Version Changes
Revised. Amendments from Version 2
The third version has benefited from the significant comments provided by Ying Li. The major revisions in this new version include updated abstract and conclusion as suggested. In addition, two subsection headings have been changed to better maintain the consistency of the manuscript.
Abstract
Aqueous-phase reforming (APR) can convert methanol and other oxygenated hydrocarbons to hydrogen and carbon dioxide at lower temperatures when compared with the corresponding gas phase process. APR favours the water-gas shift (WGS) reaction and inhibits alkane formation; moreover, it is a simpler and more energy efficient process compared to gas-phase steam reforming. For example, Pt-based catalysts supported on alumina are typically selected for methanol APR, due to their high activity at temperatures of circa 200°C. However, non-noble catalysts such as nickel (Ni) supported on metal-oxides or zeolites are being investigated with promising results in terms of catalytic activity and stability. The development of APR kinetic models and reactor designs is also being addressed to make APR a more attractive process for producing in situ hydrogen. This can also lead to the possibility of APR integration with high-temperature proton exchange membrane fuel cells. The integration can result into increased overall system efficiency and avoiding critical issues faced in the state-of-the-art fuel cells integrated with methanol steam reforming.
Keywords: Aqueous-phase reforming, methanol, hydrogen, catalyst, fuel cell, PEMFC
Abbreviations
APR: aqueous-phase reforming
BET: Brunauer-Emmett-Teller
CHP: combined heat and power
FC: fuel cell
HT-PEMFC: high temperature polymer electrolyte membrane fuel cell
HHV: higher heating value
IWI: incipient wetness impregnation
LT-PEMFC: low temperature polymer electrolyte membrane fuel cell
MeOH: methanol
MEA: membrane electrode assembly
MSR: methanol steam reforming
PR: production rate
WGS: water-gas shift
WHSV: weight hourly space velocity
WI: wetness impregnation
Nomenclature
Δ H° = enthalpy change
ε FP = thermal efficiency of fuel processor
K = equilibrium constant
k = rate constant
C = concentration
R = reaction rate
Introduction
Energy decarbonization is critically needed to mitigate climate change and to contribute to economic growth and technological progress. The US Energy Information Administration predicts that between 2018 and 2050, world energy consumption will grow 50% in total and more than 30% in the industrial sector 1 . According to the United Nations Sustainable Development Goals (goal 7), it is crucial that energy demand is met for the universal access of energy services since 13% of the world’s population still (2019) suffers from a lack of modern energy access 2 . Fossil fuels such as oil and coal continue to dominate the global energy sector. Consequentially, greenhouse gas emissions, which are pointed to as the main cause of extreme climate change around the world, continues to climb 3 ; the UN recorded 1.3 million climate-related deaths between 1997–2017. This is one of the many reasons to make the transition towards more environmentally friendly and decarbonized energy. Therefore, new sustainable and renewable energy sources are crucial to meet the energy demands created by an increased world population, technological advancements and to mitigate global warming 4, 5 .
Hydrogen has emerged as a potential energy vector for a clean energy transition 6 because it can be produced from a wide range of environmentally friendly sources and can be used for storing energy that is easily convertible into electricity and heat for long periods of time 4 . Hydrogen can be produced from renewable feedstock such as biomass and synthetic fuels as well as from fossil fuels 4, 7 . Different technological pathways can be used to produce hydrogen, both for stationary and/or mobile applications. Hydrogen has a high weight energy density of 120 MJ/kg (lower heating value) compared to a traditional fuel such as gasoline (44 MJ/kg), although it has a very low volumetric energy density 7, 8 . Hydrogen can be used in combustion engines and turbines to produce mechanical energy and in electrochemical processes to produce electricity 4, 9 . These qualities make hydrogen a promising energy vector. Examples of these are the hydrogen driven trains in Germany, UK and France, and ferries in Norway 10 . The European Commission has recently ( March 10, 2020) announced the ‘clean hydrogen alliance’ as an initiative to develop hydrogen-based technologies.
Fuel cells are considered to be the next generation of propulsion systems in automotive, portable and stationary applications 11 . In particular, hydrogen powered proton exchange membrane fuel cells (PEMFCs) are a promising clean power source due to low-emissions, rapid start times and high energy density and efficiency 12 . Because of the very low volumetric energy density, hydrogen storage and transportation remain a challenge to overcome and a network of hydrogen delivering infrastructure has to be implemented. An inexpensive strategy that fulfils all the requirements for hydrogen storage and transportation has yet to emerge 4, 7 . Therefore, in situ hydrogen production is an exciting approach to overcome the challenges of hydrogen storage and transportation. Currently, the reforming of fossil fuels, such as methane, is the most widely used process to produce hydrogen, representing 71% of total hydrogen produced globally 5, 13 . Since this source of hydrogen is not renewable, research has shifted toward hydrogen production from renewable sources such as ammonia and alcohols (glycol, methanol and ethanol) 14 . A large number of studies can be found on catalytic steam reforming of alcohols because of its high hydrogen selectivity and yield 14– 16 . In addition to steam reforming, aqueous-phase reforming (APR) has emerged recently as a promising competitive approach. Contrary to steam reforming, which employs high temperatures and mild pressures, APR converts aqueous phase organic compounds into hydrogen at lower temperatures. In fact, APR operates at temperatures between 200°C and 250°C; and pressures close to bubble point feedstock, normally from 1.5 MPa to 5.0 MPa 17 . These operating conditions make APR a potential technology for providing in situ hydrogen for automotive and mobile applications, allowing this fuel to be used as a direct hydrogen carrier. This review addresses the early state-of-the-art of APR process for in situ hydrogen production, targeting the integration with high temperature (HT)-PEMFCs.
Methanol APR
In 2002, Dumesic’s research group 18 reported APR of an oxygenated hydrocarbon for the first time. Using group VIII metal-based catalysts, this research group successfully demonstrated the APR of sugar and alcohol (e.g. ethylene glycol and methanol) at ca. 226°C in a single reactor. Additionally, these authors proposed a mechanism for hydrogen formation, which became a steppingstone in the APR field 18 .
APR offers a number of advantages compared with conventional steam or autothermal reforming; APR is more energy efficient, since it eliminates the vaporization of the aqueous/oxygenated hydrocarbon solutions 19 . Additionally, operating at lower temperatures improves the stability of catalysts, avoiding coke formation, sintering and oxidation 20 . The low operating temperatures favor the water-gas shift (WGS) reaction, which minimizes the formation of carbon monoxide; and finally, it mitigates undesired decomposition reactions related to carbohydrates 19 . Contrary to the multi-stage and multi-reactor steam reforming process, APR is an easy single-step and single reactor process 21, 22 .
Of all alcohols, methanol is the most convenient hydrogen carrier. With only one carbon, methanol is the simplest alcohol displaying the highest hydrogen to carbon ratio ( Figure 1), meaning that more molecular hydrogen can be extracted from methanol compared to other feedstock 15, 23 . Methanol displays 12.6 wt. % of hydrogen content, it is liquid at room temperature, it is biodegradable and can be easily produced from biomass and or synthesized 15 . Methanol reforms – via either steam reforming or APR – at the lowest temperature compared to ethanol or methane, because it has no C-C bonds 16, 24 .
Figure 1. Hydrogen/carbon (H/C) atomic ratio and hydrogen energy densities (10 -2 KJ mol -1 number of H atom -1) of various alcohols and alkanes (reproduced with permission from 25.
Copyright 2009. Wiley-VCH Verlag GmbH & Co. KGaA, Weinheim. John Wiley and Sons).
Reaction mechanism
The methanol APR mechanism involves cleaving C-H and O-H bonds to produce hydrogen (H 2) and CO ( Equation 1). The produced CO then reacts with water to form H 2 and CO 2 via the WGS reaction ( Equation 2) 17, 18 .
APR of oxygenated hydrocarbons ( Equation 1– Equation 2) occurs at low temperatures between 200°C and 250°C and are exothermic reactions. However, in both cases ( Equation 1– Equation 2), the reactions thermodynamically favored by higher temperatures. On the other hand, WGS reaction equilibrium ( Equation 2), which converts CO into CO 2 and H 2, is favored by lower temperatures and is an endothermic reaction 17, 18 . The reaction products of APR are broader than just CO 2 and H 2. APR conditions thermodynamically favor undesirable side reactions such as methanation and Fischer-Tropsch. The APR operating conditions make carbon oxides thermodynamically unstable, where the cleavage of the C-O bond followed by hydrogenation can form alkanes and water. For methanol APR, methane and water is produced via methanation of CO and Fischer-Tropsch reaction of CO 2 ( Equation 3– Equation 4) 17– 19 . Figure 2 shows a schematic representation of methanol APR reaction pathways.
Figure 2. Reaction mechanisms of methanol aqueous-phase reforming (APR).
Catalysts
The focus has been placed on finding suitable catalysts that show high activity for promoting reforming reactions at low temperatures ( Equation 1– Equation 2) and inhibit side reactions ( Equation 3– Equation 4) 17, 19 . Various catalysts have been used for methanol APR based on noble (Pt, Ru) and non-noble (Ni, Cu) metals, yet the most commonly used catalyst for methanol APR is Pt-based. Dumesic’s group 18 made significant advances in APR by using Pt supported on aluminum oxide catalysts. Many authors opted for Pt as it is very active and displays high selectivity towards H 2 18, 26, 27 . Despite its high catalytic activity for methanol APR, Pt shows lower C-C cleavage activity compared to other metals (Ru, Ni and Ir) 28 , needed for the APR of heavier alcohols, and reasonable activity for the WGS reaction as shown in Figure 3 29 . Additionally, metals such as Cu show the highest activity for WGS but no activity for C-C bond breaking. Different supports, with large surface areas, have also been studied with the objective of increasing the Pt loading in catalysts to increase catalytic activity of Pt-based catalysts 22, 26, 30
Figure 3. Relative rates for C-C bond cleavage (black) reported in 28, water-gas shift reaction (green) reported in 29 and methanation reaction (red) reported in 35 of various metals.
Nickel-based catalysts have also been investigated for the APR reaction. Although Ni-based catalysts show lower stability and H 2 selectivity, they favor the WGS shift reaction and are inexpensive compared to Pt. Davda et al. 17 determined the catalytic activities of various silica supported noble metals (based on the rate of CO 2 production) for ethylene glycol APR and found that they decrease in the following order: Pt ~ Ni > Ru > Rh ~ Pd > Ir. To improve stability and H 2 selectivity of Ni-based catalysts, various support materials have been considered 15, 20, 31– 33 . However, reports on methanol APR via Ni-based catalysts are still limited. Shabaker et al. 34 and Coronado et al. 31, 32 made significant advances in this field. The least studied catalysts are of Cu, Ce or Ru. These metal-based catalysts have been used as bi-metallic combinations or their complexes, however. For example, a Ru complex for methanol APR showed high H 2 yield with turnover frequency of 4700 h -1 and low amounts of unwanted side products (CO and CH 4) 36 .
