Next Article in Journal
Special Issue “Biocatalysts: Design and Application”
Next Article in Special Issue
The Effect of ZrO2 as Different Components of Ni-Based Catalysts for CO2 Reforming of Methane and Combined Steam and CO2 Reforming of Methane on Catalytic Performance with Coke Formation
Previous Article in Journal
Non-Porous Sulfonic Acid Catalysts Derived from Vacuum Residue Asphaltenes for Glycerol Valorization via Ketalization with Acetone
Previous Article in Special Issue
The Effect of Hydrogen Production Rate of the via Different Preparation of Co-Based Catalyst with Sodium Borohydride
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Experimental Study on Sulfur Deactivation and Regeneration of Ni-Based Catalyst in Dry Reforming of Biogas

1
Department of Mechanical Engineering, National Chung Hsing University, Taichung 40227, Taiwan
2
Thin Film Engineering Department, Taiwan Semiconductor Manufacturing Co., Tainan 74144, Taiwan
3
Department of Aeronautics and Astronautics, National Cheng Kung University, Tainan 70101, Taiwan
4
Research Center for Smart Sustainable Circular Economy, Tunghai University, Taichung 40704, Taiwan
5
Department of Mechanical Engineering, National Chin-Yi University of Technology, Taichung 41170, Taiwan
*
Authors to whom correspondence should be addressed.
Catalysts 2021, 11(7), 777; https://doi.org/10.3390/catal11070777
Submission received: 15 May 2021 / Revised: 17 June 2021 / Accepted: 23 June 2021 / Published: 26 June 2021
(This article belongs to the Special Issue Catalytic Reforming for Syngas and H2 Productions)

Abstract

:
The dry reforming of methane (DRM) using biogas and a Ni-based catalyst for syngas production was studied experimentally in this study under the presence of H2S. Using the nonpoisoned DRM performance as a comparison basis, it was found that the catalyst deactivation by the sulfur chemisorption onto the catalyst surface depends on both reaction temperature and time. With low reaction temperatures, a complete sulfur coverage was resulted and could not be regenerated. With higher reaction temperatures, the H2S coverage decreased, and the poisoned catalysts could be regenerated. The experimental results also indicated that a catalyst deactivation could not be avoided by using the bi-reforming of methane by adding O2 or H2O simultaneously in the reactant due to the stronger chemisorption capability of sulfur. The catalyst could only be regenerated after it was poisoned. The experimental results indicated that the high-temperature oxidation process was the most effective process for regenerating the poisoned catalyst.

1. Introduction

Energy shortage and environmental pollution are issues related to the sustainable development of human beings. With the continuous consumption of fossil energy sources, renewable energy has attracted wide attention for development and utilization. Moreover, the development of environmentally friendly energy utilization technologies for the traditional energy conversion processes also receives much attention. Biogas is one of the renewable biomass energies. The utilization of biogas is recognized as true carbon neutral. Due to the negative costs regarding the reduction of disposable wastes, the utilization of biogas as an energy source has both financial and environmental benefits [1].
Biogas is mainly composed of CH4 and CO2. Depending on the feedstock, CH4 amount varies from 45 to 75%, while CO2 accounts for 20~55% [2]. In addition to CH4 and CO2, biogas also contains a small amount of oxygen, nitrogen, sulfide, and ammonia [3]. At present, biogas is mainly used as fuel for power generation [4]. The CH4 is used as the energy source, while the CO2 lowers the heating value of biogas. Biogas-based power generation is a well-developed and commercialized technology [5].
Alternatively, biogas can also be used for synthetic gas (syngas) production in which the CH4 and CO2 are reformed into H2 and CO [6,7,8]. The traditional CH4-based syngas production technologies are steam reforms of methane (SRM) [9], partial oxidation of methane (POM), and dry reforming of methane (DRM) [10]. The biogas can be used as the feedstock for these technologies [6,7,8]. In SRM or POM, a steam or oxygen addition is required. In DRM, CH4 is reformed with CO2. The CO2 contained in biogas can be used directly in DRM. As compared with SRM and POM, the syngas production from DRM using biogas as a feedstock has several advantages. Since CO2 is one of the main gas species in biogas, there is no need for CO2 separation and sequestration. To enhance the DRM performance, the captured CO2 can be utilized in the biogas-based DRM for syngas production. The syngas produced by DRM has the H2/CO ratio of 1, which is suitable for liquid fuel synthesis from Fischer-Tropsch synthesis [11].
Although DRM is of great significance from both a renewable energy usage and environmental protection point of view, it is not ready in practical applications due to at least two problems [12]. One of the main problems is the high energy consumption because DRM is an endothermic reaction. To maintain a high conversion rate, there must be a high energy input. Another major problem of DRM is catalyst deactivation. It has been shown that catalyst deactivation is mainly caused by high-temperature active metal sintering and carbon deposition [13]. The catalyst deactivation results in a reduction in reaction conversion rate and increased catalyst renewal cost.
In addition to the problems of energy consumption and catalyst deactivation, the effect of impurities contained in actual biogas for DRM is also a key issue for syngas production. Due to its low cost and high activity, Ni-based catalysts have been widely adopted in the CH4-based reforming reaction [14]. During biogas production, it inevitably produces H2S, and its removal is necessary for various biogas applications. Similar to coal gasification and SRM, the presence of H2S, even with a small amount, will cause the catalyst poison in DRM [15,16]. In general, the catalyst poison mechanism is that the H2S reacts with the active metal to occupy the active site of the catalyst, which, in turn, leads to catalyst deactivation [17]. In the actual reaction process, the catalyst poison is affected by many factors such as reaction temperature, gas composition, reactor parameters, H2S concentration, and type of catalyst [13,18].
The catalyst poisoned by H2S has been studied extensively in SRM [19]. As compared with SRM, studies on the H2S effect on DRM performance are relatively few. In this study, a detailed experimental study was carried out to examine the catalyst poison by H2S in DRM. To reduce the cost of DRM, the regeneration of poisoned catalysts is necessary [18]. Poisoned catalyst regeneration via conventional methods and its performance were also reported in this study.

