. (a) OCV and (b) maximum power density of DAAFCs as a function of alkaline concentration for various fuels. Methanol [126], ethanol [34], EG [51] at room temperature, and EG at 60 °C [59].
Figure 5. (a) OCV and (b) maximum power density of DAAFCs as a function of alkaline concentration for various fuels. Methanol [126], ethanol [34], EG [51] at room temperature, and EG at 60 °C [59].
Energies 03 01499 g005
DAAFCs with a self-breathing or passive cathode have been investigated by several groups because of its practicality and convenience [96,126,127]. For ethanol, a power density of 55 mW cm2 was reported [96]; for methanol, a maximum power density of 12.8 mW cm2 was obtained with 7 M methanol at 25 °C with KOH in the fuel [126].
Other fuel cell configurations, such as laminar flow fuel cells [128] and mixed reactant fuel cells [129,130] have also been applied to DAAFCs. Choban et al. developed a laminar flow-based single microfluidic channel fuel cell, where an acidic anolyte containing methanol and an alkaline catholyte containing dissolved oxygen were merged in a microfluidic channel which contained electrodes at it’s walls. The streams were fed in such a way that the anolyte was in contact only with the anode and catholyte with the cathode [105]. Both streams mixed only slowly due to operation in the laminar regime and at high Peclet numbers. A novel micro fabrication method, multistream laminar flow, was used. The maximum power density of 12 mW cm2 was achieved with a mixed media: 1N KOH in anode and 1N H2SO4 in cathode. Sung et al. [107] developed a micro fuel cell using non-noble metal catalysts in alkaline media. The fuel cell was designed with both electrodes not facing each other, but being in-plane. Nickel hydroxide and silver oxide was the catalyst for anode and cathode respectively. The fuel (methanol) and the oxidant (H2O2) were both mixed into the same electrolyte, 1M KOH. Although the maximum power density obtained was only 28 µWcm2, this new technology enables high design flexibility and easy integration of the microscale fuel cell into microfluidic systems and portable devices. A more conventional approach for mixed reactant DAAFC resembling that of early studies [129] was evaluated by Zeng et al. [131]: a porous separator for NiMH cells was assembled with a Pt and a Pt/Ru gas diffusion electrode, and a mixed flow of ethanol, air and KOH were supplied to the fuel cell. The maximum power density was 35 mW cm2 and the selectivity of the electrodes was tuned by adjusting hydrophobicity of the catalyst layers.
Table 5. Summary of DAAFC performance.
Table 5. Summary of DAAFC performance.
Fuel/oxidantCatalystsElectrolyte/membraneT/°COCV/VImax/mA cm−2Pmax/mW cm−2Ref.
AnodeCathode
DMAFC w/o alkaline
2 M methanol/O2
no backpressure
no backpressure
2.5 bar back pressure
2.5 bar back pressure
Pt
4 mg cm2
Pt
4 mg cm2
Quaternised
radiation-grafted
ETFE AAEM

50
60
60
80

0.46
0.52
0.58
0.63

22
34
34
68

1.5
2.4
4.2
8.5

[123]
1 M methanol/O2Pt/Ru
1 mg cm2
Pt/C
Pd/C
.
0.5 mg cm2
Tokuyama
A201
800.57
0.64
34
34
2.6
2.6
[57]
2 M methanol/airPt/Ru
1 mg cm2
Pt
1 mg cm2
ADP30
40
50
60
0.6
0.62
0.65
0.65
32
43
57
60
5.9
6.9
7.6
9.0
[102]
16% methanol in N2/airPtRuPtPVA+10M KOH
PVA+10M KOH/Ni-LDH
400.980

100
22

35
[135]
Fuel/oxidantCatalystsElectrolyte/membraneT/°COCV/VImax/mA cm−2Pmax/mW cm−2Ref.
DMAFC w/alkaline
1 M methanol/O2
no NaOH
0.25 M NaOH
1 M NaOH
4 M NaOH (2 M methanol)
Pt/C
2 mg cm−2
Pt/C
2 mg cm−2
ADPRoom T
0.48
0.52
0.6
0.6

3
10
52
110

0.2
1
8
18

[125]
2 M methanol, 1 M NaOH/airPt/C
2 mg cm2
Pt/C
2 mg cm2
ADP600.710018[21]
2 M methanol, 1 M NaOH/ airPt/C
2 mg cm2
Pt/C
2 mg cm2
Nafion600.80234.5[101]
7 M methanol, 1 M KOH/ air (passive)PtRu
4 mg cm−2
Pt/C
1 mg cm−2
Tokuyama
Room T0.715812.8[126]
2 M methanol, 2 M KOH/ O2PtRu/C
2 mg cm−2
Pt/C
1 mg cm−2
PBI/KOH901.010530[116]
2 M methanol, 2 M KOH/ O2PtRu/C
2 mg cm−2
Pt/C
1 mg cm−2
PBI/KOH75
90
0.92
0.98
150
190
49
61
[117]
1 M methanol, 1 M KOH/saturated O2 in 1 N H2SO4PtRu
2 mg cm−2
Pt
2 mg cm−2
laminar flow-
based micro fuel cells
Room T1.45012[125]
4 M methanol, 4 M KOH/ airPt/Ru
4 mg cm−2
MnO2/C
4 mg cm−2
QPVA/Al2O3Room T0.8815336[122]
DAAFC w/ alkaline
1 M KOH /humidified O2
methanol
ethylene glycol
glycerol
erythritol
xylitol
Pt/C
1 mg cm2
Ag/C
1 mg cm2 or PtRu/C
4 mg cm2
AHA500.80
46
65
55
42
29

6.0
9.0
6.8
5.5
4.0
[13]
1 M ethanol, KOH/
humidified O2
0.1 M KOH
0.5 M KOH
1.0 M KOH
PtRu
3 mg cm2
Pt
3 mg cm2
AHARoom T

1.17
0.83
0.84


100
350
400


18
58
58
[34]
4 M KOH/air
2 M methanol
2 M ethanol
2 M isopropanol
Pt/Ru
3.6 mg cm−2
MnO2/C
PVA/TiO2 composite membraneRoom T0.80
33.6
35.7
21.5

9.3
8.0
5.5
[121]
2 M glycerol, 4 M NaOH/ O2Pt/C
Pd/C
Au/C AuPd/C
Pt/C
2 mg cm2
ADPRoom T0.68
0.59
0.60
0.49
20
16
7.5
3.2
4.2
2.4
1.0
0.3
[58]
2 M EG, 4 M NaOH/ O2Pt
PtBi
PtPdBi
2 mg cm2
Pt
2 mg cm2
ADP200.66
0.83
0.81
132
132
120
19
22
28
[51]
2 M EG, NaOH/O2
1 M NaOH
2 M NaOH
4 M NaOH
Pt
2 mg cm2
Pt
2 mg cm2
ADP20
0.54
0.63
0.66

