Next Article in Journal
Natural Products and Biological Activity from Actinomycetes Associated with Marine Algae
Next Article in Special Issue
Preparation of Cellulose-Based Activated Carbon Fibers with Improved Yield and Their Methylene Chloride Adsorption Evaluation
Previous Article in Journal
Controllable Synthesis of ZnO Nanoparticles with Improved Photocatalytic Performance for the Degradation of Rhodamine B under Ultraviolet Light Irradiation
Previous Article in Special Issue
Characterization of Dissolved Organic Matter Released from Aged Biochar: A Comparative Study of Two Feedstocks and Multiple Aging Approaches
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

A Review of the Design and Performance of Catalysts for Hydrothermal Gasification of Biomass to Produce Hydrogen-Rich Gas Fuel

1
Department of Chemical and Biological Engineering, University of Saskatchewan, Saskatoon, SK S7N 5A9, Canada
2
Department of Engineering, Faculty of Agriculture, Dalhousie University, Truro, NS B2N 5E3, Canada
*
Author to whom correspondence should be addressed.
Molecules 2023, 28(13), 5137; https://doi.org/10.3390/molecules28135137
Submission received: 15 May 2023 / Revised: 28 June 2023 / Accepted: 29 June 2023 / Published: 30 June 2023
(This article belongs to the Collection Recycling of Biomass Resources: Biofuels and Biochemicals)

Abstract

:
Supercritical water gasification has emerged as a promising technology to sustainably convert waste residues into clean gaseous fuels rich in combustible gases such as hydrogen and methane. The composition and yield of gases from hydrothermal gasification depend on process conditions such as temperature, pressure, reaction time, feedstock concentration, and reactor geometry. However, catalysts also play a vital role in enhancing the gasification reactions and selectively altering the composition of gas products. Catalysts can also enhance hydrothermal reforming and cracking of biomass to achieve desired gas yields at moderate temperatures, thereby reducing the energy input of the hydrothermal gasification process. However, due to the complex hydrodynamics of supercritical water, the literature is limited regarding the synthesis, application, and performance of catalysts used in hydrothermal gasification. Hence, this review provides a detailed discussion of different heterogeneous catalysts (e.g., metal oxides and transition metals), homogeneous catalysts (e.g., hydroxides and carbonates), and novel carbonaceous catalysts deployed in hydrothermal gasification. The article also summarizes the advantages, disadvantages, and performance of these catalysts in accelerating specific reactions during hydrothermal gasification of biomass, such as water–gas shift, methanation, hydrogenation, reforming, hydrolysis, cracking, bond cleavage, and depolymerization. Different reaction mechanisms involving a variety of catalysts during the hydrothermal gasification of biomass are outlined. The article also highlights recent advancements with recommendations for catalytic supercritical water gasification of biomass and its model compounds, and it evaluates process viability and feasibility for commercialization.