2. Thermodynamic Background

2.1. Dry Reforming of Biogas

The overall dry reforming of biogas consists of four reversible reactions, which are [20] the dry reforming of methane (DRM):
CO2 + CH4 ↔ 2CO + 2H2, ΔH298 K = 247 kJ/mol, ΔG°= 61,770 − 67.32 T
reverse water–gas shift (RWGS) reaction [21]:
CO2 + H2 ↔ H2O + CO, ΔH298 K = 41 kJ/mol, ΔG° = −8545 + 7.84 T
methane decomposition (MD):
CH4 ↔ C + 2H2, ΔH298 K =75 kJ/mol, ΔG°= 21,960 − 26.45 T
Boudouard reaction:
2CO ↔ CO2 + C, ΔH298 K = −172 kJ/mol, ΔG°= −39,810 + 40.87 T
DRM is a highly endothermic reaction and is favored at high temperatures and low pressures. With the coexistence of CH4, CO2, H2, and CO, several side reactions are possible. Equations (3) and (4) are two main side reactions related to carbon formation during the reaction. In Equations (1)–(4), the standard free energy changes ΔG° are also shown. By setting ΔG° = 0, the upper or lower limit temperatures for reactions to occur can be obtained. From Equations (1)–(4), the lower limit temperatures for the endothermic DRM, RWGS, and MD reactions to occur are 640 °C, 820 °C, and 557 °C, respectively. The upper limit temperature for an exothermic BR reaction is 700 °C. In the temperature range of 557~700 °C, carbon will be formed from methane cracking or the Boudouard reaction.
With the presence of H2S in DRM, several chemical reactions are possible. The reaction between H2S and CO2 [16,22] is
H2S + CO2 = 0.5S2 + CO + H2O
This reaction is thermodynamically favorable at relatively high temperatures. In principle, the formation of COS is also possible [22]:
H2S + CO2 = COS + H2O

2.2. Sulfur Chemisorption on Nickel

According to the study by Rostrup-Nielsen and Christiansen [17], the loss of activity of Ni-based catalysts through sulfur compounds can be due to strong sulfur chemisorption on the nickel surface, which prevents the further adsorption of reactant molecules. The reaction of sulfur chemisorption on Ni can be expressed as
H2S + Ni ⇄ Ni – S + H2
It has been pointed out by Rostrup-Nislsen [17,23] and Bartholomew [24] that the fractional surface coverage depends on the value of partial pressure ratio of p H 2 S / p H 2 . Besides, the formation of Ni−S also is influenced by the reaction temperature, gas-phase composition, and reactor parameters. Since the sulfur chemisorption process is theoretically reversible and exothermal, surface Ni−S can be regenerated by stopping H2S feeding or by temperature enhancement.

3. Experimental

Figure 1 shows the test setup of the experiment. The reactant consists of CH4, CO2, and N2. After mixing in the mixing chamber, the reactant is sent to a tubular quartz reactor in which the catalyst is filled. The reactor’s diameter is 4 mm. The reactor is placed horizontally in a temperature-controllable furnace. After the reaction, the product is sent to the condenser at which the produced H2O in the product is condensed. The dried product is sent to the GC for analyzing the molar flow rate of each species contained in the dried product. In this study, N2 is regarded as an inert gas, and its flow rate is used as the basis for computing the flow rate of other gas species in the product. As shown in Figure 1, there are two N2 supplies: one is H2S-free N2, while the other one is H2S-contained N2. To study the H2S effect in DRM, H2S-contained N2 is used.
For the poisoned catalyst regeneration, the O2 and steam treatments are to be employed. To reduce the cost of the experiment, the O2 contained in the air is used in the O2 treatment for the catalyst regeneration. As shown in Figure 1, the air is supplied from the air tank. In the steam treatment for catalyst regeneration, an HPLC pump is used to supply the liquid water. After preheating, the water vapor is mixed with dried reactant, as stated above.
Based on our previous studies [15,25], the good catalytic ability and thermal stability of DRM can be resulted by using the Ni-Ce/Al2O3 catalysts. To focus on the effect of the H2S-poisoned DRM performance, 20 wt%Ni-5 wt%CeO2/Al2O3 was adopted in this study. Detailed preparation and characterization of the catalyst can be found in our previous studies [15,25]. More detailed characterizations for the catalysts before and after use, such as TEM, TGA, and XPS, have been reported extensively in the literature [18,26,27,28].
The DRM performance is characterized by the following performance indices:
CH 4   conversion :   X C H 4 = F C H 4 , i n F C H 4 , o u t F C H 4 , i n × 100 %
CO 2   conversion :   X C O 2 = F C O 2 , i n F C O 2 , o u t F C O 2 , i n × 100 %
H 2   yield :   Y H 2 = F H 2 , o u t F C H 4 , i n
CO   yield :   Y C O = F C O , o u t F C H 4 , i n
H 2 / CO   ratio :   H 2 / C O = Y H 2 Y C O
In these equations, Fi,in and Fi,out (i = CH4, CO2, H2, CO) are the molar flow rate of species i at the reactor inlet and outlet, respectively. It is noted that the H2 and CO yields are based on the feed CH4 molar flow rate. For the stoichiometric DRM reaction, the stoichiometric yield of H2 and CO (syngas) is 2, while the H2/CO ratio is 1.