42
102
132

4
13
19
[51]
Fuel/oxidantCatalystsElectrolyte/membraneT/°COCV/VImax/mA cm−2Pmax/mW cm−2Ref.
2 M EG, 4 M NaOH /O2Pt/C
2 mg cm−2
Pt/C
2 mg cm−2
ADPRoom T0.6513018[125]
0.5 M ethanol, air, 2 M KOHPtRu/C
1.8 mg cm−2
Pt/C
2 mg cm2
Porous separatorRoom T1.011030[128]
DAAFC with non Pt catalyst, with alkaline
2 M KOH/ O2
10 wt% methanol
10 wt% ethanol
5 wt% glycerol
Pd/MWCNT
1 mg cm2
Fe-Co HypermecTM K-14
Tokuyama
A-210
80
0.89
0.87
0.87

530
350
550

95
74
79
[61]
2 M KOH/O2
10 wt% methanol
10 wt% ethanol
5 wt% glycerol
PdNiZn/C
1 mg cm2
Fe-Co HypermecTM K-14Tokuyama
A-210
80
0.76
0.88
0.81

610
800
540

120
165
119
[61]
2 M methanol, 0.05 M H2O2, 0.2M KOHNi(OH)2Ag2O-Room T0.1110.03[127]
10 wt% ethanol, 2 M KOH / O2PdNiZn/C
1 mg cm2
Fe-Co HypermecTM K-14Tokuyama
A-210
25
40
60
80
0.65
0.68
0.74
0.82
400
440
550
590
65
80
120
160
[61]
1 M ethanol, 1 M KOH/O2HypermecTM
2 mg cm−2
HypermecTM
1 mg cm−2
Tokuyama30
40
50
60
0.66
0.72
0.76
0.91
170
200
230
270
12
17
22
30
[98]
2 M ethanol, 3 M KOH/airRuV/C
4.5 mg cm−2
TMPhP/C 9mg cm−2PBI/KOH800.93278125[118]
1 M EG, KOH /O2
1 M KOH
5 M KOH
7 M KOH
9 M KOH
PdNi
2 mg cm−2
HypermecTM
1 mg cm−2
Tokuyama60
0.65
0.8
0.86
0.87

320
350
390
370

35
53
67
60
[59]

6. Summary and Outlook

Liquid alcohols have significantly higher energy densities than hydrogen, and they are easier to handle. These features make direct alcohol fuel cells a promising alternative low-cost power supply for portable electronics. Direct alkaline alcohol fuel cells can benefit from the more active reaction in alkaline media. Unlike conventional proton exchange membrane fuel cells, anion exchange membranes are used in present DAAFCs to transfer OH within the membrane electrolyte. This will also reduce the carbonation process which shortens the fuel cell life time due to the reaction between CO2 and alkaline electrolyte, particularly with liquid electrolyte.
Although the power output of DAAFCs is still lower than that of direct methanol fuel cells using proton exchange membranes, the performance of DAAFCs has improved significantly in the past decade. The developments on catalysts and anion exchange membranes provide a solid base for the further development of DAAFCs.
Various alcohol fuels have been investigated for DAAFC applications. Alkaline media has particular advantages for poly carbon alcohols over acidic media because of higher reaction activity in alkali. Apart from methanol, ethanol, isopropanol, ethylene glycol and glycerol have also shown encouraging fuel cell performances. These alcohols have higher energy density than methanol, and are not toxic, which make them potential candidates as power sources for portable electronic devices. However, due to the higher energy to break C-C bonds during the oxidation, no complete oxidation to CO2 from polycarbon alcohols was achieved with current catalysts. Research on developing more active catalysts is a significant challenge. Besides most used precious metal Pt and Pt alloys, other lower cost metal catalysts, such as Pd and Ni and metal alloys have exhibited potential to be used as alternatives for Pt catalysts. Lanthanum, strontium oxides and perovskite-type oxides have also been investigated for alcohol oxidation.
Fuel crossover from the anode to cathode is a factor that limits fuel cell performance because of the mixed potential caused by fuel oxidation on the cathode. This effect causes a significant loss in fuel cell performance. However fuel crossover is significantly less in DAAFCs due to the direction of electric osmotic drag from cathode to anode. At present, Pt is still the most popular cathode catalyst for a direct alcohol fuel cell although Pt is highly sensitive to alcohols. Several catalysts have been developed and exhibited less sensitivity on alcohol oxidation. Examples are Pd- or Ag-based catalysts, as well as non-precious metal catalysts based on Mn-oxides.
Rather than electrolytes such as KOH and NaOH, solid electrolyte membranes have been used more and more in DAAFCs. Various types of anion exchange membranes, e.g. polytetrafluoroetheneco-hexafluoropropylene or poly(2,6-dimethyl-1,4-phenylene oxide), have been developed and investigated in alkaline fuel cells.
Direct alcohol anion exchange membrane fuel cells are still at early stages of development. In order to improve the power output from DAAFCs, and produce viable alkaline fuel cell systems as portable power sources, further research and development efforts need to be focused on novel catalysts and anion exchange membrane development:
(1)
Metal oxides as the promoter for catalysts.
Some metal oxides, in particular CeO2, Mn3O4 and SnO2, have multiple oxidation states and can store and release oxygen [132]. These metal oxides can be used as promoters to improve catalyst activity and fuel cell performance.
(2)
Transition metal macrocycle based oxygen reduction catalysts with improved activity and selectivity.
These catalysts have shown highly selective catalytic activity for oxygen reduction in the presence of methanol and CO in direct methanol fuel cells (DMFC) and hydrogen fuel cells [133,134]. Metal macrocyclic catalysts are stable in neutral and alkaline media. This suggests their application will be more feasible for DAAFCs.
(3)
Novel anion exchange membranes with improved stability and ionic conductivity.
One of the key issues in the development of DAAFCs is the membrane electrolyte. There is still no AEM equivalent to Nafion for solid polymer electrolyte alkaline fuel cells, even though significant progress has been made in the last decade. Further investigation on poly(2,6-dimethyl-1,4-phenylene oxide) (PPO) based membranes, alkaline doped AEMs and composite AEMs might be possible to produce a novel AEM with improved properties suitable for DAAFC applications.
Direct alkaline alcohol fuel cell technology is a promising technology as a clean energy sources for portable electronic devices. New material developments on catalysts and membranes will further improve the power output and make DAAFCs viable power sources in the near future.