Platinum-based catalysts. As previously mentioned, Pt is the favoured metal for APR reactions, showing high activity and selectivity for reforming of oxygenated hydrocarbons 18 , although low C-C cleavage activity 17 . To enhance the stability and activity of the Pt-based catalyst, different supports have been investigated such as ZnO, ZrO 2, TiO 2 and carbon black 17 . However, studies on Pt-based catalysts for methanol APR are limited and mostly focused on Pt supported on ɣ-Al 2O 3. Shabaker et al . 19 performed a detailed study of methanol APR reaction kinetics using a Pt catalyst supported on alumina (Pt/ɣ-Al 2O 3), reporting 100% selectivity for H 2 concerning alkane formation. ɣ-Al 2O 3 showed very high H 2 yield compared to other supports, followed by α-Al 2O 3 and δ-Al 2O 3 37 . In contrast, α-Fe 2O 3 and CeO 2 showed low methanol conversion and low H 2 yield 37 . Furthermore, Pt/ɣ-Al 2O 3 showed improved WGS activity since it had the highest rate of CO 2 production amongst various Pt-based catalysts. However, under APR reaction conditions, ɣ-Al 2O 3 is hydrated to boehmite (AlO(OH)). This reaction reduces the support surface area and promotes the sintering and encapsulation of the Pt nanoparticles. Consequently, it decreases the metal dispersion over the surface of the support, which affects APR performance due to the reduction of the catalyst active sites 38, 39 . Additionally, it also increases the surface acidity of the catalyst, leading to the formation of undesired side products 39 . Therefore, it is crucial to find suitable supports for Pt-based catalysts that increase catalytic activity and stability. Following this, Li et al . 40 prepared Pt supported on nickel aluminate (NiAl 2O 4) spinel. They observed that the H 2 yield from Pt/NiAl 2O 4 catalyst was four times higher than that of a Pt/Al 2O 3 catalyst. The authors attributed the noticeably higher dehydrogenation rate of methanol to the NiAl 2O 4 spinel structure and synergy between Pt and NiAl 2O 4. The large amount of oxygen vacancies over Pt/NiAl 2O 4 stabilize Pt and significantly improved WGS activity. Interestingly, they report negligible activity of NiAl 2O 4 in APR tests, which indicates clearly a synergistic effect between Pt and NiAl 2O 4. The performance and stability of this catalyst for APR is also affected by the oxidation state of Pt on NiAl 2O 4, which should be reduced to metallic Pt from PtO x. Zeolites have also been tested as a potential support for methanol APR catalysts 26 . Pt supported on NaY zeolite shows a slight increase in methanol conversion and H 2 selectivity compared with a Pt/Al 2O 3 catalyst 26 . This fact was attributed to the synergistic effect between Pt and NaY, where the high electron density of the microporous zeolite surface stabilizes Pt and enhances the hydrogen retention capacity in Pt/NaY catalysts 26 . The highest H 2 production rate reported was achieved using Pt supported on various structures of molybdenum carbide (α-MoC, β-MoC), as can be seen in Table 1 27 . Lin et al . 27 using density functional theory (DFT) calculations showed that the enhanced APR activity is due to high interaction between Pt and MoC caused by highly atomically dispersed Pt atoms on the MoC surface. Increasing Pt loading usually leads to higher H 2 yield and production rate. They observed a sharp increase in H 2 production rate when Pt loading increased from 2 wt. % to 2.2 wt. % 27 . Higher Pt load leads to the increased number of active sites for the reaction. Figure 4 shows the methanol conversion as a function of the Pt load.
Figure 4. Effect of Pt loading in methanol aqueous-phase reforming.
Reprinted with permission from 26. Copyright 2009. American Chemical Society.
Table 1. Pt-based catalyst for hydrogen production in methanol aqueous-phase reforming (IWI – incipient wetness impregnation, WI – wetness impregnation, WHSV – weight hourly space velocity).
Support | Pt
loading (wt. %) |
Preparation
method |
Temperature
(°C) |
Pressure
(MPa) |
Methanol
concentration (wt. %) |
WHSV
(h -1) |
Methanol
conversion (%) |
H
2 production
rate (µmol/min/ g catalyst) |
Ref. |
---|---|---|---|---|---|---|---|---|---|
ɣ- Al 2O 3 | 3 | IWI | 225 | 2.5 | 1 | 0.008 | 94 | 11.6 | 18 |
ɣ- Al 2O 3 | 3 | IWI | 267 | 2.5 | 1 | 0.008 | 94 | 11.6 | 18 |
Al 2O 3 | 2 | WI | 190 | 6 | 294 | 27 | |||
ɣ- Al 2O 3 | IWI | 13.1 | 197 | 37 | |||||
ɣ-Al 2O 3 | 8 | IWI | 230 | 3.2 | 5 | 2 | 41.2 | 32 | |
ɣ- Al 2O 3 | 3 | IWI | 224 | 2.65 | 1 | 0.008 | 78.9 | 9.5 | 26 |
ɣ- Al 2O 3 | 3 | IWI | 264 | 2.65 | 1 | 0.008 | 95.5 | 11.8 | 26 |
ɣ-Al 2O 3 | 1 | IWI | 210 | 2.9 | 10 | 2.94 | 26.5 | 107.3 | 40 |
δ- Al 2O 3 | IWI | 13.8 | 164 | 37 | |||||
α- Al 2O 3 | IWI | 10.5 | 145 | 37 | |||||
CeO 2 | IWI | 3.8 | 41 | 37 | |||||
α-Fe 2O 3 | IWI | 4.5 | 24 | 37 | |||||
α-MoC | 2.2 | WI | 190 | 6 | 7776 | 27 | |||
α-MoC | 2 | WI | 170 | 6 | 3000 | 27 | |||
α-MoC | 2 | WI | 150 | 6 | 924 | 27 | |||
α-MoC | 0.2 | WI | 190 | 6 | 4572 | 27 | |||
β-MoC | 2.1 | WI | 190 | 6 | 300 | 27 | |||
MoC-2 | 2 | WI | 190 | 6 | 4092 | 27 | |||
MoC-3 | 2.3 | WI | 190 | 6 | 636 | 27 | |||
NaY | 0.5 | 224 | 2.65 | 1 | 0.08 | 81.0 | 34.8 | 26 | |
NaY | 0.5 | 264 | 2.65 | 1 | 0.08 | 98.8 | 44.4 | 26 | |
NiAl 2O 4 | 1 | IWI | 210 | 2.9 | 10 | 2.94 | 99.9 | 439.2 | 40 |
TiO 2 | 2 | WI | 190 | 6 | 252 | 27 |
Ni-based catalysts. Ni-based catalysts have been investigated for the APR process, since they show appreciable H 2 production and are less expensive than Pt-based catalysts. Table 2 summarizes the Ni-based catalysts reported in the literature for methanol APR. Ni-based catalysts are more susceptible to oxidation in hydrothermal APR conditions 41 . Additionally, sintering, agglomeration and coke formation leads to the deactivation of Ni-based catalysts 39, 42 . To overcome these issues, Ni has been supported on various materials to increase H 2 production along with stability. Raney nickel offers a notably high surface area, thermal and structural stability; adding Sn does not influence the reaction rate of Raney nickel catalysts, but increases the selectivity for H 2 by weakening CO and H adsorption, which lowers the methanation reaction rate 34 . The rate of H 2 production, selectivity and stability of Raney NiSn is comparable to Pt/Al 2O 3 catalysts. Ni has been supported on metal oxides, such as Al 2O 3 and SiO 2, but also on CeO 2 and ZrO 2 ( Table 2). It should be emphasised that both CeO 2 and ZrO 2 have catalytic activity themselves, making them attractive for APR. For instance, ZrO 2 enhances WGS activity 43 . The use of mixed oxides (bimetallic supports) such as CeO 2-ZrO 2 and La 2O 3-ZrO 2 improved methanol conversion, hydrogen production and catalyst stability 44 . The addition of CeO 2 increases oxygen storage capacity, catalyst mobility and promotes redox properties. Ni supported on [CeO 2] 0.25-[ZrO 2] 0.75, mass fractions showed superior performance compared to Ni supported on [CeO 2] 0.17-[ZrO 2] 0.83 or compared to a Ni/ZrO 2 catalyst 44 . However, compared to the performance of Ni supported on CeO 2-ZrO 2, Ni/CeO 2 based catalysts showed worse results; poor metal dispersion due to a lesser surface area was pointed to as the main reason 44, 45 . Ni has also been tested as a catalyst without using any support material. Shabaker et al. 34 used a Sn modified Ni catalyst for reforming various hydrocarbons and concluded that adding Sn was required to supress the alkane formation reaction. Nevertheless, Ni without any support showed better H 2 selectivity since it avoids the often-observed dehydrogenation of the catalyst support.
Table 2. Ni-based catalysts for hydrogen production in methanol aqueous-phase reforming (IWI – incipient wetness impregnation, WI – wetness impregnation, WHSV – weight hourly space velocity).
