Next Article in Journal
Quasi-Static Research of ATV/UTV Non-Pneumatic Tires
Next Article in Special Issue
A High-Efficiency Cooperative Control Strategy of Active and Passive Heating for a Proton Exchange Membrane Fuel Cell
Previous Article in Journal
Determining Actuator Requirements for Cyclic Varying Pitch Propeller for Ships
Previous Article in Special Issue
A Multi-Stage Fault Diagnosis Method for Proton Exchange Membrane Fuel Cell Based on Support Vector Machine with Binary Tree
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Studies on the Effects of NH3 in H2 and Air on the Performance of PEMFC

1
Henan Yuqing Power Co., Ltd., 416 Mu’ye Road, Xinxiang 453000, China
2
School of Automotive Studies, Tongji University (Jiading Campus), 4800 Cao’an Road, Shanghai 201804, China
*
Author to whom correspondence should be addressed.
Energies 2021, 14(20), 6556; https://doi.org/10.3390/en14206556
Submission received: 4 August 2021 / Revised: 25 September 2021 / Accepted: 8 October 2021 / Published: 12 October 2021
(This article belongs to the Collection Batteries, Fuel Cells and Supercapacitors Technologies)

Abstract

:
The effect of NH3 in H2 and in air was investigated at various concentrations ranging from 1.0 ppm to 100 ppm in air and ranging from 0.25 ppm to 10 ppm in fuel. The effect of NH3 on cathode caused an instantaneous decrease in cell voltage which dropped from 0.734 V to 0.712 V in 30 h and drop rates was 0.73 mV/h for 1 ppm; however, the cell voltage dropped to 0.415 V in 1 h for 100 ppm of NH3. The voltage could not be recovered after the polarization test (V-I test) but could be recovered to 84.4% after operation with neat air for 1.5 h and 98.4% after cycle voltammogram (CV). It was found that the voltage drop was obvious, and the drop rate increased with the NH3 concentration in H2. The voltage drop rates at 500 mA/cm2 were 0.54 mV/h for 0.5 ppm of NH3, 0.8 mV/h for 1 ppm, and 2 mV/h for 10 ppm. The voltage could be recovered from 70.6% to 77.3% after discharged with high purity H2 for 24 h, to 92.8% after being purged with clean air for 10 h and to 98.4% after CV scan. The tolerance concentration of NH3 in H2 for 1000 h was 40 ppb, for 2000 h was 20 ppb, and for 5000 h was 9 ppb.