Acknowledgement

Eileen Yu thanks EPSRC for a research fellowship (EP/C535456/1) to carry out this work.

References

  1. Bockris, J.O.; Conway, B.E.; White, R.E. Modern Aspects of Electrochemistry; Plenum Press: New York, NY, USA, 2001; Volume 34, p. 200. [Google Scholar]
  2. Zhang, X.Y.; Lu, W.; Da, J.Y.; Wang, H.T.; Zhao, D.Y.; Webley, P.A. Porous platinum nanowire arrays for direct ethanol fuel cell applications. Chem. Commun. 2009, 8, 195–197. [Google Scholar] [CrossRef]
  3. Varela, F.J.R.; Savadogo, O. Ethanol-tolerant Pt-alloy cathodes for direct ethanol fuel cell (DEFC) applications. Asia-Pac. J. Chem. Eng. 2009, 4, 17–24. [Google Scholar] [CrossRef]
  4. Neto, A.O.; Linardi, M.; dos Anjos, D.M.; Tremiliosi, G.; Spinace, E.V. Electro-oxidation of ethanol on PtSn/CeO2-C electrocatalyst. J. Appl. Electrochem. 2009, 39, 1153–1156. [Google Scholar] [CrossRef]
  5. Ling, J.; Longtin, G.; Savadogo, O. Comparison of ethanol and methanol crossover through different MEA components and structures by cyclic voltammetry. Asia-Pac. J. Chem. Eng. 2009, 4, 25–32. [Google Scholar] [CrossRef]
  6. Xu, C.W.; Cheng, L.Q.; Shen, P.K.; Liu, Y.L. Methanol and ethanol electrooxidation on Pt and Pd supported on carbon microspheres in alkaline media. Electrochem. Commun. 2007, 9, 997–1001. [Google Scholar] [CrossRef]
  7. Antolini, E. Catalysts for direct ethanol fuel cells. J. Power Source. 2007, 170, 1–12. [Google Scholar] [CrossRef]
  8. Qi, Z.G.; Kaufman, A. Performance of 2-propanol in direct-oxidation fuel cells. J. Power Source. 2002, 112, 121–129. [Google Scholar] [CrossRef]
  9. Qi, Z.G.; Hollett, M.; Attia, A.; Kaufman, A. Low temperature direct 2-propanol fuel cells. Electrochem. Solid State Lett. 2002, 5, A129–A130. [Google Scholar] [CrossRef]
  10. Chetty, R.; Scott, K. Catalysed titanium mesh electrodes for ethylene glycol fuel cells. J. Appl. Electrochem. 2007, 37, 1077–1084. [Google Scholar] [CrossRef]
  11. Neto, A.O.; Linardi, M.; Spinace, E.V. Electro-oxidation of ethylene glycol on PtSn/C and PtSnNi/C electrocatalysts. Ionics 2006, 12, 309–313. [Google Scholar] [CrossRef]
  12. Neto, A.O.; Vasconcelos, T.R.R.; Da Silva, R.; Linardi, M.; Spinace, E.V. Electro-oxidation of ethylene glycol on PtRu/C and PtSn/C electrocatalysts prepared by alcohol-reduction process. J. Appl. Electrochem. 2005, 35, 193–198. [Google Scholar] [CrossRef]
  13. Matsuoka, K.; Iriyama, Y.; Abe, T.; Matsuoka, M.; Ogumi, Z. Alkaline direct alcohol fuel cells using an anion exchange membrane. J. Power Source. 2005, 150, 27–31. [Google Scholar] [CrossRef]
  14. Arechederra, R.L.; Treu, B.L.; Minteer, S.D. Development of glycerol/O-2 biofuel cell. J. Power Source. 2007, 173, 156–161. [Google Scholar] [CrossRef]
  15. Gulzow, E. Alkaline fuel cells: A critical view. J. Power Source. 1996, 61, 99–104. [Google Scholar] [CrossRef]
  16. Wang, Y.; Li, L.; Hu, L.; Zhuang, L.; Lu, J.; Xu, B. A feasibility analysis for alkaline membrane direct methanol fuel cell: thermodynamic disadvantages versus kinetic advantages. Electrochem. Commun. 2003, 5, 662–666. [Google Scholar] [CrossRef]
  17. Adzic, R.R.; Avramovivic, M.L.; Tripkovic, A.V. Structural Effects in Electrocatalysis—Oxidation of Formaldehyde on Gold and Platinum Single-Crystal Electrodes in Alkaline-Solution. Electrochim. Acta 1984, 29, 1353–1357. [Google Scholar] [CrossRef]
  18. Chen, Y.G.; Zhuang, L.; Lu, J.T. Non-Pt anode catalysts for alkaline direct alcohol fuel cells. Chin. J. Catal. 2007, 28, 870–874. [Google Scholar] [CrossRef]
  19. Taraszewska, J.; Roslonek, G. Electrocatalytic oxidation of methanol on a glassy carbon electrode modified by nickel hydroxide formed by ex situ chemical precipitation. J. Electroanal. Chem. 1994, 364, 209–213. [Google Scholar] [CrossRef]
  20. Miyazaki, K.; Sugimura, N.; Matsuoka, K.; Iriyama, Y.; Abe, T.; Matsuoka, M.; Ogumi, Z. Perovskite-type oxides La1-xSrxMnO3 for cathode catalysts in direct ethylene glycol alkaline fuel cells. J. Power Source. 2008, 178, 683–686. [Google Scholar] [CrossRef]
  21. Yu, E.H.; Scott, K. Direct methanol alkaline fuel cell with catalysed metal mesh anodes. Electrochem. Commun. 2004, 6, 361–365. [Google Scholar] [CrossRef]
  22. Beden, B.; Leger, J.M.; Lamy, C. Modern Aspects of Electrochemistry; Bockris, J.O’M., Conway, B.E., White, R.E., Eds.; Plenum Press: New York, NY, USA, 1992; Volume 22, p. 97. [Google Scholar]
  23. Bagotzky, V.S.; Vassilyev, Y.B. Mechanism of electro-oxidation of methanol on the platinum electrode. Electrochim. Acta 1967, 12, 1323–1343. [Google Scholar] [CrossRef]
  24. Beden, B.; Kadirgan, F.; Lamy, C.; Leger, J.M. Oxidation of methanol on a platinum electrode in alkaline medium: Effect of metal ad-atoms on the electrocatalytic activity. J. Electroanal. Chem. 1982, 142, 171–190. [Google Scholar] [CrossRef]
  25. Lamy, C.; Leger, J.M.; Clavilier, J. Structural effects in the electrooxidation of methanol in alkaline medium: Comparison of platinum single crystal and polycrystalline electrodes. J. Electroanal. Chem. 1982, 135, 321–328. [Google Scholar] [CrossRef]