Support | Ni
loading (wt. %) |
Preparation
method |
Temperature
(°C) |
Pressure
(MPa) |
Methanol
conversion (%) |
WHSV
(h -1) |
H
2 yield
(%) |
H
2 production
rate (µmol/min/ g catalyst) |
Ref. |
---|---|---|---|---|---|---|---|---|---|
ZrO 2 | 10 | IWI | 230 | 3.2 | 2 | 2 | 20.3 | 32 | |
ZrO 2 | 7 | IWI | 230 | 3.2 | 48 | 4 | 40 | 45 | |
ZrO 2 | 10 | IWI | 230 | 3.2 | 26.4 | 3.6 | 1021 | 44 | |
[CeO 2] 0.17-[ZrO 2] 0.83 | 10 | IWI | 230 | 3.2 | 50.9 | 3.6 | 2144 | 44 | |
[CeO 2] 0.25-[ZrO 2] 0.75 | 10 | IWI | 230 | 3.2 | 58.8 | 3.6 | 2529 | 44 | |
[CeO
2]
0.17-[La
2O
3]
0.05-
[ZrO 2] 0.78 |
10 | IWI | 230 | 3.2 | 45.7 | 3.6 | 2134 | 44 | |
[La 2O 3] 0.10-[ZrO 2] 0.90 | 10 | IWI | 230 | 3.2 | 46.6 | 3.6 | 2147 | 44 | |
Calcia-stabilized
ZrO 2 |
5 | IWI | 230 | 3.2 | 75 | 4 | 64 | 45 | |
Yttria-stabilized ZrO 2 | 5.6 | IWI | 230 | 3.2 | 46 | 4 | 36 | 45 | |
Calcia-stabilized
ZrO 2 |
6.9 | IWI | 230 | 3.2 | 63 | 4 | 46 | 45 | |
CeO2 | 10 | IWI | 230 | 3.2 | 16.9 | 3.6 | 453 | 44 | |
α-Al 2O 3 | 11.8 | IWI | 230 | 3.2 | 2 | 0.8 | 8.6 | 32 | |
SiO 2 | 10 | IWI | 230 | 3.2 | 2 | 1 | 10.1 | 32 | |
ɣ-Al 2O 3 | 11.4 | IWI | 230 | 3.2 | 2 | 6 | 60.6 | 32 | |
R-NiSn
14 |
225 | 2.5 | 18.6 | 1.43 | 34 |
To increase basicity of the APR reaction and consequently the selectivity for H 2, calcia-stabilized ZrO 2 (CSZ) and yttria-stabilized ZrO 2 (YSZ) have been tested as supports 45 . The results showed that adding calcia to ZrO 2 improves the APR performance with higher methanol conversion and H 2 yield, while yttria-stabilized ZrO 2 had no impact on performance. However, increasing the loading of calcia for CSZ did not necessarily increase H 2 production. Lanthanum (La) has been employed with Ce and Zr to increase WGS activity and catalyst stability. Adding La 2O 3 to a CeO 2 support increases oxygen vacancies, provides better metal-support interactions and leads to higher H 2 production due to better Ni dispersion 44 .
Effect of dopants. Adding Cu to Ni-based catalysts supported on Al 2O 3 and SiO 2 significantly increases WGS activity and makes methane formation less favorable, as depicted in Table 3 32 . This behavior is less evident in other types of support. Coronado et al. 31 , using CeO 2-ZrO 2 supports, reported higher H 2 selectivity for alcohol dehydrogenation for Ni-Cu/[CeO 2] 0.25-[ZrO 2] 0.75, in comparison to Ni/[CeO 2] 0.25-[ZrO 2] 0.75, despite the lower methanol conversion and H 2 yield. The authors justified the lower performance with the lower loading of nickel in Ni-Cu/[CeO 2] 0.25-[ZrO 2] 0.75. On the other hand, adding Ce to Ni catalysts supported on Al 2O 3 and SiO 2 showed slight improvements in methanol conversion 32 . The opposite effect was observed for CeO 2-ZrO 2 supports 31 , due to an increase in the nickel particles’ size. Similar results were obtained by Goma et al. 45 , using a Ni-based catalyst doped with CeO 2 over different supports (ZrO 2, calcia-stabilized ZrO 2 and yttria-stabilized ZrO 2). The results showed that the CeO 2 doped catalysts have lower performances than those without CeO 2. The reduction properties and basic nature of Ce reduces the CO selectivity. Moreover, the CeO 2 layer additionally hindered the basic sites of calcia and yttria, lowering the methanol conversation and H 2 production.
Table 3. Ce and Cu doped Ni-based catalysts for hydrogen production in methanol aqueous-phase reforming at 230°C and 3.2 MPa (5 wt. % methanol concentration) (IWI – incipient wetness impregnation, WI – wetness impregnation, WHSV – weight hourly space velocity).
Metal | Support | Metal
loading (wt. %) |
Preparation
method |
Methanol
conversion (%) |
WHSV
(h -1) |
H
2 yield
(%) |
H
2 prodution
rate (µmol/min/ g catalyst) |
Ref. |
---|---|---|---|---|---|---|---|---|
Ni-Cu | [CeO 2] 0.25-[ZrO 2] 0.75 | Ni 10
Cu 5 |
IWI | 38 | 3.6 | 101 | 2100 | 31 |
Ni-Cu | SiO 2 | Ni 10
Cu 5 |
IWI | 2 | 1.0 | 10.9 | 32 | |
Ni-Cu | Al 2O 3 | Ni 13
Cu 1.3 |
IWI | 2 | 6.2 | 63.5 | 32 | |
Ni-Ce | [CeO 2] 0.25-[ZrO 2] 0.75 | Ni 10
Ce 5 |
IWI | 12 | 3.6 | 18 | 620 | 31 |
Ni-Ce | Al 2O 3 | Ni 13
Ce 1.3 |
IWI | 19 | 3.6 | 28 | 950 | 31 |
Ni-Ce | SiO 2 | Ni 10
Cu 5 |
IWI | 2 | 3.4 | 34.19 | 32 | |
Ni-Ce | Al 2O 3 | Ni 13
Cu 1.3 |
IWI | 2 | 8.3 | 84.7 | 32 | |
Ni-Ce | ZrO 2 | Ni 5.9
Ce 12.6 |
IWI | 40 | 4 | 34 | 45 | |
Ni-Ce | Calcia-stabilized ZrO 2 | Ni 5.9
Ce 13 |
IWI | 68 | 4 | 57 | 45 | |
Ni-Ce | Calcia-stabilized ZrO 2 | Ni 5.5
Ce 13.5 |
IWI | 44 | 4 | 33 | 45 | |
Ni-Ce | Yttria-stabilized ZrO 2 | Ni 5.5
Ce 12.7 |
IWI | 54 | 4 | 40 | 45 |
Noble metal-based catalysts. Besides Pt-based catalysts, other noble metals such as Ru, Ir, Rh and Pd have been employed in APR of alcohols 17, 20 , but only Ru and Ir were studied for methanol APR. The dehydrogenation of methanol at low temperatures between 89°C and 95°C, using Ru-based pincer complexes in alkaline solution, showed that decreasing the base concentration (NaOH) negatively affects the catalytic performance 36 . The presence of formate as an intermediate to produce H 2 and CO 2 was identified by Raman spectroscopy. The use of the Ru complex allows high H 2 yield with a turnover frequency of 4700 h -1 and low amounts of CO and CH 4 with long-term stability 36, 46 . Ir-based catalysts supported on SiO 2, TiO 2 and ZrO 2 have also been used in methanol APR. Adding Re to Ir/SiO 2 and Ir/ZrO 2 increased the H 2 formation rate by four times and five times, respectively 47 , by forming nanocomposite clusters that decreased the H 2 adsorption sites on Ir. Meanwhile, adding Mo to Ir/TiO 2 created a thin monolayer of MoO 2 on TiO 2 that is responsible for improving the water reactivity. This hydration process over the monolayer leads to an increased H 2 production rate.
Catalyst synthesis methods. The performance of the catalyst relies on the active surface area of the support, the dispersion of the metal over the support surface and the size of the active metal nanoparticles. Therefore, it is necessary to optimize the preparation method in order to achieve catalysts with highly dispersed metal over the support surface. The most commonly used methods for catalyst preparation are derivations from co-precipitation and wet-impregnation. It is important to opt for an appropriate aforementioned preparation method (depending on metal precursors and supports) in order to achieve optimal catalytic activity. Incipient wetness impregnation (IWI) is a commonly used method for preparation of catalysts for methanol APR. However, the use of metal nitrates often offers poor dispersion of the catalyst. Studies on Fe and Ni precursors attributed the reason for low dispersion to redistribution during the drying process. Adding viscosity-increasing agents or chelating citrate salts to aqueous nitrate solutions can avoid redistribution problems 48, 49 Figure 5 shows one example of the effect of surface area in methanol APR. As seen in this figure, the use of an alumina support with a larger surface area (α-Al 2O 3) corresponds to higher H 2 yield. The reason behind this could be that the larger surface area of the support allows for a better dispersion of Ni than for the support with lower surface area (ɣ-Al 2O 3).
Figure 5. Catalyst Brunauer–Emmett–Teller (BET) surface area and hydrogen yield in methanol aqueous-phase reforming, 230°C, 3.2 MPa, WHSV 2 h -1, Ni loading 11.8 wt. % for Ni/α-Al 2O 3 catalyst, Ni loading 13 wt. % for Ni/ɣ-Al 2O 3 (figure created using data obtained from 32).
Effect of active metal on alkane formation. In APR, methanation and Fischer-Tropsch reactions are responsible for alkane formation. For APR of methanol, only CH 4 is formed as side product alkane. The formation of undesired CH 4 can be suppressed to some extent by using suitable metal and/or support materials. Alkane formation heavily depends on the acidity of the catalyst material. Additionally, using catalysts of basic nature not only suppresses alkane formation, it is also favorable for WGS activity 50 . Huber et al. 51 reported that alkane selectivity increased as acidic sites were added to non-acidic Pt/Al catalyst.
As previously mentioned, different metals have their own methanation activity rate (Ni shows a higher tendency to form alkanes than Pt). In order to decrease alkane selectivity and favor higher H 2 formation, non-acidic dopants or supports must be used with a Ni-based catalyst during dehydrogenation reactions. Not only the metal of the catalyst, but also the support plays an equally vital part in APR reactions. Acidic supports like SiO 2-Al 2O 3 tend to favor higher alkane formation compared to basic or neutral supports such as Al 2O 3 17 . Ni-based catalysts have been doped with Cu in order to increase H 2 production and to suppress CH 4 formation; however, adding Cu had no noticeable effect on increasing H 2 production 32 . Use of a Ce doped Ni-based catalyst also limits CH 4 formation, which could be the result of a synergetic effect between Ni and Ce 31 . Calcia and yttria have been added to ZrO 2 in Ni-based catalysts to increase the surface basicity. This resulted in promoted WGS activity that decreased the selectivity of alkane formation 45 . Table 4 summarizes secondary substances producing rate on catalysts used in MeOH APR. It is clear from this Table that Ni-based catalysts display higher alkane rate production than the Pt-based catalysts. This is in accordance with Figure 3 that indicates that Ni displays a higher methanation rate than Pt.
Table 4. Alkane (CH 4 and C 2H 6) producing rate on various methanol aqueous-phase reforming catalysts.
(PR – production rate, NO – not observed).