1. Introduction

Proton exchange membrane fuel cell (PEMFC) is one of the most suitable candidates for replacing internal combustion engine (ICE) due to its high power density, zero emissions, and low operation temperature [1]. However, if the fuel of fuel cell vehicle (FCV), hydrogen, is produced from fossil fuel or by-products, it will inevitably contain some impurities, such as N2 [2], CO [2,3,4,5], H2S [5,6,7,8,9], CO2 [5,7,8,9,10], SO2 [11], and/or NH3 [5,12,13,14,15,16], etc. These impurities have potential negative influences on the performance of PEMFC.
Among the impurities, ammonia (NH3) is a common contaminant not only in the hydrogen-rich fuel stream [12] but also in ambient air in some special places, such as heavy traffic tunnels and the livestock industry. At least four different research groups have published papers regarding the ammonia contamination of PEMFC [13,14,15,16]. Uribe, F.A et al. [13], Soto, H.J et al. [14], Gomez, Y.A et al. [15], and Halseid, R et al. [16] focused on studying the poisoning of cells in operation and recovery after exposure to ammonia in H2. The poisoning mechanism was only partly identified.
Uribe, F.A et al. [13] found that the cell resistance measured by high frequency resistance (HFR) was more than doubled when the cell was exposed to 30 ppm NH3 in H2 for 15 h. Exposure to 30 ppm ammonia for about 1 h resulted in performance loss, which was recoverable in about 18 h. Extended exposure to 30 ppm NH3 for 17 h was not fully recoverable within 4 days of operation on pure H2. Through a cyclic voltammetry (CV) test, no electrochemically active adsorbed contaminant could be identified on either the anode or the cathode. It was suggested that the observed performance loss was due to a loss of proton conductivity in the anode catalyst layer.
In the work of Soto, H.J et al. [14], the cell resistance was monitored by an automated current interrupt technique. Exposure to 200 ppm NH3 in H2 for 10 h increased the measured cell resistance by about 35%, much less than that observed by Uribe, who used a lower concentration of NH3. CVs recorded on the anode did not reveal any electrochemically active species. Soto, similarly to Uribe, suggested that ammonium interfered primarily with the anode. Gomez, Y.A et al. [15] also found that 200 ppm NH3 in H2 rapidly decreases the voltage of the fuel cell. The cell voltage at a constant current density of 0.1 A/cm2 decreases from 0.83 V to 0.54 V during a 100 min period, resulting in a total voltage decay of 290 mV.
Halseid, R et al. [16] found that the poisoning of PEMFCs by NH3 would take 24 h or more to reach a steady state with 10 ppm NH3 in H2. In some cases, no steady state was reached during the experiment. The performance loss was in most cases reversible, but only after operation on high-purity H2 for several days. The introduction of 1 ppm NH3 resulted in significant performance loss. The performance loss was higher than could be explained by the observed increase in ohmic resistance in the cell. There is also a significant effect of ammonium on the ORR on the cathode.
In the present work, five different concentrations of NH3 in air were added to the cathode of a cell, and four different concentrations of NH3 in H2 were added to the anode. Effects of NH3 on both the anode and the cathode were studied over a long period (40 h on cathode and more than 100 h on anode) of durability tests. Uribe, F.A et al. [13], Soto, H.J et al. [14], and Gomez, Y.A et al. [15] studied the effect of relatively high contamination levels (200 ppm, 13–130 ppm, and 200–1000 ppm, respectively). Halseid, R et al. [16] studied FCs contaminated with lower levels of NH3 in the H2 (1–20 ppm). In this work, we studied even lower concentrations of NH3 in H2, such as 0.25–10 ppm.
The change in cell impedance was investigated by electrochemical impedance spectroscopy (EIS) tests. To reveal change in electrochemically active species of a cell poisoning with NH3, CV tests were carried out. Furthermore, a tolerant limit of NH3 in H2 for fuel cells (FCs) was suggested from experimental data of the durability tests.

2. Experimental

2.1. Materials and Equipments

All single fuel cells used in experiments were fabricated from MEAs (membrane electrode assemblies, Shanghai Hesen Electric, Shanghai, China) with a size of 50 cm2. The Pt loading was 0.3 mg/cm2 on each side of the membrane (DuPont, Wilmington, DE, USA). For each contamination experiment, a new MEA was used.
Highly pure H2 (>99.999%), pure N2 (>99.99%), and compressed air were used. A low concentration of NH3 in H2 and in air was prepared by mixing NH3 in H2 or in N2 in a gas blender to a level of 20 ppm NH3 in H2 and 200 ppm NH3 in N2, and then dynamically diluted to a desired concentration.
A fuel cell test bench equipped with an electrochemical station (VMP2/Z, Princeton Applied Research) was developed, which can introduce impurities simultaneously and dynamically into H2 stream or air (See Figure 1). An electronic load (SUN-FEL200A, SUNRISE POWER) was used for discharging in galvanostatic mode.

2.2. Experiments

2.2.1. Durability (V-t) Test

Durability tests before and after the introduction of NH3 into the anode and cathode on single cells were carried out to characterize the changes in performance. The performance loss was determined by measuring the voltage drop rate (mV/h) when the cell was discharged at 500 mA/cm2. Unless otherwise specified, the cells were operated at 70 °C and 101.3 kPa and the anode stream was humidified at 75 °C. The stoichiometric ratio for H2 and air were fixed at 1.5 and 3.0, respectively.

2.2.2. Polarization (V-I) Test

The polarization test is another method used to characterize the change in cell performance. When the V-I test was carried out, the stoichiometries ratio for H2 and air were kept unchanged. A reference V-I curve was obtained by a measurement with pure H2 right after the MEA was activated for 30–40 h. Usually, V-I tests were carried out during the V-t-tests.

2.2.3. EIS Test

EIS tests were carried out by a multi-channel potentiostat (VMP2/Z, PAR) in the galvanostatic mode to investigate the impedance change after the introduction of NH3. In EIS experiments, the amplitude of the sinusoid disturbance was set at 200 mA, which was 4% of the value of the discharge current, 5 A, i.d. 100 mA/cm2. Impedances were measured with 6 steps per decade of frequencies between 10 kHz and 100 MHz. When the EIS tests were carried out to investigate the effect of NH3 in air, the polarization of the anode could be ignored because of the invariability of pure H2 stream; therefore, all the changes in EIS could be attributed to the changes at the cathode caused by the poisoning of NH3, and vice versa for the anode.