  26. Morallón, E.; Vázquez, J.L.; Aldaz, A. Electrochemical behaviour of basal single crystal Pt electrodes in alkaline medium. J. Electroanal. Chem. 1990, 288, 217–228. [Google Scholar] [CrossRef]
  27. Tripkovic, A.V.; Popovic, K.D.; Momcilovic, J.D.; Draic, D.M. Kinetic and mechanistic study of methanol oxidation on a Pt(111) surface in alkaline media. J. Electroanal. Chem. 1996, 418, 9–20. [Google Scholar] [CrossRef]
  28. Tripkovic, A.V.; Popovic, K.D.; Momcilovic, J.D.; Drazic, D.M. Kinetic and mechanistic study of methanol oxidation on a Pt(100) surface in alkaline media. J. Electroanal. Chem. 1998, 448, 173–181. [Google Scholar] [CrossRef]
  29. Tripkovic, A.V.; Popovic, K.Ð.; Momcilovic, J.D.; Drazic, D.M. Kinetic and mechanistic study of methanol oxidation on a Pt(110) surface in alkaline media. Electrochim. Acta 1998, 44, 1135–1145. [Google Scholar] [CrossRef]
  30. Tremiliosi-Filho, G.; Gonzalez, E.R.; Motheo, A.J.; Belgsir, E.M.; Léger, J.M.; Lamy, C. Electro-oxidation of ethanol on gold: Analysis of the reaction products and mechanism. J. Electroanal. Chem. 1998, 444, 31–39. [Google Scholar] [CrossRef]
  31. De Lima, R.B.; Varela, H. Catalytic oxidation of ethanol on gold electrode in alkaline media. Gold Bull. 2008, 41, 15–22. [Google Scholar] [CrossRef]
  32. Liang, Z.X.; Zhao, T.S.; Xu, J.B.; Zhu, L.D. Mechanism study of the ethanol oxidation reaction on palladium in alkaline media. Electrochim. Acta 2009, 54, 2203–2208. [Google Scholar] [CrossRef]
  33. Nguyen, S.T.; Law, H.M.; Nguyen, H.T.; Kristian, N.; Wang, S.; Chan, S.H.; Wang, X. Enhancement effect of Ag for Pd/C towards the ethanol electro-oxidation in alkaline media. Appl. Catal. B Environ. 2009, 91, 507–515. [Google Scholar] [CrossRef]
  34. Fujiwara, N.; Siroma, Z.; Yamazaki, S.-I.; Ioroi, T.; Senoh, H.; Yasuda, K. Direct ethanol fuel cells using an anion exchange membrane. J. Power Source. 2008, 185, 621–626. [Google Scholar] [CrossRef]
  35. Fang, X.; Wang, L.Q.; Shen, P.K.; Cui, G.F.; Bianchini, C. An in situ Fourier transform infrared spectroelectrochemical study on ethanol electrooxidation on Pd in alkaline solution. J. Power Source. 2010, 195, 1375–1378. [Google Scholar] [CrossRef]
  36. Cui, G.F.; Song, S.Q.; Shen, P.K.; Kowal, A.; Bianchini, C. First-Principles Considerations on Catalytic Activity of Pd toward Ethanol Oxidation. J. Phys. Chem. C 2009, 113, 15639–15642. [Google Scholar] [CrossRef]
  37. Matsuoka, K.; Iriyama, Y.; Abe, T.; Matsuoka, M.; Ogumi, Z. Electro-oxidation of methanol and ethylene glycol on platinum in alkaline solution: Poisoning effects and product analysis. Electrochim. Acta 2005, 51, 1085–1090. [Google Scholar] [CrossRef]
  38. Roquet, L.; Belgsir, E.M.; Léger, J.M.; Lamy, C. Kinetics and mechanisms of the electrocatalytic oxidation of glycerol as investigated by chromatographic analysis of the reaction products: Potential and pH effects. Electrochim. Acta 1994, 39, 2387–2394. [Google Scholar] [CrossRef]
  39. Schell, M.; Xu, Y.; Zdraveski, Z. Mechanism for the Electrocatalyzed Oxidation of Glycerol Deduced from an Analysis of Chemical Instabilities. J. Phys. Chem. 1996, 100, 18962–18969. [Google Scholar] [CrossRef]
  40. Damjanovic, A.; Genshaw, M.A.; Bockris, J.O.M. The Mechanism of Oxygen Reduction at Platinum in Alkaline Solutions with Special Reference to H[sub 2]O[sub 2]. J. Electrochem. Soc. 1967, 114, 1107–1112. [Google Scholar] [CrossRef]
  41. Damjanovic, A.; Sepa, D.B.; Vojnovic, M.V. New Evidence Supports the Proposed Mechanism for O2 Reduction at Oxide Free Platinum-Electrodes. Electrochim. Acta 1979, 24, 887–889. [Google Scholar] [CrossRef]
  42. Hsueh, K.L.; Gonzalez, E.R.; Srinivasan, S.; Chin, D.T. Effects of Phosphoric Acid Concentration on Oxygen Reduction Kinetics at Platinum. J. Electrochem. Soc. 1984, 131, 823–828. [Google Scholar] [CrossRef]
  43. Damjanovic, A.; Brusic, V. Electrode kinetics of oxygen reduction on oxide-free platinum electrodes. Electrochim. Acta 1967, 12, 615–628. [Google Scholar] [CrossRef]
  44. Damjanovic, A.; Genshaw, M.A.; Bockris, J.O.M. The Role of Hydrogen Peroxide in Oxygen Reduction at Platinum in H2SO4 Solution. J. Electrochem. Soc. 1967, 114, 466–472. [Google Scholar] [CrossRef]
  45. Beden, B.; Kadirgan, F.; Kahyaoglu, A.; Lamy, C. Electrocatalytic oxidation of ethylene glycol in alkaline medium on paltinum-gold alloy electrodes modified by underpotential deposition of lead adatoms. J. Electroanal. Chem. 1982, 135, 329–334. [Google Scholar] [CrossRef]
  46. Bi, Y.P.; Lu, G.X. Control growth of uniform platinum nanotubes and their catalytic properties for methanol electrooxidation. Electrochem. Commun. 2009, 11, 45–49. [Google Scholar] [CrossRef]
  47. Lović, J. The kinetics and mechanism of methanol oxidation on Pt and PtRu catalysts in alkaline and acid media. J. Serb. Chem. Soc. 2007, 72, 709–712. [Google Scholar] [CrossRef]