Metal | Support | Metal
loading (wt. %) |
Temperature
(°C) |
Pressure
(MPa) |
WHSV
(h -1) |
CH
4
yield (%) |
CH
4
(mol. %) |
CH
4 PR
(µmol/ min/g catalyst) |
CH
4
selectivity (%) |
C
2H
6
yield (%) |
C
2H
6 PR
(µmol/min/ g catalyst) |
Ref. |
---|---|---|---|---|---|---|---|---|---|---|---|---|
Pt | ɣ- Al 2O 3 | 3 | 225 | 2.5 | 0.008 | 0.4 | 0.0 | 18 | ||||
Pt | ɣ- Al 2O 3 | 3 | 267 | 2.5 | 0.008 | 0.6 | 0.0 | 18 | ||||
Pt | Al 2O 3 | 2 | 190 | 6 | 27 | |||||||
Pt | ɣ- Al 2O 3 | 0.97 | 37 | |||||||||
Pt | ɣ-Al 2O 3 | 8 | 230 | 3.2 | 2 | 0.1 | 0.9 | NO | NO | 32 | ||
Pt | ɣ- Al 2O 3 | 3 | 224 | 2.65 | 0.008 | 0.5 | 26 | |||||
Pt | ɣ- Al 2O 3 | 3 | 264 | 2.65 | 0.008 | 0.6 | 26 | |||||
Pt | ɣ-Al 2O 3 | 1 | 210 | 2.9 | 2.94 | 40 | ||||||
Pt | δ- Al 2O 3 | 1.01 | 37 | |||||||||
Pt | α- Al 2O 3 | 1.10 | 37 | |||||||||
Pt | CeO 2 | 0.20 | 37 | |||||||||
Pt | α-Fe 2O 3 | 0.02 | 37 | |||||||||
Pt | α-MoC | 2.2 | 190 | 6 | 27 | |||||||
Pt | α-MoC | 2 | 170 | 6 | 27 | |||||||
Pt | α-MoC | 2 | 150 | 6 | 27 | |||||||
Pt | α-MoC | 0.2 | 190 | 6 | 27 | |||||||
Pt | β-MoC | 2.1 | 190 | 6 | 27 | |||||||
Pt | MoC-2 | 2 | 190 | 6 | 27 | |||||||
Pt | MoC-3 | 2.3 | 190 | 6 | 27 | |||||||
Pt | NaY | 0.5 | 224 | 2.65 | 0.08 | 0.6 | 26 | |||||
Pt | NaY | 0.5 | 264 | 2.65 | 0.08 | 0.8 | 26 | |||||
Pt | NiAl2O4 | 1 | 210 | 2.9 | 2.94 | 40 | ||||||
Pt | TiO 2 | 2 | 190 | 6 | 27 | |||||||
Ni | ZrO 2 | 10 | 230 | 3.2 | 2 | 1.3 | 13.4 | 0.1 | 1.5 | 32 | ||
Ni | ZrO 2 | 7 | 230 | 3.2 | 4 | 2.2 | 43 | |||||
Ni | ZrO 2 | 10 | 230 | 3.2 | 3.6 | 1.5 | 42 | |||||
Ni | [CeO 2] 0.17-[ZrO 2] 0.83 | 10 | 230 | 3.2 | 3.6 | 2.2 | 42 | |||||
Ni | [CeO 2] 0.25-[ZrO 2] 0.75 | 10 | 230 | 3.2 | 3.6 | 3.3 | 42 | |||||
Ni | [CeO
2]
0.17-
[La 2O 3]0.05- [ZrO 2] 0.78 |
10 | 230 | 3.2 | 3.6 | 2.0 | 42 | |||||
Ni | [La
2O
3]
0.10-
[ZrO 2] 0.90 |
10 | 230 | 3.2 | 3.6 | 2.6 | 42 | |||||
Ni | Calcia-
stabilized ZrO 2 |
5 | 230 | 3.2 | 4 | 2.1 | 43 | |||||
Ni | Yttria-
stabilized ZrO 2 |
5.6 | 230 | 3.2 | 4 | 1.5 | 43 | |||||
Ni | Calcia-
stabilized ZrO 2 |
6.9 | 230 | 3.2 | 4 | 2.1 | 43 | |||||
Ni | CeO2 | 10 | 230 | 3.2 | 3.6 | 0.0 | 42 | |||||
Ni | α-Al 2O 3 | 11.8 | 230 | 3.2 | 2 | 0.3 | 3.4 | 32 | ||||
Ni | SiO 2 | 10 | 230 | 3.2 | 2 | 0.2 | 2.5 | 32 | ||||
Ni | ɣ-Al 2O 3 | 11.4 | 230 | 3.2 | 2 | 3.3 | 33.4 | 0.2 | 2.1 | 32 | ||
Ni-Cu | [CeO
2]
0.25-
[ZrO 2] 0.75 |
Ni 10
Cu 5 |
230 | 3.2 | 3.6 | 31 | ||||||
Ni-Cu | SiO 2 | Ni 10
Cu 5 |
230 | 3.2 | 2 | 0.0 | 0.0 | NO | NO | 32 | ||
Ni-Cu | Al 2O 3 | Ni 13
Cu 1.3 |
230 | 3.2 | 2 | 0.7 | 7.3 | 0.0 | 0.1 | 32 | ||
Ni-Ce | [CeO
2]
0.25-
[ZrO 2] 0.75 |
Ni 10
Ce 5 |
230 | 3.2 | 3.6 | 31 | ||||||
Ni-Ce | Al 2O 3 | Ni 13
Ce 1.3 |
230 | 3.2 | 3.6 | 31 | ||||||
Ni-Ce | SiO 2 | Ni 10
Cu 5 |
230 | 3.2 | 2 | 0.7 | 7.4 | 0.0 | 0.2 | 32 | ||
Ni-Ce | Al 2O 3 | Ni 13
Cu 1.3 |
230 | 3.2 | 2 | 3.2 | 32.1 | 0.2 | 2.0 | 32 | ||
Ni-Ce | ZrO 2 | Ni 5.9
Ce 12.6 |
230 | 3.2 | 4 | 0.9 | 43 | |||||
Ni-Ce | Calcia-
stabilized ZrO 2 |
Ni 5.9
Ce 13 |
230 | 3.2 | 4 | 2.2 | 43 | |||||
Ni-Ce | Calcia-
stabilized ZrO 2 |
Ni 5.5
Ce 13.5 |
230 | 3.2 | 4 | 2.5 | 43 | |||||
Ni-Ce | Yttria-
tabilized ZrO 2 |
Ni 5.5
Ce 12.7 |
230 | 3.2 | 4 | 0.9 | 43 |
Effect of experimental conditions
Methanol APR can be carried out in the range 200°C – 250°C within the pressure range of 1.5 MPa – 3.5 MPa. Therefore, it becomes necessary to observe the effect of system pressure and temperature for optimized H 2 production. Pan et al. studied the effect of pressure on H 2 production along with the production rates of CO 2 and CH 4 in APR of 5 wt. % ethylene glycol aqueous solution for Ni 20SnAl 7 hydrotalcite derived catalyst. It was reported that H 2 production decreased significantly with increased pressure (from 2.6 MPa to 3.0 MPa) whereas the production rate of CH 4 was unaffected. The decrease in CO 2 production was also observed with increasing operating pressure. This effect was ascribed to blockage of available catalytic sites and results in dehydration and hydrogenation of intermediate products at the expense of produced H 2. This results in the decrease of H 2 production 52 .
Additionally, partial H 2 pressure becomes a limiting factor to H 2 production. Methanol APR forms H 2, CO, CO 2 and sometimes CH 4 in gaseous form, which is in the form of gas bubbles within the aqueous solution. The total pressure inside the bubble is calculated to be the same as the total system pressure considering vapor-liquid equilibrium 19 . The high H 2 pressure blocks catalytic sites and lowers the surface concentration of CO, which reduces the activity of the WGS reaction. Additionally, changes in total system pressure influence the gaseous bubble density and size distribution. Such changes affect the performance of the catalyst by modifying their wetting behavior 53 . In the same manner as pressure, system temperature also influences H 2 production. For example, at high temperature water can corrode the catalyst and active metal and/or support can be washed away into the effluent 19 . However, thermodynamic analysis of methanol APR show higher H 2 selectivity at higher temperatures (>187°C). On the contrary, CH 4 formation decreases with temperature increase (and increases with pressure increase). Since CH 4 formation is an exothermic process, the methanation reaction is thermodynamic unfavored by higher temperatures 54 .
Reactor types
The catalyst and reactor design must be accurately selected for obtaining the best APR performance. APR is a three phase reaction which reaction kinetic is often controlled by the diffusion of reactants and products in the liquid phase. Namely hydrogen, which has a very low solubility in methanol and in water, diffuses slowly from the reaction center where it is formed to the gas phase, outside the catalyst; the reaction is then normally controlled by diffusion 34 . Additionally, the evolved hydrogen may take part in the methanation and/or Fischer-Tropsch side reactions that further decreases the H2 production rate 34 . Therefore, catalyst and reactor should be designed to facilitate the fast hydrogen removal from to maximize the hydrogen production rate. Most studies consider the use of fixed packed reactors for MeOH APR, but this design is certainly not the best suitable for this reaction. In a typical lab-scale APR experimental set-up, an aqueous MeOH solution is fed upwards in a packed bed reactor using an HPLC pump. Before starting the reaction, whenever needed hydrogen is used as a reducing medium for catalyst reduction. Normally nitrogen, but also argon, are then used as sweep gas as to strip away the produced hydrogen. A condenser is used to separate the gas products from the unreacted methanol and water. The products are analyzed with an online gas chromatograph and unreacted reactants can be analyzed with a liquid chromatograph. A pressure regulator is used to maintain the operating pressure of APR such as methanal and water are in liquid phase. Figure 6 shows a schematic representation of a typical MeOH APR reaction set-up using a fix bed reactor.
Figure 6. Schematic representation of methanol aqueous-phase reforming experimental set-up.
Microchannel and membrane reactors have also been considered for the APR reaction. These reactors allow higher mass transfer rates of reactants and products and then higher selectivities and hydrogen production rates. D’Angelo et al. published very interesting studies on the use of carbon-coated ceramic membrane reactors 55 and coated microchannel reactors 56 for APR of sorbitol and ethylene glycol, respectively. The carbon-coated ceramic membrane showed 2.5 times higher hydrogen yield when compared with the fixed bed reactor tested under similar experimental conditions. Since some metals selective possess H 2 permeability, due to this fact, membrane reactor can be facilitated as both reaction and purification system in reforming reactors 57, 58 . The membrane reactor facilitates the effective removal of hydrogen from the reaction medium due to the differential pressure between retentate and permeate. This product removal from reaction system can increase H2 yield. Additionally, the authors pointed that the hydrophobic nature of membrane hinders its interaction with water, therefore increasing stability of membrane. The coated microchannel reactor displayed a similar performance when compared with the membrane reactor 55 . On another fascinating study, D’Angelo et al. 56 coated Pt/AlO(OH) catalyst particles on the inner wall of a microchannel (inner diameter of 320 µm) to perform the APR of ethylene glycol. This single microchannel reactor showed increased hydrogen selectivity and ethylene glycol conversion compared to the fixed bed reactor performance used under identical conditions. The authors attributed this greater performance to the hydrodynamic behavior under Taylor flow regime, which facilitates the removal of hydrogen from catalyst surface; such reactors should also be assessed for MeOH APR.