2.2.4. CV Test

Cyclic voltammetry is a sensitive technique for investigating the behaviors of impurities on both the anode and cathode. While CV was carried out to investigate the NH3 effect on the anode, N2 at a rate of 300 mL/min was introduced into the anode, which acted as a working electrode. Meanwhile, H2 at a rate of 300 mL min−1 was introduced into the cathode, which acted as a counter electrode and reference electrode. The anode was scanned by the electrochemical station from 0.05 V to 1.4 V vs. RHE and the scan rate was set at 20 mV sec−1 (and vice versa when the cathode was used as a working electrode).

3. Results and Discussion

3.1. Effect of NH3 on the Cathode

A typical cell response to exposure to air containing five different concentrations of NH3 ranging from 1.0 to 100 ppm is shown in Figure 2. The cell was operated galvanostatically at 500 mA/cm2 and allowed to stabilize before NH3 was added to air for 30–40 h. Each concentration of ammonia introduction was continued until the cell voltage dropped to a stable value.
It can be noticed that the cell performance decreased instantaneously when NH3 was added to the cathode, and the cell voltage drop rate increased with the concentration of NH3. It can be also found that the cell voltage decreases operating with neat air as the running time increases because of the unstable and unsuitable operation conditions, unbalanced thermos-water management, slowly but certainly proceeding material aging process. However, in the V-t curve of 1.0 ppm of NH3, the cell voltage dropped from 0.734 V to 0.712 V in 30 h, and in that of 100 ppm of NH3, the cell voltage dropped to 0.415 V in 1 h.
Figure 3 shows that the cell voltage at the constant current changes with different tests. It can be seen that the cell voltage rapidly decreases after adding the NH3. After turning off the NH3 addition, we carried out the following tests: V-I test, operating with neat air, and CV scan. The cell voltage had little change after the V-I test but recovered to 0.620 V after operating with neat air for 1.5 h. However, through CV scanning, the cell voltage could be recovered to 0.720 V, with a recovery rate of 98.2%.
When the cell voltage dropped to a stable state after poisoning with NH3, the addition of NH3 was turned off, and then a V-I test was carried out. The V-I test results were shown in Figure 4. It can be seen that the polarization loss is closely linked with the NH3 concentration, and the influence is much more severe when the concentration of NH3 is above 5.0 ppm. Moreover, the cell could no longer work efficiently at the current density of 900 mA/cm2, if the NH3 concentration was 5.0 ppm. Similarly, the cell would lose its work ability at 700 mA/cm2 under the effect of 20 ppm of NH3, and at 600 mA/cm2 under the effect of 100 ppm of NH3.

3.2. Effect of NH3 on the Anode

In Figure 5, the results of contamination of a cell with four different concentrations of NH3 in fuel are shown. The cell was discharged at 500 mA/cm2 and the addition of NH3 was continued for more than 100 h.
It can be seen that, in each curve, the voltage drop rate remained approximately constant with the addition of NH3. For the same reason as operating with neat air, the voltage with pure H2 also decreases. However, the voltage would fluctuate occasionally, which could be attributed to the adsorption of NH3 in the membrane. When NH3 was adsorbed in the membrane, reaction between NH3 and H+ took place (R1), and ammonium would be formed, which would reduce the acidity and proton conductivity of the membrane [17]. In that case, the transfer of proton would be blocked. Simultaneously, NH3 would be swept by H2 and H2O stream intermittently. The adsorption and desorption of NH3 cost the fluctuation of cell voltage, which is chemisorption.
NH 3 + H +   NH 4 +
The voltage drop rate increased with the NH3 concentration. For example, in the curve of 0.25 ppm of NH3, the voltage drop from 0.697 V to 0.67 V in 100 h, the voltage drop rate is 0.27 mV/h. In the curve of other concentrations, the voltage drop rates are 0.54 mV/h for 0.5 ppm of NH3, 0.8 mV/h for 1.0 ppm of NH3, and 2.0 mV/h for 10 ppm of NH3.
Serial tests were performed to study the recovery of cell performance after exposed to 10 ppm NH3/H2 for more than 100 h. As seen in Figure 6, the cell was recovered through (a) cutting off NH3 and carrying out the polarization curve test and EIS test, (b) running with pure H2, (c) purging the anode with clean air, (d) and testing with CV. It was found that after the NH3 poisoning, the cell voltage dropped to ca. 0.492 V, and even discharged with high-purity H2 for as long as 24 h, the cell voltage was lightly recovered. It merely rosed from 0.492 V to 0.539 V, i.e., 77.3% of the original value. After being purged with clean air for 10 h, the cell voltage was further recovered to 0.647 V, i.e., 92.8% of the original value. After CV scan, the cell voltage could be recovered to 0.686 V, i.e., 98.4% of the original value. During the composite processes treatment, the anode surface was cleaned with moist H2 at first; meanwhile, some NH4+ cations could be washed out. When clean air was introduced, with CV scan in higher potential, the substance adsorbed on the surface was removed. Therefore, cell performance was mostly, if not completely, recovered.
Figure 7 shows the polarization curves after poisoning with different concentrations of NH3/H2. Each polarization test was carried out 100 h after the addition of NH3. It seems that the polarization losses did not strengthen too much in the curves of 0.25 ppm and 0.5 ppm of NH3. However, it was much more severe when 1 ppm or higher concentrations of NH3 were introduced into the anode. Compared to the effect of NH3 in air, it was found that the polarization losses were much more severe with the same concentration levels in H2.