4. Results and Discussion

4.1. Experimental Parameters

Based on the thermodynamic background described above, the reaction temperature T is chosen in the range of 700–900 °C, and the reaction pressure is fixed as 1 atm. The weight of the filled catalyst is fixed as 0.5 mg. This results in a volume of the catalyst bed of 0.033 mm3. The total volume flow rate of the reactant is fixed as 50 standard cubic centimeters per minute (sccm), and the detailed composition of the reactant is shown in Table 1. For the CH4/CO2 = 1/1 case, it corresponds to a stoichiometric DRM reaction. For CH4/CO2 = 1/2, it corresponds to the excess CO2 supply for the DRM reaction. In the CH4/CO2 = 1/0.5 case, it corresponds to the lean CO2 supply for the DRM reaction. Moreover, this also corresponds to the typical components in the biogas [8].
The poisoning effect is always correlated with the H2S concentration contained in the reactant. Based on the studies by Ashrafi et al. [26] and Yang [27], the magnitude of the H2S concentration contained in the reactant affects the speed of the stabilization of the catalyst activity, and the phenomenon of catalyst deactivation is similar for different H2S concentrations. Based on these studies, a single H2S concentration with a value of 100 ppm was chosen in this study.

4.2. DRM without H2S

Figure 2 shows the effect of reaction temperature on the DRM performance without the H2S effect for the CH4/CO2 = 1/1 case. The experimental data shown in Figure 2 were taken when the reaction temperature reached the designated value, and the data collection lasted for 18 h. For every 15 min, the data from the experiment were reported. Due to endothermic reaction, it can be seen that the DRM performance is enhanced as the reaction temperature increases. The thermally stable results of the CH4 conversion, CO2 conversion, H2 yield, and CO yield, as shown in Figure 2a–d, can be obtained when the reaction time increases for T = 800 and 900 °C. For T = 700 °C, a slight decrease in the DRM performance can be found. Since there is no H2S effect, the decrease in the DRM performance is mainly due to the catalyst deactivation from the carbon deposition. By comparing the CH4 and CO2 conversions shown in Figure 2a,b, one can see that CO2 conversion is higher than the CH4 conversion for the three reaction temperatures studied. From the theoretical background, this is apparently due to the contribution from RWGS in which CO2 reacted with H2 to form CO and H2O. In Figure 2c,d, the H2 and CO yields are shown. The variations of the H2 and CO yields are similar to those of XCH4 and XCO2. Additionally, based on the thermodynamic background, the theoretical maximum yields of H2 and CO would be 2. As shown in Figure 2c,d, the yields of H2 and CO exceed 2 slightly for the T = 900 °C case due to the uncertainty of the experimental measurements. From the theoretical background, the H2/CO ratio has a value of unity. Figure 2e shows the H2/CO ratio based on the results shown in Figure 2c,d. H2/CO is less than unified due to the consumption of H2 and the formation of CO in the RWGS reaction. As noted above, the DRM reaction is the dominant reaction at high reaction temperatures. The H2 and CO yields follow the similar variation trend of XCH4. In the following discussions, only XCH4 and XCO2 are reported.
Using the data from 15 to 18 h shown in Figure 2a,b, the averaged conversions of CH4 and CO2 can be obtained and listed in Table 2. As shown in Table 2, the experimental results of the present study agree well with those reported in the literature [29,30,31].

4.3. DRM with H2S

From the results shown in Figure 2, the DRM reaction is the dominant reaction at high reaction temperatures. The H2 and CO yields follow the similar variation trend of XCH4. In the following discussions, only XCH4 and XCO2 are reported. In Figure 3, the effect of H2S on the DRM performance is shown for three reaction temperatures. The test procedure and reactant used are the same as those for Figure 2, except that the H2S is introduced in the reactant. This is achieved by switching the H2S-free N2 flow to a H2S-contained N2 flow. It is seen that the catalyst activity is decreased as the H2S is introduced in the reactant. From Figure 3, it is also clearly seen that the catalyst poisoned by H2S depends on the reaction temperature and reaction time. For T = 700 °C, linear decreases in both XCH4 and XCO2 are obtained as the reaction time increases, as shown in Figure 3a,b, respectively. For T = 800 °C, slight decreases in XCH4 and XCO2 resulted in an earlier reaction time. As the reaction time is greater than 4 h, exponential decays in both XCH4 and XCO2 result. As the reaction time is greater than 12 h, both XCH4 and XCO2 approach steady-state values, indicating that the sulfur coverage on the catalyst surface reaches a saturated condition. For T = 900 °C, the catalyst poisoned by H2S is to a smaller extent as compared with T = 700 °C and 800 °C under the same reaction time. That is, the sulfur coverage decreases with the increased reaction temperatures. It is also expected that the sulfur coverage on the catalyst surface will reach a saturated condition when the test time is increased.
From Figure 3, it is seen that the rate of the poison of Ni catalyst depended on both the reaction temperature and reaction time. For T = 700 °C, the catalyst activity lost immediately and exhibited a constant rate. From the variation trend, it was expected that zero XCH4 and XCO2 would be reached for a longer reaction time. For T = 800 °C and 900 °C, the rate of poison was about the same for the first 4 h of the reaction time. After that, it was seen that the rate of poison dropped exponentially for the T = 800 °C case, while a constant rate of poison was found for T = 900 °C. In the study of the catalyst performance for SRM by Appari et al. [32] and Yang [27], an exponential decay of catalyst activity was also found for reaction temperatures in the 700~900 °C range for the SRM case. The results shown in Figure 3 agree with those reported by Appari et al. [32] and Yang [27] for the T = 800 °C case. The above results were related to the sulfur chemisorption ability on Ni catalysts under different reaction temperatures. As pointed out by many studies [16,33,34], it is therefore suggested that higher reaction temperature is preferred in DRM when H2S-contained biogas is used as the feedstock.
In Figure 4, the effect of the CO2 amount on the H2S-poisoned DRM performance is shown for reaction temperature T = 800 °C. For the higher CO2 amount case, i.e., CH4/CO2 = 1/2, the DRM performance follows the same trend as the CH4/CO2 = 1/1 case discussed above. From the thermodynamic analysis, carbon formulation can be enhanced by reducing the CO2 amount in the DRM [35,36]. For the CH4/CO2 = 1/0.5 case, due to more carbon formulations, it can be seen that the variation trend of XCH4 is different from the CH4/CO2 = 1/1 and 1/2 cases. It is seen that CH4 conversion drops immediately in the early reaction time and then reaches a steady-state value. As compared with the CH4/CO2 = 1/1 and 1/2 cases, the initial drop in XCH4 is due to catalyst deactivation caused by both carbon deposition and sulfur poison. It is noted that the XCH4 can be enhanced by increasing the CH4/CO2 ratio for both the non-poisoned and poisoned catalysts shown in Figure 4a.
In Figure 4b, the effect of the CH4/CO2 ratio on XCO2 during the catalyst poison is shown. It is interesting to note that, for the CH4/CO2 = 1/0.5 case, XCO2 was found to continuous linearly decrease after an immediate drop when test time increases. As mentioned above, this variation is related to catalyst deactivation due to both the carbon deposition and sulfur poison. With an increased CH4/CO2 ratio, a lower XCO2 was obtained due to the excess CO2 supply in the reaction. Based on the results shown in Figure 4b, this applies to both poisoned and nonpoisoned catalysts.
By comparing the results shown in Figure 2, Figure 3 and Figure 4, it can be realized that a quick catalyst deactivation is dominated by sulfur poison. The same conclusion was also made in the study by Appai et al. [32]. Moreover, the Ce contained in the catalyst enhances the capability of carbon deposition resistance. Therefore, the effect of carbon deposition on catalyst deactivation would be to a lesser extent as compared with the sulfur poison. Further support of this conclusion can be made in the discussion on the poisoned catalyst regeneration using high temperatures.