APR kinetics and modeling
Kinetic modelling is a way to shed light on the reaction pathways leading to rate of formation of the products of various catalysts at given temperatures. Ribeirinha et al. 59 studied empirical and mechanistic models for the high performance catalyst CuO/ZnO/Ga 2O 3 for methanol steam reforming (MSR). Simulating a packed bed reactor, they found that these models accurately fit to the experimental results, after appropriate parameter estimation. Lotrič et al. 60 developed a power-law empirical model for the same catalyst CuO/ZnO/Ga 2O 3 and reported good fitting with experimental data. There are several kinetic models reported in the literature for MSR for various catalysts, reaction temperatures and types of reactor 61– 67 In contrast, APR kinetic models are scarce as the technology is still in its development phase. Coronado et al. 68 presented a kinetic model for C 1-C 4 alcohols derived from Fischer-Tropsch synthesis over Ni supported on ceria-zirconia using power law and phenomenological models. The authors evaluated three different reaction rate equations. The first equation was a Langmuir–Hinshelwood reaction rate expression based on the work by Murzin et al. 69 for xylitol APR ( Equation 5), and the two other equations ( Equation 6– Equation 7) were simple power law models.
The authors considered only alcohols as being able to adsorb on the active sites of the catalyst. The adsorption equilibrium constants of alcohols ( K), the rate constants ( k) and the reaction orders ( m) were considered as lumped parameters to avoid system over-parametrization. Water concentration was considered constant and the WGS reaction was assumed to be dependent on the concentration of CO and H 2O. The Arrhenius and Van’t Hoff equations were used to calculate the temperature dependency of the reaction rate constants and adsorption equilibrium constants, respectively 68 . However, the availability of APR kinetics modelling is still limited and in its preliminary stage. Further study and development in the field of APR kinetic models are required in order to better understand APR reaction kinetics that can help optimize and upscale APR.
Feasibility of reforming for fuel cell applications
Low-temperature PEMFCs (LT-PEMFCs) such as Nafion based polymer electrolyte membrane fuel cells are state-of-the-art fuel cells that operate below 100°C. Despite the excellent proton conductivity, the poor CO tolerance at the anode, slow oxygen reduction reaction kinetics and complex water management system hinders cost reduction and performance improvement of LT-PEMFCs. Some studies reveal that LT-PEMFCs suffer significant performance loss in the presence of > 10-20 ppm of CO in fuel feed 70 . Operating at temperatures above 100°C has the benefits of enhanced catalytic activities and simplified water management. HT-PEMFCs are also more tolerant to fuel impurities and have low dependency on heat removal systems 71, 72 . HT-PEMFCs also have a better CO tolerance compared to LT-PEMFCs. A HT-PEMFC operating at 130°C can tolerate up to 1000 ppm of CO compared to LT-PEMFCs operating temperature 80°C, which can tolerate only up to 10 ppm of CO 70, 73 . Such attributes of HT-PEMFCs provide the potential of in situ hydrogen production using fuel processors based on reforming or oxidation processes. One of the most important reasons for in situ H 2 production is the non-availability of a viable H 2 storage material 74– 76 . In this sense, in situ H 2 production has been extensively researched in recent years as a source of H 2 feedstock for fuel cells. The integration of fuel cells with a steam reformer is thermally advantageous since fuel cells are exothermic, while reformers are endothermic. Several studies investigating the thermal integration of low temperature steam reforming with HT-PEMFCs can be found in the literature 23, 77– 81 . However, the operating temperature remains a hurdle to overcome in the integration of steam reformers and HT-PEMFCs. HT-PEMFCs have the operating temperature range of 120°C – 200°C 82 , whereas the steam reforming of alcohols occurs at temperatures between 150°C and 500°C 83– 85 . The operating temperature for state-of-the-art HT-PEMFCs cannot exceed 200°C, as it would lead to problems such as degradation and loss of proton conductivity. Additionally, the temperature difference between HT-PEMFCs and steam reformers is more of a challenge for rapid startup 86, 87 .
Fuel processors for fuel cells applications
Fuel processors are complex devices that convert hydrocarbons into H 2. A typical fuel processor consists of a steam generator, preheater, catalytic fuel reformer, WGS reactor, exhausters, condenser and catalytic after burner ( Figure 7). Catalytic fuel reforming is undoubtedly the most crucial part of the cogeneration system as it produces H 2 for the fuel cell. A vaporizer or heater is another important component of the fuel processor that is required to heat fuel and water to meet the catalytic reaction conditions of the hydrocarbon. Many fuel processers are equipped with CO removal systems, such as WGS reactors 88, 89 . When using LT-PEMFCs, it becomes critical to feed in pure H 2 to avoid the contamination of the fuel cell anode catalyst; therefore, more complex purification systems are required, such as pressure-swing adsorption (PSA) units. Since methanol reforming involves production of CO 2, it is critical to install a CO 2 capture system at the outlet to avoid adding to CO 2 emissions 90– 92 .
Figure 7. Schematic diagram of an integrated fuel (methanol) processor with fuel cell.
Reproduced from 93 (open access, Creative Commons Attribution (CC BY 4.0) license. Licensee MDPI, Basel, Switzerland.).
Fuel processors make the overall HT/LT-PEMFCs a complicated system, as it is vital to maintain the temperatures of various subsystems. This in turn can make the thermal integration of the system a challenging task (especially for relatively small fuel cell systems) since the temperatures for different subsystems can vary across wide ranges. The thermal efficiency of a fuel processor ( ε FP) ( Equation 8) is calculated using the higher heating value (HHV) of H 2 produced via reforming divided by the HHV of the input fuel (e.g. methanol). The total thermal efficiency of the fuel processor is brought down by the energy required to run the fuel processor subsystem (e.g. pumps, condenser, exhaust fans, etc.). Brett et al. 89 reported that if one counts the (electrical) efficiency of the whole system, in reality it is one third to one fifth lower than the efficiency counted for only the fuel cell stack.
APR integration with HT-PEMFC
Internal reforming of methanol or integration of a methanol reformer with a fuel cell is distinctly advantageous compared to fuel processor systems due to the complex thermal management, water management and eventual loss in overall efficiency due to many subsystem components. To overcome this issue, researchers have studied lowering the temperature of MSR for integration with HT-PEMFC 24, 94– 96 . Lotrič et al. 80 followed a different approach, by considering a higher HT-PEMFC stack operating temperature of 255°C. The authors investigated the integration of MSR with fuel cells and reported that HT-PEMFC shows higher system efficiency than LT-PEMFC in an integrated system. However, operating fuel cell stacks at 255°C requires novel electrolytes capable of proton conduction at such a temperature in absence of water, using expensive gaskets to withstand high temperatures, and development of a new catalyst support, as widely used carbon supports corrode faster at high temperatures 88 .
These issues can be avoided, to some extent, via methanol APR. Integrating methanol APR with HT-PEMFCs can be seen as a potential technology to increase efficiency and to overcome the complexity of MSR integration into HT-PEMFCs. As discussed earlier, methanol APR, using catalysts such as Pt/MoC, can be performed at temperatures below 200°C with considerably high H 2 production 27 . Additionally, APR is an exothermic process, which makes its integration with fuel cell a simpler and more efficient method of H 2 and electricity production 80 . Using methanol APR with HT-PEMFCs is advantageous since the undesired byproducts of reforming are easy to control, except for CH 4 and CO, where the concentration is below the limits of HT-PEMFC tolerance. Integrating APR with HT-PEMFCs can also make the whole system more compact since methanol APR is an easy single-step, single reactor process. The methanol APR reaction conditions are far better suited for integration with HT-PEMFCs than MSR. APR integrated HT-PEMFCs have potential for higher CHP fuel cell system efficiency and performance than MSR integrated HT-PEMFCs.
Challenges with integration. Although the integration of APR and HT-PEMFCs has several advantages, there are also distinct barriers. Challenges are related to the low maturity of the APR technology, elementary APR catalyst research, and the feasible conditions of the integrated system.
To improve the overall efficiency of the HT-PEMFC system and to make the integration of APR and HT-PEMFCs more attractive, the waste heat recovered from the FC stack should be utilized as the heat source of the APR unit. In thermal integration, the waste heat of the FC stack is utilized with a heat transfer fluid recycled in the overall system. Thermal integration would enhance the benefits of APR and compensate for the lower methanol conversions and hydrogen production rates compared to the high values attained with MSR technology. Due to thermal losses in the heat recovery process, the operation temperature of the APR unit is below the operation temperature of the FC stack. Moreover, even though the operation temperature of the HT-PEMFC may vary in the range of 120–200°C, the degradation of the stack is faster at higher operation temperatures 97 . Therefore, in the thermal integration of APR and HT-PEMFC technology, the selection of operation temperature is based on the optimization of APR catalyst activity and stack lifetime.
Since the reforming temperature should remain below 200°C, a major challenge is the lack of data related to APR systems at decreased temperatures. As presented in Table 1 and Table 2, the methanol conversion and hydrogen production rate drop when the reforming temperature is decreased. A high conversion rate is essential for the system to reach high system efficiency. Furthermore, the anode of the HT-PEMFC tolerates only limited amounts of vaporized feed solution; even a 3% volume fraction of methanol-water vapor would degrade the FC performance 98 . To avoid this, either a more efficient catalyst or a cold trap for vapor collection is necessary for the integrated system.
Additionally, a challenge with the APR catalysts is the lack of long-term activity experiments. Li et al. 27 reached excellent hydrogen production rates with a Pt/α-MoC catalyst, but their experiments lasted less than two hours, which provides no information on long-term catalyst activity. When the APR unit is integrated with the HT-PEMFC stack, the stability of the APR catalyst should correspond with the durability of the HT-PEM membrane electrode assemblies (MEAs), which have recently reached a lifetime of 9,000 hours without any problems 99 . In situ re-reduction of deactivated catalyst demands more complexity from the FC system and therefore it should be avoided if possible.
Even though high-pressure hydrogen storage is not required when hydrogen is produced in situ with APR, the pressure conditions of APR are relatively high. To perform APR, the reforming pressure should be above the bubble point pressure of the feed mixture. If the methanol molar fraction in the feed solution is 30%, the bubble point pressure of the solution is over 9 bar at 150°C and over 29 bar at 200°C 100, 101 . Since the HT-PEMFC is operated at around atmospheric pressure, the reforming pressure should be maintained with a back-pressure regulator.