3.3. Recovery Mechanism Study

3.3.1. Impedance Study

In order to clarify the effect of NH3 in air and H2 on the impedance of PEMFC, EIS tests were carried out to study the impedance changes during the durability tests. Nyquist plots drawn in EIS tests show the same pattern as shown in Figure 8a,b: an inductive line at a high frequency range and a depressed capacitive semicircle at intermediate and low frequency ranges [18,19]. The inductive lines intersecting with the real axis correspond to the total ohmic resistance of the cell. This single semicircle loop, often called the “kinetic loop”, occurs when the electrode process is dominated only by the interfacial kinetics of the ORR process. The radii of the semicircle show the polarization resistance of the cell. The electrochemical parameters (Table 1) of the MEAs were analyzed via an equivalent circuit (EC) method (Figure 8c). Rct represents the charge transfer resistance of the ORR; Rs represents the total ohmic resistance of the cell, which can be expressed as the sum of the contributions from contact resistances between components and ohmic resistances of the cell components; Rm represents the mass transfer impedance.
In Figure 8a, three Nyquist plots of the cathode under different states are shown: in clean air, after being poisoned by 100 ppm of NH3, and after being discharged with clean air and CV scanned.
Under the effect of NH3, the Rs increases; however, the capacitive semicircle varies slightly. This means the impedance change is mainly attributable to the increase of ohmic resistance, which is due to the NH4+ absorption on the surface of catalyst and its transfer into the membrane. After the addition of NH3 was shut down, the cell was discharged with neat air for 3 h, and the cathode was scanned by CV; the impedance spectra were greatly recovered. The immovability in polarization resistance showed that the Pt/C catalyst was not affected by NH3. However, in the catalysis of Pt/C, NH3 could be oxidized to N2 or nitrogen oxides, and then be swept out of the cathode by air flow and water. The Reactions are proposed as follows:
4 NH 3 +   3 O 2     2 N 2   +   6 H 2 O
2 NH 3 + O 2   2 N O   +   6 H +   +   4 e
NH 3 + O 2     N O 2   +   3 H +   +   2 e
2 NH 3 + 3 O 2     2 N O 3   +   6 H +   +   4 e
Figure 8b shows the impedance test results of cell poisoning with 10 ppm NH3 in H2 recovered by certain methods, such as purging with pure H2 and CV scanning.
It can be seen that both the ohmic and polarization resistance increased with the addition of NH3. After the methods mentioned above were taken to recover the cell, the polarization resistance decreased but could not return to the original level, as well as the ohmic resistance. This means that some of the NH3 added to the anode was absorbed in the membrane. Ammonium formed by reaction (R1) could not be easily swept out of the membrane, which caused the increase in ohmic resistance. Other NH3 might be adsorbed by Pt/C catalyst [20], which caused the increase in polarization resistance. Under the oxidation circumstance of the anode and the absence of oxygen, the adsorbed NH3 molecule can be oxidized by losing one or two proton(s):
Pt + NH 3     Pt NH 3
Pt + NH 3     Pt NH 2   +   H +
Pt + NH 3     Pt NH 2   +   2 H +