4.4. Bi-Reforming of Methane with H2S

Since H2S chemisorption onto the catalyst surface is a reversible process, surface-adsorbed sulfur can be removed. High-temperature, O2 oxidation, and steam treatments are conventional methods used to regenerate the sulfur-poisoned reforming catalyst. Based on these regeneration technologies, it would also be interesting to examine the catalyst activity as the reactant contains O2 or steam, in addition to H2S. With O2 or steam added to the reactant, the reaction system becomes a bi-reforming of methane [37,38]. For the O2 addition case, the partial oxidation of methane (POM) or complete oxidation of methane (COM), depending on the O2 amount added, along with DRM, may occur simultaneously during the reaction. The stoichiometric POM and COM are written as
POM: CH4 + 0.5O2 ↔ 2H2 + CO
COM: CH4 + 2O2 ↔ CO2 + 2H2O
Figure 5 shows the effect of a O2 addition on DRM with and without H2S poison at T = 800 °C. As mentioned in the experimental setup, the O2 supply is from the air, and the volume flow rate ratio of the reactant is CH4/CO2/O2 = 15/15/2 (sccm). For the nonpoisoned catalyst case, the O2 addition can enhance the XCH4, as shown in Figure 5a, but lowers the XCO2, as shown in Figure 5b, due to the CO2 production from POM or COM. For the poisoned catalyst case, the catalyst activity loss can still be found with a O2 addition in the reactant for reaction times greater than 5 h. This implies that the chemisorption of H2S onto catalyst activity sites is stronger than the catalytic reaction between CH4 and O2. The DRM performance can only be improved when the coverage of H2S reaches a saturated condition.
The effect of a H2O addition on the DRM performance with and without H2S poison at T = 800 °C is shown in Figure 6. The volumetric flow rate ratio of the reactant is CH4/CO2/H2O = 15/15/10 (sccm). With the H2O addition, the steam reforming of methane (SRM), along with DRM, may occur simultaneously during the reaction. The stoichiometric SRM is written as
SRM: CH4 + H2O ↔ 3H2 + CO
For the nonpoisoned case, it can be seen that the H2O addition can enhance the XCH4 due to SRM, as shown in Figure 6a. However, due to the H2O addition, the water–gas shift reaction is also enhanced and results in a lower XCO2 due to more CO2 production, as shown in Figure 6b. Similar to the O2 addition, the H2S poison is the dominant reaction, and the catalyst activity loss is still found for the H2O addition, as shown in Figure 6. The DRM performance can only be improved when the coverage of sulfur reaches a saturated condition.