In three-phase reforming systems, such as APR, mass transfer limitations cause challenges. Typically, they are solved with nitrogen co-feeding, which enhances the removal of hydrogen molecules from the porous catalyst surfaces. Nevertheless, when the APR is integrated with the HT-PEMFC, nitrogen co-feeding would dilute the feed gas, which is unfavorable. The separation of nitrogen and hydrogen is possible with novel membrane separation techniques 102, 103 . However, such a separation system would increase the complexity and capital expenditure (CAPEX) of the FC system, and thus is not considered a reasonable solution. Instead, usage of a micro-channel reactor or novel catalyst coating on a structured catalyst have been suggested to improve the mass transfer properties of the APR system 20 .
Conclusion
In recent years, research on methanol APR for hydrogen production has been increasing, with a focus on developing new catalysts. Pt-based catalysts are the best performing catalysts and the highest hydrogen production was reported using MoC and NiAl 2O 4 as supports. Unconventional supports such as Zeolite and MoC performed better than conventional supports such as alumina and silica. Non-noble metal-based catalysts, such as Ni-based catalysts, show promising results and are the focus of most methanol APR catalytic developments. Ni has been supported on various metal oxides such as CeO 2 and ZrO 2, which increased overall hydrogen production by promoting WGS reaction. The addition of Cu to produce a bi-metallic Ni-based catalyst leads to higher WGS activity and suppressed methanation, whereas adding Ce only leads to increased hydrogen yield. Ru, Re and Ir are other less frequently used metals that showed promising results and need to be further explored.
Methanol APR has the potential to be considered as a prominent hydrogen production technology. However, methanol APR using non-noble metal catalysts need further development to reach Pt-based catalyst activity. The challenges of the catalyst development for APR is achieving high H 2 production under APR operating temperatures. By using various dopants and novel supports, to some extent, the hydrothermal stability and performance of catalyst can be increased. Therefore, catalyst development should be focused on a highly durable and active catalyst under APR reaction conditions. Additionally, mass transfer limitation also remains a huge hurdle in APR that limits the APR performance. Therefore, focus should also be given in studying various reactors design to minimize mass transport limitation. Furthermore, kinetic models need to be developed for better understanding of APR reaction mechanisms, as well as being important tools for scaling, designing and optimizing processes.
Data availability
No data are associated with this article.
Funding Statement
This research was financially supported by the European Union’s Horizon 2020 research and innovation programme under the grant agreement No 875081 (project EMPOWER). This project has also received Base Funding - UIDB/00511/2020 of the Laboratory for Process Engineering, Environment, Biotechnology and Energy – LEPABE - funded by national funds through the FCT/MCTES (PIDDAC).
The funders had no role in study design, data collection and analysis, decision to publish, or preparation of the manuscript.
[version 3; peer review: 2 approved]
References
- 1. U.S. Energy Information Administration: International Energy Outlook 2019. 2019; [cited 2020 Mar 13]. Reference Source [Google Scholar]
- 2. Sustainable Development Goals: UNITED NATIONS. [cited 2020 Nov 9]. Reference Source [Google Scholar]
- 3. Ritchie H, Roser M: Energy. Our World Data, 2020. Reference Source [Google Scholar]
- 4. Abe JO, Popoola API, Ajenifuja E, et al. : Hydrogen energy, economy and storage: Review and recommendation. Int J Hydrogen Energy. 2019;44(29):15072–86. 10.1016/j.ijhydene.2019.04.068 [DOI] [Google Scholar]
- 5. Dawood F, Anda M, Shafiullah GM: Hydrogen production for energy: An overview. Int J Hydrogen Energy. 2020;45(7):3847–69. 10.1016/j.ijhydene.2019.12.059 [DOI] [Google Scholar]
- 6. Cany C, Mansilla C, da Costa P, et al. : Adapting the French nuclear fleet to integrate variable renewable energies via the production of hydrogen: Towards massive production of low carbon hydrogen? Int J Hydrogen Energy. 2017;42(19):13339–56. 10.1016/j.ijhydene.2017.01.146 [DOI] [Google Scholar]
- 7. Rusman NAA, Dahari M: A review on the current progress of metal hydrides material for solid-state hydrogen storage applications. Int J Hydrogen Energy. 2016;41(28):12108–26. 10.1016/j.ijhydene.2016.05.244 [DOI] [Google Scholar]
- 8. Hwang HT, Varma A: Hydrogen storage for fuel cell vehicles. Curr Opin Chem Eng. 2014;5:42–8. 10.1016/j.coche.2014.04.004 [DOI] [Google Scholar]
- 9. Martin A, Agnoletti MF, Brangier E: Users in the design of Hydrogen Energy Systems: A systematic review. Int J Hydrogen Energy. 2020;45(21):11889–900. 10.1016/j.ijhydene.2020.02.163 [DOI] [Google Scholar]
- 10. Hirscher M, Yartys VA, Baricco M, et al. : Materials for hydrogen-based energy storage - past, recent progress and future outlook. J Alloys Compd. 2020;827:153548. 10.1016/j.jallcom.2019.153548 [DOI] [Google Scholar]
- 11. Jiang J, Zhu X, Qian H, et al. : Cross-linked sulfonated poly(ether ether ketone) electrolytes bearing pendent imidazole groups for high temperature proton exchange membrane fuel cells. Sustain Energ Fuels. 2019;3(9):2426–34. 10.1039/C9SE00318E [DOI] [Google Scholar]
- 12. Han M, Zhang G, Li M, et al. : Considerations of the morphology in the design of proton exchange membranes: Cross-linked sulfonated poly(ether ether ketone)s using a new carboxyl-terminated benzimidazole as the cross-linker for PEMFCs. Int J Hydrogen Energy. 2011;36(3):2197–206. 10.1016/j.ijhydene.2010.11.065 [DOI] [Google Scholar]
- 13. Ajanovic A, Haas R: Prospects and impediments for hydrogen and fuel cell vehicles in the transport sector. Int J Hydrogen Energy. 2021;46(16):10049–10058. 10.1016/j.ijhydene.2020.03.122 [DOI] [Google Scholar]
- 14. Kubacka A, Fernández-García M, Martínez-Arias A: Catalytic hydrogen production through WGS or steam reforming of alcohols over Cu, Ni and Co catalysts. Appl Catal A Gen. 2016;518:2–17. 10.1016/j.apcata.2016.01.027 [DOI] [Google Scholar]
- 15. Bepari S, Kuila D: Steam reforming of methanol, ethanol and glycerol over nickel-based catalysts-A review. Int J Hydrogen Energy. 2020;45(36):18090–18113. 10.1016/j.ijhydene.2019.08.003 [DOI] [Google Scholar]
- 16. Sá S, Silva H, Brandão L, et al. : Catalysts for methanol steam reforming-A review. Appl Catal B Environ. 2010;99(1–2):43–57. 10.1016/j.apcatb.2010.06.015 [DOI] [Google Scholar]
- 17. Davda RR, Shabaker JW, Huber GW, et al. : A review of catalytic issues and process conditions for renewable hydrogen and alkanes by aqueous-phase reforming of oxygenated hydrocarbons over supported metal catalysts. Appl Catal B Environ. 2005;56(1–2):171–86. 10.1016/j.apcatb.2004.04.027 [DOI] [Google Scholar]
- 18. Cortright RD, Davda RR, Dumesic JA: Hydrogen from catalytic reforming of biomass-derived hydrocarbons in liquid water. Mater Sustain Energy A Collect Peer-Reviewed Res Rev Artic from Nat Publ Gr. 2010;418:289–92. 10.1142/9789814317665_0043 [DOI] [PubMed] [Google Scholar]
- 19. Shabaker JW, Davda RR, Huber GW, et al. : Aqueous-phase reforming of methanol and ethylene glycol over alumina-supported platinum catalysts. J Catal. 2003;215(2):344–52. 10.1016/S0021-9517(03)00032-0 [DOI] [Google Scholar]
- 20. Coronado I, Stekrova M, Reinikainen M, et al. : A review of catalytic aqueous-phase reforming of oxygenated hydrocarbons derived from biorefinery water fractions. Int J Hydrogen Energy. 2016;41(26):11003–32. 10.1016/j.ijhydene.2016.05.032 [DOI] [Google Scholar]
- 21. Davda RR, Shabaker JW, Huber GW, et al. : Aqueous-phase reforming of ethylene glycol on silica-supported metal catalysts. Appl Catal B Environ. 2003;43(1):13–26. 10.1016/S0926-3373(02)00277-1 [DOI] [Google Scholar]
- 22. Godina LI, Tokarev AV, Simakova IL, et al. : Aqueous-phase reforming of alcohols with three carbon atoms on carbon-supported Pt. Catal Today. 2018;301:78–89. 10.1016/j.cattod.2017.03.042 [DOI] [Google Scholar]
- 23. Iulianelli A, Ribeirinha P, Mendes A, et al. : Methanol steam reforming for hydrogen generation via conventional and membrane reactors: A review. Renew Sustain Energy Rev. 2014;29:355–68. 10.1016/j.rser.2013.08.032 [DOI] [Google Scholar]
- 24. Cai F, Ibrahim JJ, Fu Y, et al. : Low-temperature hydrogen production from methanol steam reforming on Zn-modified Pt/MoC catalysts. Appl Catal B Environ. 2020;264:118500. 10.1016/j.apcatb.2019.118500 [DOI] [Google Scholar]
- 25. Hohn KL, Lin YC: Catalytic partial oxidation of methanol and ethanol for hydrogen generation. ChemSusChem. 2009;2(10):927–40. 10.1002/cssc.200900104 [DOI] [PubMed] [Google Scholar]