1. Introduction

Owing to increased growth in the population as well as urban and industrial development, global energy consumption has witnessed a dramatic rise over the years. Currently, 80% of the global energy demand is met by fossil fuels such as coal, natural gas, gasoline, and diesel. It cannot be denied that fossil fuels have long-term adverse effects on the environment and ecosystems, including global warming, an increase in greenhouse gas emissions, acid rain, and changes in weather patterns, to name a few [1]. On a global scale, CO2 emissions from the usage of fossil fuels such as coal, crude oil, and natural gas amount to 15, 12, and 8 billion tons, respectively [2]. Gradually phasing away from fossil fuels and seeking alternative and renewable sources of energy are urgently required.
Biofuels produced from renewable sources such as lignocellulosic biomass, livestock manure, microalgae, municipal solid waste, and sewage sludge are desirable alternatives to fossil fuels for meeting future energy demands and reducing carbon emissions [3,4]. Hydrogen (H2) has proven to be a clean alternative source to fossil fuels for meeting energy demands because of its zero carbon emissions, higher heating value of 140 MJ/kg, and adiabatic flame temperature of approximately 2100 °C. The combustion products of H2 are water and heat energy, compared to the combustion of fossil fuels which emits greenhouse gases such as CO2, CO, CH4, SOx, and NOx. In addition to being considered the fuel of the future, hydrogen is also utilized in a wide variety of commercial applications such as fuel cells, upgrading crude oil, synthesis of fine chemicals, metallurgy, pharmaceuticals, and the aerospace industry [5]. A main advantage of H2 is its ability to produce clean electricity through fuel cells [6]. H2 is also a valuable precursor in the production of various commodity and specialty chemicals, such as methanol, ammonia, alcohol, and aldehydes, through various catalytic and non-catalytic thermochemical conversion processes [7,8]. H2 is also extensively used by refineries in hydrotreating processes such as hydrodeoxygenation [9], hydrodenitrogenation [10], hydrodesulfurization [10], and hydrodemetallization [11] to upgrade crude oil and bio-oil to transportation-grade fuels. The sustainable nature of H2 and its increasing demand in many industrial and commercial sectors has entrenched it as an integral component of the circular economy.
Although hydrogen gas has no color, its production routes have designated it different colors categorization. Hydrogen can be categorized as brown, grey, blue, green, pink, yellow, turquoise, and white based on its production from a wide variety of sources and technologies (Figure 1). Based on the source and production technology employed, hydrogen can be classified into different colors such as brown H2 (gasification of coal), grey H2 (steam reforming of methane), blue H2 (steam reforming of methane with carbon capture), green H2 (electrolysis using electricity from renewables), pink H2 (electrolysis using electricity from nuclear energy), turquoise H2 (methane pyrolysis), yellow H2 (electrolysis using electricity from solar power), and white H2 (geological H2 in underground deposits) [12].
Currently, the major route for the synthesis of H2 is the steam reforming of methane, which contributes to approximately 95% of global H2 production [13]. Nearly 250,000 standard cubic feet of CO2 is emitted per 1 million standard cubic feet of H2 produced from the steam reforming of CH4 [14]. Despite the significantly larger carbon footprint of the steam reforming of methane process, it is still commercially applied today. Although not widely commercialized, several sustainable pathways for hydrogen production from alternative sources are also available, such as electrolysis, photocatalysis, hydrothermal gasification, dark fermentation, and photo-fermentation [15].
The hydrothermal gasification conversion route is capable of sustainably producing H2 via renewable lignocellulosic biomass sources. This process utilizes water at either subcritical or supercritical conditions as a green solvent and reaction medium to disintegrate complex organic substrates to gases such as H2, CH4, CO, and CO2 [16]. When the temperature and pressure of water exceed its critical points of 374 °C and 22.1 MPa, respectively, supercritical water (SCW) is generated [17]. On the other hand, water is transformed into subcritical water when the temperature and pressure of water are slightly below or near its critical points.
SCW experiences a significant change in its properties compared to liquid water at room conditions, imparting unique properties such as faster kinetics, a non-polar nature, and excellent solubility of gaseous molecules with the absence of interphase transfer boundaries [18,19]. Due to these versatile properties, supercritical water gasification (SCWG) can convert recalcitrant feedstocks with high moisture content into gaseous fuels enriched with combustible gases such as H2 and CH4. SCWG also does not require biomass drying because of its aqueous reaction medium, making the process energy efficient [20]. Due to these advantages, SCWG has recently gained popularity as an environmentally friendly process to produce H2 from waste feedstocks.
The main products of SCWG are gases (e.g., H2, CO, CO2, CH4, and C2+), hydrochar, and liquid effluents. Hydrochar is a carbon-rich solid product resulting from depolymerization, dehydrogenation, decarboxylation, deamination, and aromatization of the organic feedstock used in SCWG [21]. Further activation and functionalization of hydrochar can enhance its surface area and properties for a wide variety of applications, such as solid fuel, adsorbent, catalyst support, activated carbon, carbon sequestration product, reinforcing material for composites, and soil amendment agent [22,23,24]. The liquid effluents resulting from the hydrothermal decomposition of biomass contain alcohols, furfurals, carboxylic acids, esters, ethers, aliphatics, aldehydes, ketones, and phenolics [25]. Some of these degradation compounds may further polymerize to form tar, which is a challenging component that causes plugging as well as heat and mass transfer limitations in the processors [26]. It should be noted that process conditions such as temperature, reaction time, pressure, and feedstock concentration largely impact the yields and composition of gases, liquids, and hydrochar from the SCWG of biomass [20].
Catalysts also play an important role in improving the process efficiencies of SCWG, especially carbon gasification efficiency and selectivities of gases, by regulating specific reaction mechanisms [27,28]. Several homogeneous and heterogeneous catalysts have been designed and investigated for the SCWG of different biomasses. However, the literature on the application of different catalysts in SCWG appears to be scattered. In addition, in-depth knowledge is scarce on understanding the different reaction pathways, mechanisms, and product properties impacted by homogeneous and heterogeneous catalysts in SCWG. Hence, this review paper attempts to categorically summarize the recent advancements in different homogeneous and heterogeneous catalysts used in SCWG. Furthermore, the challenges and shortcomings of different catalysts are also identified, followed by a discussion and recommendations for the effective design of catalysts, catalytic supports, and promoters used in the SCWG of biomass to produce high-value gaseous fuels.