  48. Tripkovic, A.V.; Popovic, K.D.; Grgur, B.N.; Blizanac, B.; Ross, P.N.; Markovic, N.M. Methanol electrooxidation on supported Pt and PtRu catalysts in acid and alkaline solutions. Electrochim. Acta 2002, 47, 3707–3714. [Google Scholar] [CrossRef]
  49. Verma, A.; Basu, S. Direct alkaline fuel cell for multiple liquid fuels: Anode electrode studies. J. Power Source. 2007, 174, 180–185. [Google Scholar] [CrossRef]
  50. Bai, Y.X.; Wu, J.J.; Xi, J.Y.; Wang, J.S.; Zhu, W.T.; Chen, L.Q.; Qiu, X.P. Electrochemical oxidation of ethanol on Pt-ZrO2/C catalyst. Electrochem. Commun. 2005, 7, 1087–1090. [Google Scholar] [CrossRef]
  51. Demarconnay, L.; Brimaud, S.; Coutanceau, C.; Leger, J.M. Ethylene glycol electrooxidation in alkaline medium at multi-metallic Pt based catalysts. J. Electroanal. Chem. 2007, 601, 169–180. [Google Scholar] [CrossRef]
  52. Zheng, H.T.; Li, Y.L.; Chen, S.X.; Shen, P.K. Effect of support on the activity of Pd electrocatalyst for ethanol oxidation. J. Power Source. 2006, 163, 371–375. [Google Scholar] [CrossRef]
  53. Xu, C.W.; Tian, Z.Q.; Chen, Z.T.; Jiang, S.P. Pd/C promoted by Au for 2-propanol electrooxidation in alkaline media. Electrochem. Commun. 2008, 10, 246–249. [Google Scholar] [CrossRef]
  54. Ye, J.; Liu, J.; Xu, C.; Jiang, S.P.; Tong, Y. Electrooxidation of 2-propanol on Pt, Pd and Au in alkaline medium. Electrochem. Commun. 2007, 9, 2760–2763. [Google Scholar] [CrossRef]
  55. Liu, J.; Ye, J.; Xu, C.; Jiang, S.P.; Tong, Y. Electro-oxidation of methanol, 1-propanol and 2-propanol on Pt and Pd in alkaline medium. J. Power Source. 2008, 177, 67–70. [Google Scholar] [CrossRef]
  56. Xu, C.; Shen, P.K.; Liu, Y. Ethanol electrooxidation on Pt/C and Pd/C catalysts promoted with oxide. J. Power Source. 2007, 164, 527–531. [Google Scholar] [CrossRef]
  57. Bunazawa, H.; Yamazaki, Y. Ultrasonic synthesis and evaluation of non-platinum catalysts for alkaline direct methanol fuel cells. J. Power Source. 2009, 190, 210–215. [Google Scholar] [CrossRef]
  58. Lamy, C.; Coutanceau, C.; Leger, J.-M. Electrocatalytic Oxidation of Glycerol in a Solid Alkaline Membrane Fuel Cell (SAMFC). In Proceedings of ECS 216th Meeting, Vienna, Austria, 4–9 October, 2009.
  59. An, L.; Zhao, T.S.; Shen, S.Y.; Wu, Q.X.; Chen, R. Performance of a direct ethylene glycol fuel cell with an anion-exchange membrane. Int. J. Hydrogen Energ. 2010, 35, 4329–4335. [Google Scholar] [CrossRef]
  60. Bambagioni, V.; Bianchini, C.; Filippi, J.; OberhauserIal, W.; Marchionni, A.; Vizza, F.; Psaro, R.; Sordelli, L.; Foresti, M.L.; Innocenti, M. Ethanol Oxidation on Electrocatalysts Obtained by Spontaneous Deposition of Palladium onto Nickel-Zinc Materials. Chemsuschem 2009, 2, 99–112. [Google Scholar] [CrossRef] [PubMed]
  61. Bambagioni, V.; Bianchini, C.; Marchionni, A.; Filippi, J.; Vizza, F.; Teddy, J.; Serp, P.; Zhiani, M. Pd and Pt-Ru anode electrocatalysts supported on multi-walled carbon nanotubes and their use in passive and active direct alcohol fuel cells with an anion-exchange membrane (alcohol = methanol, ethanol, glycerol). J. Power Source. 2009, 190, 241–251. [Google Scholar] [CrossRef]
  62. Singh, R.N.; Singh, A. Anindita Electrocatalytic activity of binary and ternary composite films of Pd, MWCNT and Ni, Part II: Methanol electrooxidation in 1 M KOH. Int. J. Hydrogen Energ. 2009, 34, 2052–2057. [Google Scholar] [CrossRef]
  63. Xu, C.W.; Tian, Z.Q.; Shen, P.K.; Jiang, S.P. Oxide (CeO2, NiO,Co(3)O(4)and Mn3O4)-promoted Pd/C electrocatalysts for alcohol electrooxidation in alkaline media. Electrochim. Acta 2008, 53, 2610–2618. [Google Scholar] [CrossRef]
  64. Borkowska, Z.; Tymosiak-Zielinska, A.; Nowakowski, R. High catalytic activity of chemically activated gold electrodes towards electro-oxidation of methanol. Electrochim. Acta 2004, 49, 2613–2621. [Google Scholar] [CrossRef]
  65. Borkowska, Z.; Tymosiak-Zielinska, A.; Shul, G. Electrooxidation of methanol on polycrystalline and single crystal gold electrodes. Electrochim. Acta 2004, 49, 1209–1220. [Google Scholar] [CrossRef]
  66. Fleischmann, M.; Korinek, K.; Pletcher, D. The oxidation of organic compounds at a nickel anode in alkaline solution. J. Electroanal. Chem. 1971, 31, 39–49. [Google Scholar] [CrossRef]
  67. Van Effen, R.M.; Evans, D.H. A study of aldehyde oxidation at glassy carbon, mercury, copper, silver, gold and nickel anodes. J. Electroanal. Chem. 1979, 103, 383–397. [Google Scholar] [CrossRef]
  68. Motheo, A.J.; Machado, S.A.S.; Rabelo, F.J.B.; Santos, J.R., Jr. Electrochemical study of ethanol oxidation on nickel in alkaline media. J. Braz. Chem. Soc. 1994, 5, 161–165. [Google Scholar] [CrossRef]
  69. Abdel Rahim, M.A.; Abdel Hameed, R.M.; Khalil, M.W. Nickel as a catalyst for the electro-oxidation of methanol in alkaline medium. J. Power Source. 2004, 134, 160–169. [Google Scholar] [CrossRef]