3.3.2. Cyclic Voltammetry Study

CV was employed to investigate the adsorption and desorption of NH3 on the surface of Pt, because it is a sensitive tool for the diagnostic of the surface state change, including the determination of the electrochemical active surface (EAS) of an electrode [21,22].
In Figure 9a combined with the Table 2, the CV results of a cell poisoning with 100 ppm NH3 in air for 2 h (curve b) and then discharged with neat air for 3 h (curve c) are shown, as well as the one tested with pure H2 (reference curve a). With the contamination of 100 ppm NH3, the CV curve changed significantly. The broader oxidation curve in ca. 0.8 V can be attributed to the oxidation of adsorbed ammonia molecules. On the other hand, the double layer region became smaller after contamination, changing from 0.28 V~0.68 V to 0.35 V~0.68 V, because of the weakened capacity of the catalyst layers due to the adsorption of ammonia. However, the H-desorption peaks in the range of 0.1 to 0.3 V showed little variation. After air purging for 3 h, curve b was close but not similar to the original one (curve a), which clarified the irreversibility of the cell performance after the cell was poisoned by NH3 in the cathode. The reason to explain why the CV curve changes so much needs to be found through deeper research.
The CV results poisoning with 10 ppm NH3 in H2 for 100 h (curve b) and then discharging with pure H2 for 24 h and purging with air for 10 h (curve c) are shown in Figure 9b, as well as the one tested with pure H2 (reference curve a). It can be seen in curve b that the H-desorption peaks in the range of 0.1 to 0.3 V decrease after the introduction of NH3, and due to the strong adsorption of NH3 on Pt, the EAS of Pt is blocked by NH3, which in turn suppresses the HOR, leading to a decrease in cell performance. Furthermore, there exists a reduced oxidation peak of Pt at 0.85 V in curve b. This peak could be caused by the reduction of the EAS of Pt. Indeed, the reduction peak for PtO or PtO2 at 0.7 V during the cathodic scan became weaker, because less Pt oxides were generated during the previous anodic scan. After being discharged with pure H2 and purged with neat air, curve c recovered almost completely, but the H-desorption peaks in the range of 0.1 to 0.3 V became smaller than that of the reference one (curve a), implying that the surface properties (EAS, particle size, etc.) of Pt/C might be changed. Therefore, the poisoning effect of NH3 cannot be completely recovered by CV oxidation.

3.4. Determination of Tolerant Limit of NH3 in H2

Compared with air, the lower concentration of NH3 in H2 has a more severe performance degradation. Therefore, it is important to set a tolerance limit of NH3 in H2 for automotive applications, under which the fuel cell should not be impacted to lose too much work efficiency in a certain period and the cell should be able to perform under rather low degradation rates. In automotive application, the ideal lifetime should be 5000 h or more [23]. However, nowadays, the technologies for the lifespan of FC do not reach that high a level. If the level of 5000 h is set, the tolerance limit would be set extremely low, and the cost of hydrogen would be impractically high. Therefore, it could be considered that the tolerance limit of NH3 can be set in three phases. The running time of FC must reach 1000 h in the first phase, 2000 h in the second phase, and 5000 h in the third phase. Additionally, the limitation of the decrease in cell voltage at 500 mA/cm2 was set at 25 mV.
In Figure 5, the original voltage and the dropped voltage poisoning with NH3 were indicated by two straight lines. The line at 0.697 V indicated the original voltage and the one at 0.672 V as the voltage drop tolerance limit. Based on the experimental results given in Figure 5, the time spent for 25 mV of cell voltage decrease, and for the four corresponding concentrations of NH3, the time spent to reach 25 mV voltage decrease under the effect of certain concentration of NH3 can be obtained. In Figure 10, CNH3, the concentration of NH3, is plotted against 1/t. A polynomial relationship between the two variables can be drawn out:
C N H 3 = 44888 ( 1 t ) 3 2170.3 ( 1 t ) 2 + 45.251 ( 1 t )
where t is time of fuel cell running, h, and CNH3 is NH3 concentration, ppm.
From Equation (1), the tolerance limit for a 25 mV voltage drop can be calculated. For example, to reach a running time of 1000 h, the concentration of NH3 in H2 should be controlled at a level of 40 ppb. For 2000 h of operation time, the concentration of NH3 should be decreased to 20 ppb. For 5000 h of operation time, the concentration of NH3 should be controlled at 9 ppb, which is very close to the level set by the national standard [24].
In this work, the determination of the tolerance limit was discussed based on the durability test results. However, the influence factors on the behavior of NH3 effect may be multiple. Therefore, more experiments should be carried out in various operation conditions, such as Pt loading, the operating current density, or voltage and cell temperature, etc. Nevertheless, the method in this paper can be applied in the same way; furthermore, the effects of other hydrogen impurities, such as CO and H2S, etc., can also be investigated as well.