4.5. Catalyst Regeneration

From the results shown in Figure 5 and Figure 6, it can be concluded that the simultaneous addition of O2 or H2O in the DRM cannot avoid the catalyst deactivation due to the stronger chemisorption of sulfur onto the catalyst as compared to O2 or H2O. Therefore, catalyst regeneration can only be carried out after the catalyst has been poisoned. To regenerate this sulfur-poisoned reforming catalyst, three different regeneration methods were adopted in this study, including the high-temperature sulfur-free treatment, high-temperature oxidation treatment, and high-temperature steam treatment, as listed in Table 3 [32,39]. The poisoned catalyst regeneration under the high-temperature sulfur-free treatment is shown in Figure 7 for various reaction temperatures. The poison experiment was performed for the first 8 h. After that, the H2S-contained N2 gas was switched to H2S-free N2 gas, and the measurements lasted for 10 h. As shown in Figure 7a,b, the catalyst activity recovery depends on the reaction temperature. For the T = 700 °C case, the catalyst activity cannot be regenerated. As pointed out by many studies [26,40], the regenerating reaction is irreversible when the reaction is below 700 °C. For the T = 800 °C and 900 °C cases, the catalyst activity can be regenerated. With the increasing reaction time, it can be expected that the catalyst activity can approach the H2S-free performance, as listed in Table 2. The results shown in Figure 7 demonstrate that sulfur poison is the dominant factor causing catalyst deactivation, as mentioned above. In the study by Izquierdo et al. [18], the tri-reforming of biogas using a Ni-based catalyst was studied. From the TGA analysis, they identified that the carbon deposition amount onto the catalyst was low. From the quick catalyst deactivation, they made the conclusion that the catalyst deactivation was mainly due to sulfur poison.
In Figure 8, the result of the poisoned catalyst regeneration under the O2999 treatment is shown. During this process, the surface-adsorbed sulfur may be oxidized, and the Ni sites are recovered back to NiO:
Ni-S + 3/2 O2 ↔ NiO + SO2
With the introduction of H2 gas flow for two h, NiO is reduced into the active Ni sites. It is seen by using the regenerated catalyst that XCH4 and XCO2 increase with the increased test time and approach to constant values that can result, as shown in Figure 2. This indicated that the sulfur can be almost completely removed by the formation of SO2 based on Equation (16). It was pointed out by Li et al. [41] that NiSO4 is formed instead of SO2 when a high O2 flow was used based on the reaction
Ni-S + 2O2 ↔ NiSO4
In this case, the catalyst would lose its activity in the DRM.
Similar to the O2 treatment, the steam treatment for poisoned catalyst regeneration used the steam flow instead of the O2 flow. The steam treatment removes sulfur in the form of SO2 and H2S and oxidizes Ni to NiO via the following reactions [42]:
Ni-S + H2O ↔ NiO + H2S
H2S + 2H2O ↔ SO2 + 3H2
Ni + H2O ↔ NiO + H2
By introducing a H2 gas flow for two h, NiO is reduced into the active Ni sites. In Figure 9, the performance of the steam-regenerated catalyst is shown. It is seen that a certain amount of sulfur absorbed on the catalyst is not effectively removed during the regeneration treatment. This may be due to the relatively short time of the steam treatment. As a result, decreases in both XCH4 and XCO2 were observed by using the partially regenerated catalyst.

5. Conclusions

In this study, the performance of DRM using biogas as a feedstock was studied, including the catalyst poison, by the presence of H2S in the reactant. Using the nonpoisoned DRM performance as a comparison basis, the following conclusions can be made:
(1)
The catalyst poison depends on both the reaction temperature and time. The H2S coverage onto the catalyst surface decreases with the increased reaction temperature.
(2)
Due to the stronger chemisorption of sulfur onto the catalyst as compared to O2 or H2O, catalyst deactivation cannot be regenerated by the bi-reforming of methane in which DRM is combined with POM, COM, or SRM.
(3)
The catalyst cannot be regenerated for the poison that occurs at low temperatures.
(4)
The poisoned catalyst can be effectively regenerated using a high-temperature oxidation process. A higher reaction time is required for the catalyst regenerated by the high-temperature steam process.

Author Contributions

R.-Y.C. created the research concept, performed the experiments, analyzed data, and wrote the paper. Y.-C.C. performed the experiments and collected and analyzed the data. W.-H.C. organized the work, designed the experiments, analyzed data, and wrote the paper. All authors have read and agreed to the published version of the manuscript.

Funding

This research was funded by the Ministry of Science and Technology, Taiwan, R.O.C., under the contracts MOST 108-2221-E-006-127-MY3 and MOST 109-2622-E-006-006-CC1.

Data Availability Statement

Not applicable.