  70. Khalil, M.W.; Abdel Rahim, M.A.; Zimmer, A.; Hassan, H.B.; Abdel Hameed, R.M. Nickel impregnated silicalite-1 as an electro-catalyst for methanol oxidation. J. Power Source. 2005, 144, 35–41. [Google Scholar] [CrossRef]
  71. Kumar, K.S.; Haridoss, P.; Seshadri, S.K. Synthesis and characterization of electrodeposited Ni-Pdalloy electrodes for methanol oxidation. Surf. Coat. Tech. 2008, 202, 1764–1770. [Google Scholar] [CrossRef]
  72. Shobba, T.; Mayanna, S.M.; Sequeira, C.A.C. Preparation and characterization of Co-W alloys as anode materials for methanol fuel cells. J. Power Source. 2002, 108, 261–264. [Google Scholar] [CrossRef]
  73. Qi, W.; Zhou, D.B.; Chen, S.L.; Huang, Y.; Cheng, X. Preparation and Electrocatalytic Properties of Fe, Co, Ni-Polymer-C Complex Catalysts for Ethanol Electro-oxidation. Acta Chim. Sin. 2009, 67, 917–922. [Google Scholar]
  74. Raghuveer, V.; Viswanathan, B. Can La2-xSrxCuO4 be used as anodes for direct methanol fuel cells? Fuel 2002, 81, 2191–2197. [Google Scholar] [CrossRef]
  75. Raghuveer, V.; Ravindranathan Thampi, K.; Xanthopoulos, N.; Mathieu, H.J.; Viswanathan, B. Rare earth cuprates as electrocatalysts for methanol oxidation. Solid State Ionics 2001, 140, 263–274. [Google Scholar] [CrossRef]
  76. Yu, H.-C.; Fung, K.-Z.; Guo, T.-C.; Chang, W.-L. Syntheses of perovskite oxides nanoparticles La1-xSrxMO3-[delta] (M = Co and Cu) as anode electrocatalyst for direct methanol fuel cell. Electrochim. Acta 2004, 50, 811–816. [Google Scholar] [CrossRef]
  77. Singh, R.N.; Sharma, T.; Singh, A.; Anindita; Mishra, D.; Tiwari, S.K. Perovskite-type La2-xSrxNiO4 (0 ≤ x ≤ 1) as active anode materials for methanol oxidation in alkaline solutions. Electrochim. Acta 2008, 53, 2322–2330. [Google Scholar] [CrossRef]
  78. Tsivadze, A.; Tarasevich, M.; Bogdanovskaya, V.; Ehrenburg, M. Platinum-free nanosize electrocatalysts for glycerol oxidation. Dokl. Chem. 2008, 419, 54–56. [Google Scholar] [CrossRef]
  79. Yu, E.Y.; Scott, K.; Reeve, R.W. Electrochemical Reduction of Oxygen on Carbon Supported Pt and Pt/Ru Fuel Cell Electrodes in Alkaline Solutions. Fuel Cells 2003, 3, 169–176. [Google Scholar] [CrossRef]
  80. Gamburzev, S.; Petrov, K.; Appleby, A.J. Silver-carbon electrocatalyst for air cathodes in alkaline fuel cells. J. Appl. Electrochem. 2002, 32, 805–809. [Google Scholar] [CrossRef]
  81. Kostowskyj, M.A.; Gilliam, R.J.; Kirk, D.W.; Thorpe, S.J. Silver nanowire catalysts for alkaline fuel cells. Int. J. Hydrogen Energ. 2008, 33, 5773–5778. [Google Scholar] [CrossRef]
  82. Meng, H.; Wu, M.; Hu, X.X.; Nie, M.; Wei, Z.D.; Shen, P.K. Selective cathode catalysts for mixed-reactant alkaline alcohol fuel cells. Fuel Cells 2006, 6, 447–450. [Google Scholar] [CrossRef]
  83. Jiang, L.; Hsu, A.; Chu, D.; Chen, R. Oxygen Reduction Reaction on Carbon Supported Pt and Pd in Alkaline Solutions. J. Electrochem. Soc. 2009, 156, B370–B376. [Google Scholar] [CrossRef]
  84. Kim, J.; Momma, T.; Osaka, T. Cell performance of Pd-Sn catalyst in passive direct methanol alkaline fuel cell using anion exchange membrane. J. Power Source. 2009, 189, 999–1002. [Google Scholar] [CrossRef]
  85. Klápste, B.; Vondrák, J.; Velická, J. MnOx/C composites as electrode materials II. Reduction of oxygen on bifunctional catalysts based on manganese oxides. Electrochim. Acta 2002, 47, 2365–2369. [Google Scholar] [CrossRef]
  86. Tachibana, K.; Matsuki, K. Development of in situ a.c. impedance measurement system under constant-current conditions and its application to galvanostatic discharge of electrolytic manganese dioxide in alkaline solution. J. Power Source. 1998, 74, 29–33. [Google Scholar] [CrossRef]
  87. Mao, L.; Sotomura, T.; Nakatsu, K.; Koshiba, N.; Zhang, D.; Ohsaka, T. Electrochemical Characterization of Catalytic Activities of Manganese Oxides to Oxygen Reduction in Alkaline Aqueous Solution. J. Electrochem. Soc. 2002, 149, A504–A507. [Google Scholar] [CrossRef]
  88. Elzing, A.; van der Putten, A.; Visscher, W.; Barendrecht, E. The cathodic reduction of oxygen at cobalt phthalocyanine: Influence of electrode preparation on electrocatalysis. J. Electroanal. Chem. 1986, 200, 313–322. [Google Scholar] [CrossRef]
  89. Kiros, Y.; Schwartz, S. Pyrolyzed macrocycles on high surface area carbons for the reduction of oxygen in alkaline fuel cells. J. Power Source. 1991, 36, 547–555. [Google Scholar] [CrossRef]
  90. Sarangapani, S.; Lessner, P.; Manoukian, M.; Giner, J. Non-noble electrocatalysts for alkaline fuel cells. J. Power Source. 1990, 29, 437–442. [Google Scholar] [CrossRef]
  91. Kiros, Y.; Lindström, O.; Kaimakis, T. Cobalt and cobalt-based macrocycle blacks as oxygen-reduction catalysts in alkaline fuel cells. J. Power Source. 1993, 45, 219–227. [Google Scholar] [CrossRef]
  92. Van Den Brink, F.; Visscher, W.; Barendrecht, E. Electrocatalysis of cathodic oxygen reduction by metal phthalocyanines: Part I. Introduction, cobalt phthalocyanine as electrocatalyst: experimental part. J. Electroanal. Chem. 1983, 157, 283–304. [Google Scholar] [CrossRef]