4. Conclusions

The present work investigated the poisoning effect of a low concentration of NH3 in air and in fuel on PEMFC. From the experimental results, the following conclusions can be drawn:
(1)
The effect of NH3 on the cathode in the concentration range from 1.0 to 100 ppm was investigated. The cell voltage would drop to a stable value when the NH3 was added to the cathode, and the value decreased with NH3 concentration. Through discharging with neat air and then CV, the cell voltage could be recovered to 98.2% of the original level. NH3 addition in air caused the increase in polarization resistance, which could be recovered through discharging with neat air. However, the CV curve changed significantly under the effect of NH3, which could not be recovered by neat air purging.
(2)
The effect of NH3 on the anode in the concentration range from 0.25 to 10 ppm was investigated. The cell voltage drop rate remained approximately constant with the addition of NH3; however, the voltage would fluctuate occasionally. The cell voltage could be recovered to 98.4% of the original level through discharging with pure H2, neat air purging, and CV scan. NH3 in fuel would cause the ohmic and polarization resistance to increase, which was irreversible when the methods mentioned above were taken to recover the cell. The irreversibility was also shown in the CV curves. The H-adsorption peaks in the range of 0.1–0.3 V after poisoning with NH3 could not be recovered to the original one.
(3)
The tolerance limits of NH3 concentration in H2 were determined and a polynomial relationship between the concentration of NH3 for a cell voltage drop of 25 mV and the corresponding time of operation was established. The allowable concentration of NH3 in H2 for a fuel cell operation time of 1000 h is 40 ppb, while for 2000 h, the limit should be lowered to 20 ppb. For 5000 h, the limit should be controlled at 9 ppb.

Author Contributions

Conceptualization, K.H. and D.Y.; Methodology, K.H. and D.Y.; Software, K.H.; Validation, D.Y.; Formal analysis, K.H. and D.Y.; Investigation, K.H. and D.Y; Resources, D.Y.; date curation, K.H.; writing-original draft, K.H.; writing-review and editing, D.Y. All authors have read and agreed to the published version of the manuscript.

Funding

This research was funding by the Major Science and Technology Projects of Xinxiang City (ZD18008).

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

Data sharing not applicable.

Acknowledgments

The author wishes to thank Henan Yuqing Power Co., Ltd. for the support. The present work is partly supported by the Major Science and Technology Projects of Xinxiang City (ZD18008).