- 26. Tang Z, Monroe J, Dong J, et al. : Platinum-loaded NaY zeolite for aqueous-phase reforming of methanol and ethanol to hydrogen. Ind Eng Chem Res. 2009;48(5):2728–33. 10.1021/ie801222f [DOI] [Google Scholar]
- 27. Lin L, Zhou W, Gao R, et al. : Low-Temperature hydrogen production from water and methanol using Pt/α-MoC catalysts. Nature. 2017;544(7648):80–3. 10.1038/nature21672 [DOI] [PubMed] [Google Scholar]
- 28. Sinfelt JH, Yates DJC: Catalytic hydrogenolysis of ethane over the noble metals of Group VIII. J Catal. 1967;8(1):82–90. 10.1016/0021-9517(67)90284-9 [DOI] [Google Scholar]
- 29. Grenoble DC, Estadt MM, Ollis DF: The chemistry and catalysis of the water gas shift reaction: 1. The kinetics over supported metal catalysts. J Catal. 1981;67(1):90–102. 10.1016/0021-9517(81)90263-3 [DOI] [Google Scholar]
- 30. Vaidya PD, Lopez-Sanchez JA: Review of Hydrogen Production by Catalytic Aqueous-Phase Reforming. ChemistrySelect. 2017;2(22):6563–76. 10.1002/slct.201700905 [DOI] [Google Scholar]
- 31. Coronado I, Pitínová M, Karinen R, et al. : Aqueous-phase reforming of Fischer-Tropsch alcohols over nickel-based catalysts to produce hydrogen: Product distribution and reaction pathways. Appl Catal A Gen. 2018;567:112–21. 10.1016/j.apcata.2018.09.013 [DOI] [Google Scholar]
- 32. Coronado I, Stekrova M, Moreno LG, et al. : Aqueous-phase reforming of methanol over nickel-based catalysts for hydrogen production. Biomass Bioenergy. 2017;106:29–37. 10.1016/j.biombioe.2017.08.018 [DOI] [Google Scholar]
- 33. Xiang Y, Li X, Lu C, et al. : Reaction performance of hydrogen from aqueous-phase reforming of methanol or ethanol in hydrogenation of phenol. Ind Eng Chem Res. 2011;50(6):3139–44. 10.1021/ie101411h [DOI] [Google Scholar]
- 34. Shabaker JW, Dumesic JA: Kinetics of aqueous-phase reforming of oxygenated hydrocarbons: Pt/Al 2O 3 and Sn-modified Ni catalysts. Ind Eng Chem Res. 2004;43(12):3105–12. 10.1021/ie049852o [DOI] [Google Scholar]
- 35. Vannice MA: The catalytic synthesis of hydrocarbons from H 2 CO mixtures over the Group VIII metals. V. The catalytic behavior of silica-supported metals. J Catal. 1977;50(2):228–236. 10.1016/0021-9517(77)90031-8 [DOI] [Google Scholar]
- 36. Nielsen M, Alberico E, Baumann W, et al. : Low-temperature aqueous-phase methanol dehydrogenation to hydrogen and carbon dioxide. Nature. 2013;495(7439):85–9. 10.1038/nature11891 [DOI] [PubMed] [Google Scholar]
- 37. Chen GY, Li WQ, Chen H, et al. : Progress in the aqueous-phase reforming of different biomass-derived alcohols for hydrogen production. J Zhejiang Univ Sci A. 2015;16(6):491–506. 10.1631/jzus.A1500023 [DOI] [Google Scholar]
- 38. Jongerius AL, Copeland JR, Foo GS, et al. : Stability of Pt/γ-Al 2O 3 catalysts in lignin and lignin model compound solutions under liquid phase reforming reaction conditions. ACS Catal. 2013;3(3):464–73. 10.1021/cs300684y [DOI] [Google Scholar]
- 39. El Doukkali M, Iriondo A, Cambra JF, et al. : Deactivation study of the Pt and/or Ni-based γ-Al 2O 3 catalysts used in the aqueous phase reforming of glycerol for H 2 production. Applied Catalysis A: General. 2014;472:80–91. 10.1016/j.apcata.2013.12.015 [DOI] [Google Scholar]
- 40. Li D, Li Y, Liu X, et al. : NiAl 2O 4 Spinel Supported Pt Catalyst: High Performance and Origin in Aqueous-Phase Reforming of Methanol. ACS Catal. 2019;9(10):9671–82. 10.1021/acscatal.9b02243 [DOI] [Google Scholar]
- 41. Ziemniak SE, Guilmette PA, Turcotte RA, et al. : Oxidative dissolution of nickel metal in hydrogenated hydrothermal solutions. Corros Sci. 2008;50(2):449–62. 10.1016/j.corsci.2007.07.013 [DOI] [Google Scholar]
- 42. Stekrova M, Rinta-Paavola A, Karinen R: Hydrogen production via aqueous-phase reforming of methanol over nickel modified Ce, Zr and La oxide supports. Catal Today. 2018;304:143–52. 10.1016/j.cattod.2017.08.030 [DOI] [Google Scholar]
- 43. Goma D, Delgado JJ, Lefferts L, et al. : Catalytic Performance of Ni/CeO 2/X-ZrO 2 (X = Ca, Y) Catalysts in the Aqueous-Phase Reforming of Methanol. Nanomaterials (Basel). 2019;9(11):1582. 10.3390/nano9111582 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 44. Han Y, Zhu J: Surface science studies on the zirconia-based model catalysts. Top Catal. 2013;56(15–17):1525–41. 10.1007/s11244-013-0156-5 [DOI] [Google Scholar]
- 45. Van Haasterecht T, Ludding CCI, De Jong KP, et al. : Stability and activity of carbon nanofiber-supported catalysts in the aqueous phase reforming of ethylene glycol. J Energy Chem. 2013;22(2):257–69. 10.1016/S2095-4956(13)60032-7 [DOI] [Google Scholar]
- 46. Strobel V, Schuster JJ, Braeuer AS, et al. : Shining light on low-temperature methanol aqueous-phase reforming using homogeneous Ru-pincer complexes-operando Raman-GC studies. React Chem Eng. 2017;2(3):390–6. 10.1039/C6RE00228E [DOI] [Google Scholar]
- 47. Sakamoto T, Miyao T, Yoshida A, et al. : Effect of Re and Mo addition upon liquid phase methanol reforming with water over SiO 2, ZrO 2 and TiO 2 supported Ir catalysts. Int J Hydrogen Ener. 2010;35(12):6203–9. 10.1016/j.ijhydene.2010.03.113 [DOI] [Google Scholar]
- 48. Sietsma JRA, van Dillen AJ, de Jongh PE, et al. : Application of ordered mesoporous materials as model supports to study catalyst preparation by impregnation and drying. Studies in Surface Science and Catalysis. Elsevier Masson SAS;2006;162:95–102. 10.1016/S0167-2991(06)80895-5 [DOI] [Google Scholar]
- 49. Mehrabadi BAT, Eskandari S, Khan U, et al. : A Review of Preparation Methods for Supported Metal Catalysts.1st ed. Advances in Catalysis. Elsevier Inc.;2017;61:1–35. 10.1016/bs.acat.2017.10.001 [DOI] [Google Scholar]
- 50. Guo Y, Azmat MU, Liu X, et al. : Effect of support’s basic properties on hydrogen production in aqueous-phase reforming of glycerol and correlation between WGS and APR. Appl Energy. 2012;92:218–23. 10.1016/j.apenergy.2011.10.020 [DOI] [Google Scholar]
- 51. Huber GW, Cortright RD, Dumesic JA: Renewable alkanes by aqueous-phase reforming of biomass-derived oxygenates. Angew Chem Int Ed Engl. 2004;43(12):1549–51. 10.1002/anie.200353050 [DOI] [PubMed] [Google Scholar]
- 52. Pan G, Ni Z, Cao F, et al. : Hydrogen production from aqueous-phase reforming of ethylene glycol over Ni/Sn/Al hydrotalcite derived catalysts. Appl Clay Sci. 2012;58:108–113. 10.1016/j.clay.2012.01.023 [DOI] [Google Scholar]
- 53. Neira D’Angelo MF, Ordomsky V, Van Der Schaaf J, et al. : Aqueous phase reforming in a microchannel reactor: The effect of mass transfer on hydrogen selectivity. Catal Sci Technol. 2013;3(10):2834–2842. 10.1039/C3CY00577A [DOI] [Google Scholar]
- 54. Xie J, Su D, Yin X, et al. : Thermodynamic analysis of aqueous phase reforming of three model compounds in bio-oil for hydrogen production. Int J Hydrogen Energy. 2011;36(24):15561–72. 10.1016/j.ijhydene.2011.08.103 [DOI] [Google Scholar]
- 55. D’Angelo MFN, Ordomsky V, Schouten JC, et al. : Carbon-coated ceramic membrane reactor for the production of hydrogen by aqueous-phase reforming of sorbitol. ChemSusChem. 2014;7(7):2007–15. 10.1002/cssc.201301324 [DOI] [PubMed] [Google Scholar]
- 56. D’Angelo MFN, Ordomsky V, Paunovic V, et al. : Hydrogen production through aqueous-phase reforming of ethylene glycol in a washcoated microchannel. ChemSusChem. 2013;6(9):1708–16. 10.1002/cssc.201200974 [DOI] [PubMed] [Google Scholar]
- 57. Lytkina AA, Orekhova NV, Yaroslavtsev AB: Methanol Steam Reforming in Membrane Reactors. Pet Chem. 2018;58(11):911–22. 10.1134/S096554411811004X [DOI] [Google Scholar]
- 58. Lin YM, Rei MH: Study on the hydrogen production from methanol steam reforming in supported palladium membrane reactor. Catal Today. 2001;67(1–3):77–84. 10.1016/S0920-5861(01)00267-X [DOI] [Google Scholar]
- 59. Ribeirinha P, Mateos-Pedrero C, Boaventura M, et al. : CuO/ZnO/Ga 2O 3 catalyst for low temperature MSR reaction: Synthesis, characterization and kinetic model. Appl Catal B Environ. 2018;221:371–379. 10.1016/j.apcatb.2017.09.040 [DOI] [Google Scholar]
- 60. Lotrič A, Sekavčnik M, Pohar A, et al. : Conceptual design of an integrated thermally self-sustained methanol steam reformer - High-temperature PEM fuel cell stack manportable power generator. Int J Hydrogen Energy. 2017;42(26):16700–16713. 10.1016/j.ijhydene.2017.05.057 [DOI] [Google Scholar]
- 61. Sá S, Sousa JM, Mendes A: Steam reforming of methanol over a CuO/ZnO/Al 2O 3 catalyst, part I: Kinetic modelling. Chem Eng Sci. 2011;66(20):4913–4921. 10.1016/j.ces.2011.06.063 [DOI] [Google Scholar]
- 62. Chiu YJ, Chiu HC, Hsieh RH, et al. : Simulations of hydrogen production by methanol steam reforming. Energy Procedia. 2019;156:38–42. 10.1016/j.egypro.2018.11.081 [DOI] [Google Scholar]
- 63. Ribeirinha P, Abdollahzadeh M, Sousa JM, et al. : Modelling of a high-temperature polymer electrolyte membrane fuel cell integrated with a methanol steam reformer cell. Appl Energy. 2017;202:6–19. 10.1016/j.apenergy.2017.05.120 [DOI] [Google Scholar]
- 64. Israni SH, Harold MP: Methanol steam reforming in single-fiber packed bed Pd-Ag membrane reactor: Experiments and modeling. J Memb Sci. 2011;396(1–2):375–387. 10.1016/j.memsci.2010.12.029 [DOI] [Google Scholar]