Acknowledgments

The authors acknowledge the financial support of the Ministry of Science and Technology, Taiwan, R.O.C., under the contracts MOST 108-2221-E-006-127-MY3 and MOST 109-2622-E-006-006-CC1 for this research.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Sadhwani, N.; Adhikari, S.; Eden, M. Biomass Gasification Using Carbon Dioxide: Effect of Temperature, CO2/C Ratio, and the Study of Reactions Influencing the Process. Ind. Eng. Chem. Res. 2016, 55, 2883–2891. [Google Scholar] [CrossRef]
  2. Chen, W.-H.; Lin, S.-C. Biogas partial oxidation in a heat recirculation reactor for syngas production and CO2 utilization. Appl. Energy 2018, 217, 113–125. [Google Scholar] [CrossRef]
  3. Hernández, B.; Martín, M. Optimal Process Operation for Biogas Reforming to Methanol: Effects of Dry Reforming and Biogas Composition. Ind. Eng. Chem. Res. 2016, 55, 6677–6685. [Google Scholar] [CrossRef]
  4. Kang, J.Y.; Kang, D.W.; Kim, T.S.; Hur, K.B. Comparative economic analysis of gas turbine-based power generation and combined heat and power systems using biogas fuel. Energy 2014, 67, 309–318. [Google Scholar] [CrossRef]
  5. Tamoor, M.; Tahir, M.S.; Sagir, M.; Iqbal, S.; Nawaz, T. Design of 3 kW integrated power generation system from solar and biogas. Int. J. Hydrogen Energy 2020, 45, 12711–12720. [Google Scholar] [CrossRef]
  6. Gao, Y.; Jiang, J.; Meng, Y.; Yan, F.; Aihemaiti, A. A review of recent developments in hydrogen production via biogas dry reforming. Energy Convers. Manag. 2018, 171, 133–155. [Google Scholar] [CrossRef]
  7. Zhao, X.; Joseph, B.; Kuhn, J.; Ozcan, S. Biogas Reforming to Syngas: A Review. iScience 2020, 23, 101082. [Google Scholar] [CrossRef] [PubMed]
  8. Jung, S.; Lee, J.; Moon, D.H.; Kim, K.-H.; Kwon, E.E. Upgrading biogas into syngas through dry reforming. Renew. Sustain. Energy Rev. 2021, 143, 110949. [Google Scholar] [CrossRef]
  9. Chen, W.-H.; Lin, M.-R.; Lu, J.-J.; Chao, Y.; Leu, T.-S. Thermodynamic analysis of hydrogen production from methane via autothermal reforming and partial oxidation followed by water gas shift reaction. Int. J. Hydrogen Energy 2010, 35, 11787–11797. [Google Scholar] [CrossRef]
  10. Cañete, B.; Gigola, C.E.; Brignole, N.B. Synthesis Gas Processes for Methanol Production via CH4Reforming with CO2, H2O, and O2. Ind. Eng. Chem. Res. 2014, 53, 7103–7112. [Google Scholar] [CrossRef]
  11. James, O.O.; Mesubi, A.M.; Ako, T.C.; Maity, S. Increasing carbon utilization in Fischer–Tropsch synthesis using H2-deficient or CO2-rich syngas feeds. Fuel Process. Technol. 2010, 91, 136–144. [Google Scholar] [CrossRef]
  12. Wang, C.; Wang, Y.; Chen, M.; Liang, D.; Yang, Z.; Cheng, W.; Tang, Z.; Wang, J.; Zhang, H. Recent advances during CH4 dry reforming for syngas production: A mini review. Int. J. Hydrogen Energy 2021, 46, 5852–5874. [Google Scholar] [CrossRef]
  13. Argyle, M.D.; Bartholomew, C.H. Heterogeneous Catalyst Deactivation and Regeneration: A Review. Catalysts 2015, 5, 145. [Google Scholar] [CrossRef] [Green Version]
  14. Meloni, E.; Martino, M.; Palma, V. A Short Review on Ni Based Catalysts and Related Engineering Issues for Methane Steam Reforming. Catalysts 2020, 10, 352. [Google Scholar] [CrossRef] [Green Version]
  15. Chein, R.; Yang, Z. H2S effect on dry reforming of biogas for syngas production. Int. J. Energy Res. 2019, 43, 3330–3345. [Google Scholar] [CrossRef]
  16. Jablonski, W.S.; Villano, S.M.; Dean, A.M. A comparison of H2S, SO2, and COS poisoning on Ni/YSZ and Ni/K2O-CaAl2O4 during methane steam and dry reforming. Appl. Catal. A Gen. 2015, 502, 399–409. [Google Scholar] [CrossRef] [Green Version]
  17. Rostrup-Nielsen, J.; Christiansen, L.J. Concepts in Syngas Manufacture; World Scientific: Singapore, 2011. [Google Scholar]
  18. Izquierdo, U.; García-García, I.; Gutierrez, Á.M.; Arraibi, J.R.; Barrio, V.L.; Cambra, J.F.; Arias, P.L. Catalyst Deactivation and Regeneration Processes in Biogas Tri-Reforming Process. The Effect of Hydrogen Sulfide Addition. Catalysts 2018, 8, 12. [Google Scholar] [CrossRef] [Green Version]
  19. Wachter, P.; Gaber, C.; Raic, J.; Demuth, M.; Hochenauer, C. Experimental investigation on H2S and SO2 sulphur poisoning and regeneration of a commercially available Ni-catalyst during methane tri-reforming. Int. J. Hydrogen Energy 2021, 46, 3437–3452. [Google Scholar] [CrossRef]
  20. Wang, S.; Lu, G.Q.; Millar, G.J. Carbon Dioxide Reforming of Methane To Produce Synthesis Gas over Metal-Supported Catalysts: State of the Art. Energy Fuels 1996, 10, 896–904. [Google Scholar] [CrossRef]
  21. Chen, W.-H.; Chen, C.-Y. Water gas shift reaction for hydrogen production and carbon dioxide capture: A review. Appl. Energy 2020, 258, 114078. [Google Scholar] [CrossRef]
  22. Fidalgo, B.; Muradov, N.; Menéndez, J.A. Effect of H2S on carbon-catalyzed methane decomposition and CO2 reforming reactions. Int. J. Hydrogen Energy 2012, 37, 14187–14194. [Google Scholar] [CrossRef]