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. EG&G Technical Services, Inc. Fuel Cell Handbook, 7th ed.; US Department of Energy: West Virginia, WV, USA, 2004. [Google Scholar]
  2. Devrim, Y.; Albostan, A.; Devrim, H. Experimental investigation of CO tolerance in high temperature PEM fuel cells. Int. J. Hydrogen Energy 2018, 43, 18672–18681. [Google Scholar] [CrossRef]
  3. De Bruijn, F.A.; Papageorgopoulos, D.C.; Sitters, E.F.; Janssen, G.J.M. The influence of carbon dioxide on PEM fuel cell anodes. J. Power Sources 2002, 110, 117–124. [Google Scholar] [CrossRef]
  4. Janssen, G.J.M. Modelling study of CO2 poisoning on PEMFC anodes. J. Power Sources 2004, 136, 45–54. [Google Scholar] [CrossRef]
  5. Zhao, Y.Q.; Mao, Y.J.; Zhang, W.X.; Tang, Y.L.; Wang, P.Z. Reviews on the effects of contaminations and research methodologies for PEMFC. Int. J. Hydrogen Energy 2020, 45, 23174–23200. [Google Scholar] [CrossRef]
  6. Cremers, C.; Rau, M.S.; Niedergesass, A.; Pinkwart, K.; Tuebke, J. Effect of Model Fuel Impurities for Reformed Jet Fuels on the Hydrogen Oxidation at Platinum Based Catalysts under HT-PEMFC Conditions. ECS Trans. 2016, 75, 919–929. [Google Scholar] [CrossRef]
  7. Araya, S.S.; Andreasen, S.J.; Kaer, S.K. Experimental Characterization of the Poisoning Effects of Methanol-Based Reformate Impurities on a PBI-Based High Temperature PEM Fuel Cell. Energies 2012, 5, 4251–4267. [Google Scholar] [CrossRef]
  8. De Wild, P.J.; Nyqvist, R.G.; De Bruijn, F.A.; Stobbe, E.R. Removal of sulphur-containing odorants from fuel gases for fuel cell-based combined heat and power applications. J. Power Sources 2006, 159, 995–1004. [Google Scholar] [CrossRef]
  9. Mohtadi, R.; Lee, W.K.; Van Zee, J.W. The effect of temperature on the adsorption rate of hydrogen sulfide on Pt anodes in a PEMFC. Appl. Catal. B 2005, 56, 37–42. [Google Scholar] [CrossRef]
  10. Jimenez, S.; Soler, J.; Valenzuela, R.X.; Daza, L. Assessment of the performance of a PEMFC in the presence of CO. J. Power Sources 2005, 151, 69–73. [Google Scholar] [CrossRef]
  11. Prithi, J.A.; Rajalakshmi, N.; Dhathathereyan, K.S. Mesoporous platinum as sulfur-tolerant catalyst for PEMFC cathodes. J. Solid State Electrochem. 2017, 21, 3479–3485. [Google Scholar] [CrossRef]
  12. Chellappa, A.S.; Fischer, C.M.; Thomson, W.J. Ammonia decomposition kinetics over Ni-Pt/Al2O3 for PEM fuel cell applications. Appl. Catal. A-Gen. 2002, 227, 231–240. [Google Scholar] [CrossRef]
  13. Uribe, F.A.; Gottesfeld, S.; Zawodzinski, T.A. Effect of ammonia as potential fuel impurity on proton exchange membrane fuel cell performance. J. Electrochem. Soc. 2002, 149, A293–A296. [Google Scholar] [CrossRef]
  14. Soto, H.J.; Lee, W.K.; Van Zee, J.W.; Murthy, M. Effect of transient ammonia concentrations on PEMFC performance. Electrochem. Solid ST 2003, 6, A133–A135. [Google Scholar] [CrossRef]
  15. Gomez, Y.A.; Oyarce, A.; Lindbergh, G.; Lagergren, C. Ammonia Contamination of a Proton Exchange Membrane Fuel Cell. J. Electrochem. Soc. 2018, 165, F189–F197. [Google Scholar] [CrossRef]
  16. Halseid, R.; Viel, P.J.S.; Tunold, R. Effect of ammonia on the performance of polymer electrolyte membrane fuel cells. J. Power Sources 2006, 154, 343–350. [Google Scholar] [CrossRef]
  17. Hongririkarn, K. Effect of Impurities on Performance of Proton Exchange Membrane Fuel Cell Componets; Celmson University: Clemson, CA, USA, 2010; pp. 14–20. [Google Scholar]
  18. Rodat, S.; Sailler, S.; Druart, F.; Thivel, P.X.; Bultel, Y.; Ozil, P. EIS measurements in the diagnosis of the environment within a PEMFC stack. J. Appl. Electrochem. 2010, 40, 911–920. [Google Scholar] [CrossRef]
  19. Tant, S.; Rosini, S.; Thivel, P.X.; Druart, F.; Rakotondrainibe, A.; Geneston, T.; Bultel, Y. An algorithm for diagnosis of proton exchange membrane fuel cells by electrochemical impedance spectroscopy. Electrochim. Acta 2014, 135, 368–379. [Google Scholar] [CrossRef]
  20. Ford, D.C.; Xu, Y.; Mavrikakis, M. Atomic and molecular adsorption on Pt(111). Surf. Sci. 2005, 587, 159–174. [Google Scholar] [CrossRef]