  93. Van Den Brink, F.; Visscher, W.; Barendrecht, E. Electrocatalysis of cathodic oxygen reduction by metal phthalocyanines: Part II. Cobalt phthalocyanine as electrocatalyst: A mechanism of oxygen reduction. J. Electroanal. Chem. 1983, 157, 305–318. [Google Scholar] [CrossRef]
  94. Van den Brink, F.; Visscher, W.; Barendrecht, E. Electrocatalysis of cathodic oxygen reduction by metal phthalocyanines: Part IV. Iron phthalocyanine as electrocatalyst: Mechanism. J. Electroanal. Chem. 1984, 175, 279–289. [Google Scholar] [CrossRef]
  95. Van Den Brink, F.; Visscher, W.; Barendrecht, E. Electrocatalysis of cathodic oxygen reduction by metal phthalocyanines: Part III. Iron phthalocyanine as electrocatalyst: Experimental part. J. Electroanal. Chem. 1984, 172, 301–325. [Google Scholar] [CrossRef]
  96. Bianchini, C.; Bambagioni, V.; Filippi, J.; Marchionni, A.; Vizza, F.; Bert, P.; Tampucci, A. Selective oxidation of ethanol to acetic acid in highly efficient polymer electrolyte membrane-direct ethanol fuel cells. Electrochem. Commun. 2009, 11, 1077–1080. [Google Scholar] [CrossRef]
  97. Zhiani, M.; Gasteiger, H.A.; Piana, M.; Catanorchi, S.; Bert, P. Comparative Study between Pt/C and Low Cost Cathode Nano-Particle Catalyst in Alkalien Direct Ethanol Fuel Cell (DEFC). In Proceedings of ECS 216th Meeting, Vienna, Austria, 4–9 October 2009.
  98. Li, Y.S.; Zhao, T.S.; Liang, Z.X. Performance of alkaline electrolyte-membrane-based direct ethanol fuel cells. J. Power Source. 2009, 187, 387–392. [Google Scholar] [CrossRef]
  99. Yu, E.H.; Scott, K. Development of direct methanol alkaline fuel cells using anion exchange membranes. J. Power Source. 2004, 137, 248–256. [Google Scholar] [CrossRef]
  100. Yu, E.H.; Scott, K. Direct methanol alkaline fuel cells with catalysed anion exchange membrane electrodes. J. Appl. Electrochem. 2005, 35, 91–96. [Google Scholar] [CrossRef]
  101. Yu, E.H.; Scott, K.; Reeve, R.W. Application of sodium conducting membranes in direct methanol alkaline fuel cells. J. Appl. Electrochem. 2006, 36, 25–32. [Google Scholar] [CrossRef]
  102. Baldauf, M.; Preidel, W. Status of the development of a direct methanol fuel cell. J. Power Source. 1999, 84, 161–166. [Google Scholar] [CrossRef]
  103. Scott, K.; Yu, E.; Vlachogiannopoulos, G.; Shivare, M.; Duteanu, N. Performance of a direct methanol alkaline membrane fuel cell. J. Power Source. 2008, 175, 452–457. [Google Scholar] [CrossRef]
  104. Yanagi, H.; Fukuta, K. Anion Exchange Membrane and Ionomer for Alkaline Membrane Fuel Cells (AMFCs). ECS Trans. 2008, 16, 257–262. [Google Scholar]
  105. Danks, T.N.; Slade, R.C.T.; Varcoe, J.R. Comparison of PVDF- and FEP-based radiation-grafted alkaline anion-exchange membranes for use in low temperature portable DMFCs. J. Mater. Chem. 2002, 12, 3371–3373. [Google Scholar] [CrossRef] [Green Version]
  106. Danks, T.N.; Slade, R.C.T.; Varcoe, J.R. Alkaline anion-exchange radiation-grafted membranes for possible electrochemical application in fuel cells. J. Mater. Chem. 2003, 13, 712–721. [Google Scholar] [CrossRef] [Green Version]
  107. Kang, J.J.; Li, W.Y.; Lin, Y.A.; Li, X.P.; Xiao, X.R.; Fang, S.B. Synthesis and ionic conductivity of a polysiloxane containing quaternary ammonium groups. Polym. Adv. Tech. 2004, 15, 61–64. [Google Scholar] [CrossRef]
  108. Yi, F.; Yang, X.P.; Li, Y.J.; Fang, S.B. Synthesis and ion conductivity of poly(oxyethylene) methacrylates containing a quaternary ammonium group. Polym. Adv. Tech. 1999, 10, 473–475. [Google Scholar] [CrossRef]
  109. Fang, J.; Shen, P.K. Quaternized poly(phthalazinon ether sulfone ketone) membrane for anion exchange membrane fuel cells. J. Membrane Sci. 2006, 285, 317–322. [Google Scholar] [CrossRef]
  110. Li, L.; Wang, Y.X. Sulfonated polyethersulfone Cardo membranes for direct methanol fuel cell. J. Membrane Sci. 2005, 246, 167–172. [Google Scholar] [CrossRef]
  111. Pan, Y.; Huang, Y.H.; Liao, B.; Cong, G.M. Synthesis and characterization of aminated poly(2,6-dimethyl-1,4-phenylene oxide). J. Appl. Polym. Sci. 1996, 61, 1111–1115. [Google Scholar] [CrossRef]
  112. Wu, L.; Xu, T.W.; Wu, D.; Zheng, X. Preparation and characterization of CPPO/BPPO blend membranes for potential application in alkaline direct methanol fuel cell. J. Membrane Sci. 2008, 310, 577–585. [Google Scholar] [CrossRef]
  113. Wu, L.; Xu, T.W. Improving anion exchange membranes for DMAFCs by inter-crosslinking CPPO/BPPO blends. J. Membrane Sci. 2008, 322, 286–292. [Google Scholar] [CrossRef]
  114. Li, L.; Wang, Y.X. Quaternized polyethersulfone Cardo anion exchange membranes for direct methanol alkaline fuel cells. J. Membrane Sci. 2005, 262, 1–4. [Google Scholar] [CrossRef]
  115. Xing, B.; Savadogo, O. Hydrogen/oxygen polymer electrolyte membrane fuel cells (PEMFCs) based on alkaline-doped polybenzimidazole (PBI). Electrochem. Commun. 2000, 2, 697–702. [Google Scholar] [CrossRef]