  23. Rostrup-Nielsen, J.R. New aspects of syngas production and use. Catal. Today 2000, 63, 159–164. [Google Scholar] [CrossRef]
  24. Bartholomew, C.; Agrawal, P.; Katzer, J. Sulfur Poisoning of Metals. Adv. Catal. 1982, 31, 135–242. [Google Scholar]
  25. Chein, R.-Y.; Fung, W.-Y. Syngas production via dry reforming of methane over CeO2 modified Ni/Al2O3 catalysts. Int. J. Hydrogen Energy 2019, 44, 14303–14315. [Google Scholar] [CrossRef]
  26. Ashrafi, M.; Pfeifer, C.; Pröll, T.; Hofbauer, H. Experimental Study of Model Biogas Catalytic Steam Reforming: 2. Impact of Sulfur on the Deactivation and Regeneration of Ni-Based Catalysts. Energy Fuels 2008, 22, 4190–4195. [Google Scholar] [CrossRef]
  27. Yang, X. An experimental investigation on the deactivation and regeneration of a steam reforming catalyst. Renew. Energy 2017, 112, 17–24. [Google Scholar] [CrossRef]
  28. Chen, X.; Jiang, J.; Yan, F.; Li, K.; Tian, S.; Gao, Y.; Zhou, H. Dry Reforming of Model Biogas on a Ni/SiO2 Catalyst: Overall Performance and Mechanisms of Sulfur Poisoning and Regeneration. ACS Sustain. Chem. Eng. 2017, 5, 10248–10257. [Google Scholar] [CrossRef]
  29. Kalai, D.Y.; Stangeland, K.; Tucho, W.M.; Jin, Y.; Yu, Z. Biogas reforming on hydrotalcite-derived Ni-Mg-Al catalysts: The effect of Ni loading and Ce promotion. J. CO2 Util. 2019, 33, 189–200. [Google Scholar] [CrossRef]
  30. Niu, J.; Liland, S.E.; Yang, J.; Rout, K.R.; Ran, J.; Chen, D. Effect of oxide additives on the hydrotalcite derived Ni catalysts for CO2 reforming of methane. Chem. Eng. J. 2019, 377, 119763. [Google Scholar] [CrossRef]
  31. Newnham, J.; Mantri, K.; Amin, M.H.; Tardio, J.; Bhargava, S.K. Highly stable and active Ni-mesoporous alumina catalysts for dry reforming of methane. Int. J. Hydrogen Energy 2012, 37, 1454–1464. [Google Scholar] [CrossRef]
  32. Appari, S.; Janardhanan, V.M.; Bauri, R.; Jayanti, S. Deactivation and regeneration of Ni catalyst during steam reforming of model biogas: An experimental investigation. Int. J. Hydrogen Energy 2014, 39, 297–304. [Google Scholar] [CrossRef] [Green Version]
  33. Chattanathan, S.A.; Adhikari, S.; McVey, M.; Fasina, O. Hydrogen production from biogas reforming and the effect of H2S on CH4 conversion. Int. J. Hydrogen Energy 2014, 39, 19905–19911. [Google Scholar] [CrossRef]
  34. Dou, X.; Veksha, A.; Chan, W.P.; Oh, W.-D.; Liang, Y.N.; Teoh, F.; Mohamed, D.K.B.; Giannis, A.; Lisak, G.; Lim, T.-T. Poisoning effects of H2S and HCl on the naphthalene steam reforming and water-gas shift activities of Ni and Fe catalysts. Fuel 2019, 241, 1008–1018. [Google Scholar] [CrossRef]
  35. Chein, R.; Chen, Y.; Yu, C.; Chung, J. Thermodynamic analysis of dry reforming of CH4 with CO2 at high pressures. J. Nat. Gas Sci. Eng. 2015, 26, 617–629. [Google Scholar] [CrossRef]
  36. Nikoo, M.K.; Amin, N. Thermodynamic analysis of carbon dioxide reforming of methane in view of solid carbon formation. Fuel Process. Technol. 2011, 92, 678–691. [Google Scholar] [CrossRef] [Green Version]
  37. Kumar, N.; Shojaee, M.; Spivey, J. Catalytic bi-reforming of methane: From greenhouse gases to syngas. Curr. Opin. Chem. Eng. 2015, 9, 8–15. [Google Scholar] [CrossRef] [Green Version]
  38. Olah, G.A.; Goeppert, A.; Czaun, M.; Prakash, G.K.S. Bi-reforming of Methane from Any Source with Steam and Carbon Dioxide Exclusively to Metgas (CO–2H2) for Methanol and Hydrocarbon Synthesis. J. Am. Chem. Soc. 2013, 135, 648–650. [Google Scholar] [CrossRef]
  39. Pawar, V.; Appari, S.; Monder, D.S.; Janardhanan, V.M. Study of the Combined Deactivation Due to Sulfur Poisoning and Carbon Deposition during Biogas Dry Reforming on Supported Ni Catalyst. Ind. Eng. Chem. Res. 2017, 56, 8448–8455. [Google Scholar] [CrossRef]
  40. Sehested, J. Four challenges for nickel steam-reforming catalysts. Catal. Today 2006, 111, 103–110. [Google Scholar] [CrossRef]
  41. Li, L.; Howard, C.; King, D.L.; Gerber, M.; Dagle, R.; Stevens, D. Regeneration of Sulfur Deactivated Ni-Based Biomass Syngas Cleaning Catalysts. Ind. Eng. Chem. Res. 2010, 49, 10144–10148. [Google Scholar] [CrossRef]
  42. Lombard, C.; Ledoze, S.; Marencak, E.; Marquaire, P.; Lenoc, D.; Bertrand, G.; Lapicque, F. In situ regeneration of the Ni-based catalytic reformer of a 5kW PEMFC system. Int. J. Hydrogen Energy 2006, 31, 437–440. [Google Scholar] [CrossRef]
Figure 1. A schematic of the experimental setup.
Figure 1. A schematic of the experimental setup.
Catalysts 11 00777 g001
Figure 2. Effect of the reaction temperature on the DRM performance without a H2S presence in the reactant. CO2/CH4 = 1. (a) CH4 conversion, (b) CO2 conversion, (c) H2 yield, (d) CO yield, and (e) H2/CO ratio.
Figure 2. Effect of the reaction temperature on the DRM performance without a H2S presence in the reactant. CO2/CH4 = 1. (a) CH4 conversion, (b) CO2 conversion, (c) H2 yield, (d) CO yield, and (e) H2/CO ratio.
Catalysts 11 00777 g002