  21. Marken, F.; Neudeck, A.; Bood, A.M. Electroanalytical Methods; Scholz, F., Ed.; Springer: Berlin/Heidelberg, Germany, 2010; Chapter II.1; pp. 57–106. [Google Scholar]
  22. Zhang, X.Y.; Pasaogullari, U.; Molter, T. Influence of ammonia on membrane-electrode assemblies in polymer electrolyte fuel cells. Int. J. Hydrogen Energy 2009, 34, 9188–9194. [Google Scholar] [CrossRef]
  23. Gasteiger, H.A.; Kocha, S.S.; Sompalli, B.; Wagner, F.T. Activity benchmarks and requirements for Pt, Pt-alloy, and non-Pt oxygen reduction catalysts for PEMFCs. Appl. Catal. B 2005, 56, 9–35. [Google Scholar] [CrossRef]
  24. Chinese Standard GB/T. Fuel Specification for Proton Exchange Membrane Fuel Cell Vehicles-Hydrogen; GB/T 37244-2018; Chinese Standard GB/T: Beijing, China, 2018. [Google Scholar]
Figure 1. Schematic of PEM fuel cell test bench.
Figure 1. Schematic of PEM fuel cell test bench.
Energies 14 06556 g001
Figure 2. Performance changes of PEMFC after poisoning with NH3 in air at 500 mA cm−2.
Figure 2. Performance changes of PEMFC after poisoning with NH3 in air at 500 mA cm−2.
Energies 14 06556 g002
Figure 3. Recovery of cell voltage with discharging with neat air, V-I test, and CV at 500 mA/cm2.
Figure 3. Recovery of cell voltage with discharging with neat air, V-I test, and CV at 500 mA/cm2.
Energies 14 06556 g003
Figure 4. Effects of NH3 in air on polarization.
Figure 4. Effects of NH3 in air on polarization.
Energies 14 06556 g004
Figure 5. Performance changes of PEMFC after poisoning with NH3 in fuel.
Figure 5. Performance changes of PEMFC after poisoning with NH3 in fuel.
Energies 14 06556 g005
Figure 6. Recovery of cell voltage with discharging with pure H2, air purging, and CV.
Figure 6. Recovery of cell voltage with discharging with pure H2, air purging, and CV.
Energies 14 06556 g006
Figure 7. Effects of NH3 in H2 on polarization.
Figure 7. Effects of NH3 in H2 on polarization.
Energies 14 06556 g007
Figure 8. Nyquist plots of the impedance spectra measured before and after a cell poisoning with NH3 in air (a) and fuel (b); (c) the equivalent circuit.
Figure 8. Nyquist plots of the impedance spectra measured before and after a cell poisoning with NH3 in air (a) and fuel (b); (c) the equivalent circuit.
Energies 14 06556 g008
Figure 9. CV curves before and after poisoning with NH3 in air (a) and in fuel (b).
Figure 9. CV curves before and after poisoning with NH3 in air (a) and in fuel (b).
Energies 14 06556 g009
Figure 10. Relationship between the concentration of NH3 and the maintained time allowed for a cell voltage drop of 25 mV.
Figure 10. Relationship between the concentration of NH3 and the maintained time allowed for a cell voltage drop of 25 mV.
Energies 14 06556 g010
Table 1. The electrochemical parameters of the MEAs under different conditions.
Table 1. The electrochemical parameters of the MEAs under different conditions.
Rs (ohm)Rct (ohm)Rm (ohm)
Neat air0.000501020.00103460.00033082
Poisoning with 100 ppm NH30.00053230.00114210.0003605
After neat air purging for 3 h0.00048840.000991230.00028034
Pure H20.000558080.000426530.00043754
Poisoning with 10 ppm NH3 in H20.000559640.000454340.00051119
After operating with pure H2 for 24 h0.000558350.000438460.00046464
Table 2. The ECSAs before and after poisoning with NH3 in air and fuel.
Table 2. The ECSAs before and after poisoning with NH3 in air and fuel.
ECSA   ( m 2 g P t 1 )
Neat air53.95
Poisoning with 100 ppm NH350.775
After neat air purging for 3 h59.395
Pure H279.97
Poisoning with 10 ppm NH3 in H261.175
After operating with pure H2 for 24 h67.415
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Hu, K.; Yang, D. Studies on the Effects of NH3 in H2 and Air on the Performance of PEMFC. Energies 2021, 14, 6556. https://doi.org/10.3390/en14206556

AMA Style

Hu K, Yang D. Studies on the Effects of NH3 in H2 and Air on the Performance of PEMFC. Energies. 2021; 14(20):6556. https://doi.org/10.3390/en14206556

Chicago/Turabian Style

Hu, Kefeng, and Daijun Yang. 2021. "Studies on the Effects of NH3 in H2 and Air on the Performance of PEMFC" Energies 14, no. 20: 6556. https://doi.org/10.3390/en14206556

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Back to TopTop