  116. Hou, H.Y.; Sun, G.Q.; He, R.H.; Sun, B.Y.; Jin, W.; Liu, H.; Xin, Q. Alkali doped polybenzimidazole membrane for alkaline direct methanol fuel cell. Int. J. Hydrogen Energ. 2008, 33, 7172–7176. [Google Scholar]
  117. Hou, H.Y.; Sun, G.Q.; He, R.H.; Wu, Z.M.; Sun, B.Y. Alkali doped polybenzimidazole membrane for high performance alkaline direct ethanol fuel cell. J. Power Source. 2008, 182, 95–99. [Google Scholar] [CrossRef]
  118. Modestov, A.D.; Tarasevich, M.R.; Leykin, A.Y.; Filimonov, V.Y. MEA for alkaline direct ethanol fuel cell with alkali doped PBI membrane and non-platinum electrodes. J. Power Source. 2009, 188, 502–506. [Google Scholar] [CrossRef]
  119. Yang, C.C.; Chiu, S.J.; Chien, W.C. Development of alkaline direct methanol fuel cells based on crosslinked PVA polymer membranes. J. Power Source. 2006, 162, 21–29. [Google Scholar] [CrossRef]
  120. Yang, C.C.; Lee, Y.J.; Chiu, S.J.; Lee, K.T.; Chien, W.C.; Lin, C.T.; Huang, C.A. Preparation of a PVA/HAP composite polymer membrane for a direct ethanol fuel cell (DEFC). J. Appl. Electrochem. 2008, 38, 1329–1337. [Google Scholar] [CrossRef]
  121. Yang, C.C.; Chiu, S.J.; Lee, K.T.; Chien, W.C.; Lin, C.T.; Huang, C.A. Study of poly(vinyl alcohol)/titanium oxide composite polymer membranes and their application on alkaline direct alcohol fuel cell. J. Power Source. 2008, 184, 44–51. [Google Scholar] [CrossRef]
  122. Yang, C.-C.; Chiu, S.-J.; Chien, W.-C.; Chiu, S.-S. Quaternized poly(vinyl alcohol)/alumina composite polymer membranes for alkaline direct methanol fuel cells. J. Power Source. 2010, 195, 2212–2219. [Google Scholar] [CrossRef]
  123. Cairns, E.J.; Bartosik, D.C. A Methanol Fuel Cell with an Invariant Alkaline Electrolyte. J. Electrochem. Soc. 1964, 111, 1205–1210. [Google Scholar] [CrossRef]
  124. Varcoe, J.R.; Slade, R.C.T. An electron-beam-grafted ETFE alkaline anion-exchange membrane in metal-cation-free solid-state alkaline fuel cells. Electrochem. Commun. 2006, 8, 839–843. [Google Scholar] [CrossRef] [Green Version]
  125. Coutanceau, C.; Demarconnay, L.; Lamy, C.; Léger, J.M. Development of electrocatalysts for solid alkaline fuel cell (SAFC). J. Power Source. 2006, 156, 14–19. [Google Scholar] [CrossRef]
  126. Kim, J.-H.; Kim, H.-K.; Hwang, K.-T.; Lee, J.-Y. Performance of air-breathing direct methanol fuel cell with anion-exchange membrane. Int. J. Hydrogen Energ. 2010, 35, 768–773. [Google Scholar] [CrossRef]
  127. Bidault, F.; Kucernak, A. A novel cathode for alkaline fuel cells based on a porous silver membrane. J. Power Source. 2010, 195, 2549–2556. [Google Scholar] [CrossRef]
  128. Choban, E.R.; Spendelow, J.S.; Gancs, L.; Wieckowski, A.; Kenis, P.J.A. Membraneless laminar flow-based micro fuel cells operating in alkaline, acidic, and acidic/alkaline media. Electrochim. Acta 2005, 50, 5390–5398. [Google Scholar] [CrossRef]
  129. Priestnall, M.A.; Kotzeva, V.P.; Fish, D.J.; Nilsson, E.M. Compact mixed-reactant fuel cells. J. Power Source. 2002, 106, 21–30. [Google Scholar] [CrossRef]
  130. Sung, W.; Choi, J.-W. A membraneless microscale fuel cell using non-noble catalysts in alkaline solution. J. Power Source. 2007, 172, 198–208. [Google Scholar] [CrossRef]
  131. Zeng, R.; Shen, P.K. Selective membrane electrode assemblies for bipolar plate-free mixed-reactant fuel cells. J. Power Source. 2007, 170, 286–290. [Google Scholar] [CrossRef]
  132. Shen, P.K.; Xu, C. Alcohol oxidation on nanocrystalline oxide Pd/C promoted electrocatalysts. Electrochem. Commun. 2006, 8, 184–188. [Google Scholar] [CrossRef]
  133. Faubert, G.; Cote, R.; Guay, D.; Dodelet, J.P.; Denes, G.; Bertrand, P. Iron catalysts prepared by high-temperature pyrolysis of tetraphenylporphyrins adsorbed on carbon black for oxygen reduction in polymer electrolyte fuel cells. Electrochim. Acta 1998, 43, 341–353. [Google Scholar] [CrossRef]
  134. Scott, K.; Shukla, A.K.; Jackson, C.L.; Meuleman, W.R.A. A mixed-reactants solid-polymer-electrolyte direct methanol fuel cell. J. Power Source. 2004, 126, 67–75. [Google Scholar] [CrossRef]
  135. Ganley, C.J.; Karikari, N.K.; Raghavan, D. Performance enhancement of alkaline direct methanol fuel cells by Ni/Al layered double hydroxides. J. Fuel Cell Sci. Tech. 2010, 7, 031019. [Google Scholar] [CrossRef]

Share and Cite

MDPI and ACS Style

Yu, E.H.; Krewer, U.; Scott, K. Principles and Materials Aspects of Direct Alkaline Alcohol Fuel Cells. Energies 2010, 3, 1499-1528. https://doi.org/10.3390/en3081499

AMA Style

Yu EH, Krewer U, Scott K. Principles and Materials Aspects of Direct Alkaline Alcohol Fuel Cells. Energies. 2010; 3(8):1499-1528. https://doi.org/10.3390/en3081499

Chicago/Turabian Style

Yu, Eileen Hao, Ulrike Krewer, and Keith Scott. 2010. "Principles and Materials Aspects of Direct Alkaline Alcohol Fuel Cells" Energies 3, no. 8: 1499-1528. https://doi.org/10.3390/en3081499

Article Metrics

Back to TopTop