- 65. Sari A, Sabziani J: Modeling and 3D-simulation of hydrogen production via methanol steam reforming in copper-coated channels of a mini reformer. J Power Sources. 2017;352:64–76. 10.1016/j.jpowsour.2017.03.120 [DOI] [Google Scholar]
- 66. Chein R, Chen YC, Chung JN: Numerical study of methanol-steam reforming and methanol-air catalytic combustion in annulus reactors for hydrogen production. Appl Energy. 2013;102:1022–1034. 10.1016/j.apenergy.2012.06.010 [DOI] [Google Scholar]
- 67. Hao Y, Du X, Yang L, et al. : Numerical simulation of configuration and catalyst-layer effects on micro-channel steam reforming of methanol. Int J Hydrogen Energy. 2011;36(24):15611–15621. 10.1016/j.ijhydene.2011.09.038 [DOI] [Google Scholar]
- 68. Coronado I, Arandia A, Reinikainen M, et al. : Kinetic modelling of the aqueous-phase reforming of fischer-tropsch water over ceria-zirconia supported nickel-copper catalyst. Catalysts. 2019;9(11):936. 10.3390/catal9110936 [DOI] [Google Scholar]
- 69. Murzin DY, Garcia S, Russo V, et al. : Kinetics, Modeling, and Process Design of Hydrogen Production by Aqueous Phase Reforming of Xylitol. Ind Eng Chem Res. 2017;56(45):13240–13253. 10.1021/acs.iecr.7b01636 [DOI] [Google Scholar]
- 70. Li Q, He R, Gao JA, et al. : The CO Poisoning Effect in PEMFCs Operational at Temperatures up to 200°C. J Electrochem Soc. 2003;150(12):A1599. 10.1149/1.1619984 [DOI] [Google Scholar]
- 71. Sun X, Simonsen SC, Norby T, et al. : Composite Membranes for High Temperature PEM Fuel Cells and Electrolysers: A Critical Review. Membranes (Basel). 2019;9(7):83. 10.3390/membranes9070083 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 72. Lee KS, Maurya S, Kim YS, et al. : Intermediate temperature fuel cells via an ion-pair coordinated polymer electrolyte. Energy Environ Sci. 2018;11(4):979–87. 10.1039/C7EE03595K [DOI] [Google Scholar]
- 73. Gang X, Li Q, Hjuler HA, et al. : Hydrogen Oxidation on Gas Diffusion Electrodes for Phosphoric Acid Fuel Cells in the Presence of Carbon Monoxide and Oxygen. J Electrochem Soc. 1995;142(9):2890–3. 10.1149/1.2048660 [DOI] [Google Scholar]
- 74. Sengodan S, Lan R, Humphreys J, et al. : Advances in reforming and partial oxidation of hydrocarbons for hydrogen production and fuel cell applications. Renew Sustain Energy Rev. 2018;82:761–80. 10.1016/j.rser.2017.09.071 [DOI] [Google Scholar]
- 75. Nazir H, Muthuswamy N, Louis C, et al. : Is the H 2 economy realizable in the foreseeable future? Part II: H 2 storage, transportation, and distribution. Int J Hydrogen Energy. 2020;45(41):20693–708. 10.1016/j.ijhydene.2020.05.241 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 76. Wu W, Ben LK, Abbas S, et al. : Exergy-based modular design of an on-board MeOH-to-H 2 processor for fuel cell vehicles. Int J Hydrogen Energy. 2020;45(38):19880–90. 10.1016/j.ijhydene.2020.05.132 [DOI] [Google Scholar]
- 77. Avgouropoulos G, Paxinou A, Neophytides S: In situ hydrogen utilization in an internal reforming methanol fuel cell. Int J Hydrogen Energy. 2014;39(31):18103–8. 10.1016/j.ijhydene.2014.03.101 [DOI] [Google Scholar]
- 78. Zhang S, Zhang Y, Chen J, et al. : Design, fabrication and performance evaluation of an integrated reformed methanol fuel cell for portable use. J Power Sources. 2018;389:37–49. 10.1016/j.jpowsour.2018.04.009 [DOI] [Google Scholar]
- 79. Andreasen SJ, Kær SK, Sahlin S: Control and experimental characterization of a methanol reformer for a 350 W high temperature polymer electrolyte membrane fuel cell system. Int J Hydrogen Energy. 2013;38(3):1676–84. 10.1016/j.ijhydene.2012.09.032 [DOI] [Google Scholar]
- 80. Lotrič A, Sekavčnik M, Hočevar S: Effectiveness of heat-integrated methanol steam reformer and polymer electrolyte membrane fuel cell stack systems for portable applications. J Power Sources. 2014;270:166–82. 10.1016/j.jpowsour.2014.07.072 [DOI] [Google Scholar]
- 81. Ribeirinha P, Alves I, Vázquez FV, et al. : Heat integration of methanol steam reformer with a high-temperature polymeric electrolyte membrane fuel cell. Energy. 2017;120:468–477. 10.1016/j.energy.2016.11.101 [DOI] [Google Scholar]
- 82. Krishnan NN, Lee S, Ghorpade RV, et al. : Polybenzimidazole (PBI-OO) based composite membranes using sulfophenylated TiO 2 as both filler and crosslinker, and their use in the HT-PEM fuel cell. J Memb Sci. 2018;560:11–20. 10.1016/j.memsci.2018.05.006 [DOI] [Google Scholar]
- 83. Shanmugam V, Neuberg S, Zapf R, et al. : Hydrogen production over highly active Pt based catalyst coatings by steam reforming of methanol: Effect of support and co-support. Int J Hydrogen Energy. 2020;45(3):1658–70. 10.1016/j.ijhydene.2019.11.015 [DOI] [Google Scholar]
- 84. Amiri TY, Ghasemzageh K, Iulianelli A: Membrane reactors for sustainable hydrogen production through steam reforming of hydrocarbons: A review. Chem Eng Process - Process Intensif. 2020;157:108148. 10.1016/j.cep.2020.108148 [DOI] [Google Scholar]
- 85. Palma V, Ruocco C, Cortese M, et al. : Bioalcohol reforming: An overview of the recent advances for the enhancement of catalyst stability. Catalysts. 2020;10(6):665. 10.3390/catal10060665 [DOI] [Google Scholar]
- 86. Quartarone E, Angioni S, Mustarelli P: Polymer and composite membranes for proton-conducting, high-temperature fuel cells: A critical review. Materials (Basel). 2017;10(7):687. 10.3390/ma10070687 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 87. Azenha CSR, Mateos-Pedrero C, Queirós S, et al. : Innovative ZrO 2-supported CuPd catalysts for the selective production of hydrogen from methanol steam reforming. Appl Catal B Environ. 2017;203:400–7. 10.1016/j.apcatb.2016.10.041 [DOI] [Google Scholar]
- 88. O’Hayre R, Cha SW, Colella W, et al. : Fuel Cell Fundamentals. Fuel Cell Fundamentals. 2016. 10.1002/9781119191766 [DOI] [Google Scholar]
- 89. Brett DJL, Brandon NP, Hawkes AD, et al. : Fuel cell systems for small and micro combined heat and power (CHP) applications. Small Micro Comb Heat Power Syst Adv Des Performance, Mater Appl. 2011;233–61. 10.1533/9780857092755.2.233 [DOI] [Google Scholar]
- 90. Bae J: Fuel Processor Lifetime and Reliability in Solid Oxide Fuel Cells. In: Solid Oxide Fuel Cell Lifetime and Reliability: Critical Challenges in Fuel Cells.2017;145–171. 10.1016/B978-0-08-101102-7.00008-8 [DOI] [Google Scholar]
- 91. Barbir F: PEM fuel cells : theory and practice.Elsevier,2013. Reference Source [Google Scholar]
- 92. Shekhawat D, Spivey JJ, Berry DA: Fuel Cells: Technologies for Fuel Processing. Fuel Cells: Technologies for Fuel Processing. 2011. 10.1016/C2009-0-20328-X [DOI] [Google Scholar]
- 93. Araya SS, Liso V, Cui X, et al. : A review of the methanol economy: The fuel cell route. Energies. 2020;13(3):596. 10.3390/en13030596 [DOI] [Google Scholar]
- 94. Eblagon KM, Concepción PH, Silva H, et al. : Ultraselective low temperature steam reforming of methanol over PdZn/ZnO catalysts-Influence of induced support defects on catalytic performance. Appl Catal B Environ. 2014;154–155:316–328. 10.1016/j.apcatb.2014.02.032 [DOI] [Google Scholar]
- 95. Silva H, Mateos-Pedrero C, Ribeirinha P, et al. : Low-temperature methanol steam reforming kinetics over a novel CuZrDyAl catalyst. React Kinet Mech Catal. 2015;115:321–339. 10.1007/s11144-015-0846-z [DOI] [Google Scholar]
- 96. Palo DR, Dagle RA, Holladay JD: Methanol steam reforming for hydrogen production. Chem Rev. 2007;107(10):3992–4021. 10.1021/cr050198b [DOI] [PubMed] [Google Scholar]
- 97. Zhang C, Zhou W, Ehteshami MM, et al. : Determination of the optimal operating temperature range for high temperature PEM fuel cell considering its performance, CO tolerance and degradation. Energy Convers Manag. 2015;105:433–41. 10.1016/j.enconman.2015.08.011 [DOI] [Google Scholar]
- 98. Simon Araya S, Juhl Andreasen S, Venstrup Nielsen H, et al. : Investigating the effects of methanol-water vapor mixture on a PBI-based high temperature PEM fuel cell. Int J Hydrogen Energy. 2012;37(23):18231–42. 10.1016/j.ijhydene.2012.09.009 [DOI] [Google Scholar]
- 99. Rastedt M, Büsselmann J, Klicpera T, et al. : Determination of Long-Term Stability and Quality of HT-PEM MEAs. ECS Trans. 2018;86(13):301–14. 10.1149/08613.0301ecst [DOI] [Google Scholar]
- 100. Sentenac P, Bur Y, Rauzy E, et al. : Density of Methanol + Water between 250 K and 440 K and up to 40 MPa and Vapor-Liquid Equilibria from 363 K to 440 K. J Chem Eng Data. 1998;43(4):592–600. 10.1021/je970297p [DOI] [Google Scholar]
- 101. Olevskii VM, Golubev IF: The vapor-liquid equilibrium at elevated pressures: V. Tr GIAP. 1956;45–51. [Google Scholar]
- 102. Du H, Li J, Zhang J, et al. : Separation of hydrogen and nitrogen gases with porous graphene membrane. J Phys Chem C. 2011;115(47):23261–6. 10.1021/jp206258u [DOI] [Google Scholar]
- 103. Zhu J, Meng X, Zhao J, et al. : Facile hydrogen/nitrogen separation through graphene oxide membranes supported on YSZ ceramic hollow fibers. J Memb Sci. 2017;535:143–50. 10.1016/j.memsci.2017.04.032 [DOI] [Google Scholar]