Figure 3. Effect of the reaction temperature on the DRM performance with a H2S presence in the reactant. CO2/CH4 = 1. (a) CH4 conversion and (b) CO2 conversion °C.
Figure 3. Effect of the reaction temperature on the DRM performance with a H2S presence in the reactant. CO2/CH4 = 1. (a) CH4 conversion and (b) CO2 conversion °C.
Catalysts 11 00777 g003
Figure 4. Effect of the CH4/CO2 ratio on the DRM performance with the presence of H2S in the reactant. T = 800 °C. (a) CH4 conversion and (b) CO2 conversion.
Figure 4. Effect of the CH4/CO2 ratio on the DRM performance with the presence of H2S in the reactant. T = 800 °C. (a) CH4 conversion and (b) CO2 conversion.
Catalysts 11 00777 g004
Figure 5. Effect of O2 addition on the DRM performance. T = 800 °C. (a) CH4 conversion and (b) CO2 conversion.
Figure 5. Effect of O2 addition on the DRM performance. T = 800 °C. (a) CH4 conversion and (b) CO2 conversion.
Catalysts 11 00777 g005
Figure 6. Effect of the H2O addition on the DRM performance. T = 800 °C. (a) CH4 conversion and (b) CO2 conversion.
Figure 6. Effect of the H2O addition on the DRM performance. T = 800 °C. (a) CH4 conversion and (b) CO2 conversion.
Catalysts 11 00777 g006
Figure 7. Effect of the reaction temperature on catalyst regeneration. CH4/CO2 = 1. (a) CH4 conversion and (b) CO2 conversion.
Figure 7. Effect of the reaction temperature on catalyst regeneration. CH4/CO2 = 1. (a) CH4 conversion and (b) CO2 conversion.
Catalysts 11 00777 g007
Figure 8. Poisoned catalyst regeneration using the high-temperature oxidation treatment. T = 800 °C and CH4/CO2 = 1. The catalyst was poisoned from 0 to 12 h, regenerated by 10-sccm air from 12 to 14 h to form NiO, reduced NiO to Ni by 20-sccm H2 from 14 to 16 h, and DRM-tested using the regenerated catalyst from 16 to 33 h.
Figure 8. Poisoned catalyst regeneration using the high-temperature oxidation treatment. T = 800 °C and CH4/CO2 = 1. The catalyst was poisoned from 0 to 12 h, regenerated by 10-sccm air from 12 to 14 h to form NiO, reduced NiO to Ni by 20-sccm H2 from 14 to 16 h, and DRM-tested using the regenerated catalyst from 16 to 33 h.
Catalysts 11 00777 g008
Figure 9. Poisoned catalyst regeneration using steam. T = 800 °C and CH4/CO2 = 1. The catalyst was poisoned from 0 to 12 h, regenerated by 10-sccm oxygen from 12 to 14 h to form NiO, reduced NiO to Ni by 20-sccm H2 from 14 to 16 h, and DRM-tested using the regenerated catalyst from 16 to 33 h.
Figure 9. Poisoned catalyst regeneration using steam. T = 800 °C and CH4/CO2 = 1. The catalyst was poisoned from 0 to 12 h, regenerated by 10-sccm oxygen from 12 to 14 h to form NiO, reduced NiO to Ni by 20-sccm H2 from 14 to 16 h, and DRM-tested using the regenerated catalyst from 16 to 33 h.
Catalysts 11 00777 g009
Table 1. Composition of simulated biogas used in the experiment.
Table 1. Composition of simulated biogas used in the experiment.
CaseMolar RatioCH4 (sccm)CO2 (sccm)N2 (sccm)
1CH4/CO2 = 1/0.5157.527.5
2CH4/CO2 = 1/1151520
3CH4/CO2= 1/215305
Table 2. Comparison between the present study and data reported in the literature without H2S poison. CH4/CO2 = 1/1.
Table 2. Comparison between the present study and data reported in the literature without H2S poison. CH4/CO2 = 1/1.
Present StudyRef [29]Ref [30]Ref [31]
TemperatureXCH4XCO2XCH4XCO2XCH4XCO2XCH4XCO2
700 °C55%68%53%68%53%64%56%70%
800 °C78%91%82%88%80%85%80%89%
900 °C88%96%N/AN/AN/AN/A94%98%
Table 3. Details of the poisoned catalyst regeneration treatments.
Table 3. Details of the poisoned catalyst regeneration treatments.
RegenerationDetail Conditions
High-temperature reactionT = 700~900 °C. The catalyst was poisoned from 0 to 7 h, followed by a H2S-free DRM test for 8 h.
high-temperature steamT = 800 °C. The catalyst was poisoned from 0 to 12 h, regenerated by 10 sccm steam from 12 to 14 h to form NiO, reduced NiO to Ni by 20 sccm H2 from 14 to 16 h, and DRM-tested using regenerated catalyst from 16 to 33 h.
High-temperature oxidationT = 800 °C. The catalyst was poisoned from 0 to 12 h, regenerated by 10 sccm air from 12 to 14 h to form NiO, reduced NiO to Ni by 20 sccm H2 from 14 to 16 h, and DRM test using regenerated catalyst from 16 to 33 h.
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Chein, R.-Y.; Chen, Y.-C.; Chen, W.-H. Experimental Study on Sulfur Deactivation and Regeneration of Ni-Based Catalyst in Dry Reforming of Biogas. Catalysts 2021, 11, 777. https://doi.org/10.3390/catal11070777

AMA Style

Chein R-Y, Chen Y-C, Chen W-H. Experimental Study on Sulfur Deactivation and Regeneration of Ni-Based Catalyst in Dry Reforming of Biogas. Catalysts. 2021; 11(7):777. https://doi.org/10.3390/catal11070777

Chicago/Turabian Style

Chein, Rei-Yu, Yen-Chung Chen, and Wei-Hsin Chen. 2021. "Experimental Study on Sulfur Deactivation and Regeneration of Ni-Based Catalyst in Dry Reforming of Biogas" Catalysts 11, no. 7: 777. https://doi.org/10.3390/catal11070777

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