2. Homogeneous Catalysts Used in Hydrothermal Gasification

Homogeneous catalysts used in SCWG generally consist of alkali metal and hydroxide catalysts. Table 1 summarizes some notable studies on the application of homogeneous catalysts in the SCWG process [29,30,31,32,33,34,35]. Homogeneous catalysts promote water–gas shift reactions by favoring C–C bond breakup, thus improving H2 yields [36]. The water–gas shift reaction results in the formation of H2 and CO2 because of the reaction of CO and H2O. The produced H2 can further react with the reactive intermediates generated by the catalytic action of homogeneous catalysts to increase overall gas yields [31]. Homogeneous catalysts usually have rapid conversion rates and can be used in both batch and continuous reactors. Homogeneous catalysts are also cost-effective with negligible sintering [26].
Su et al. [37] reported a base-catalyzed mechanism of alkali metals that enabled the water–gas shift reaction. The degradation intermediates were anions comprising hydroxides, carbonates and formates. Mixing the carbonates in water produced CO2 and hydroxides. Hydroxides can further combine with CO to produce formate. Further decomposition of formaldehyde can generate H2. Watanabe et al. [38] reported an ionic-catalyzed mechanism of alkali metals in the SCWG of methanol. They proposed that the ionic species stabilized the methanol by protonation or disassociation. Methanol then oxidized into CO, and protons stabilized the produced CO. CO2 was formed via the oxidation of CO, and hydroxide ions favored the water–gas shift reaction to convert CO into CO2. Thus, the oxidization of CO to CO2 was enhanced by the hydroxyl ions.
Figure 2 represents a simplified catalytic mechanism of potassium metal in the SCWG of biomass [39]. Sınaǧ et al. [32] compared K2CO3 (a homogeneous catalyst) with Raney nickel (a heterogeneous catalyst) in the SCWG of glucose. Their results showed that the catalytic action of K2CO3 enhanced H2 production while suppressing the formation of phenols for improved gasification efficiency. K2CO3 demonstrated superior catalytic activity as compared to Raney nickel. K2CO3 showed higher yields of H2 and CO2 than Raney nickel, which confirmed its catalytic action to promote the water–gas shift reaction. The catalytic mechanism of K2CO3 in enhancing water–gas shift via formate (HCOOK+) formation is presented in the following equations. The produced potassium formate further reacts with excess water to generate H2 with KHCO3, which decomposes into CO2 and K2CO3.
K2CO3 + H2O → KHCO3 + KOH
HCOOK + H2O → KHCO3 + H2
2KHCO3 → CO2 + K2CO3 + H2O
Madenoğlu et al. [30] studied the kinetics effects of K2CO3 in the SCWG of cellulose, lignin, and their mixtures. Their results showed that K2CO3 promoted the rates of gasification reactions and prevented the formation of char. Both gas and aqueous phase yields increased at the expense of char yield due to the catalytic effects of K2CO3. K2CO3 also favored the water–gas shift reaction, thus increasing the H2 yield.
Sınaǧ et al. [40] studied the catalytic effect of K2CO3 on glucose, phyto-mass (plant residues without proteins), and zoo-mass (meat residues containing proteins). The addition of K2CO3 had a significant influence in promoting the water–gas shift reaction during the SCWG of glucose and enhanced H2 production. However, its catalytic effects in promoting water–gas shift during the SCWG of phyto-mass and zoo-mass were minimal.
Nanda et al. [35] compared four different homogeneous catalysts, Na2CO3, K2CO3, NaOH, and KOH, in the SCWG of Timothy grass. An increase in catalyst loading from 1% to 3% increased the total gas yield, as well as H2, CH4, and CO2 yields, but decreased the CO yield for all catalysts. This indicated the catalytic action of alkali catalysts promoted gasification efficiency and the water–gas shift reaction. KOH showed the highest H2 yield of 9 mol/kg, followed by K2CO3, NaOH, and Na2CO3. A similar trend was observed for total gas yields. The highest total gas and H2 yields with KOH were explained by its catalytic action to promote the water–gas shift reaction. On the other hand, NaOH enhanced the methanation reaction with the consumption of H2, increasing CH4 yields. Nanda et al. [41] also confirmed the superior catalytic effects of KOH in the SCWG of fructose where KOH showed a higher H2 yield than NaOH with nearly three times more H2 yield than non-catalytic reactions.
Yanik et al. [42] compared the activities and selectivities of K2CO3, Trona, red mud, and Raney nickel catalysts in the SCWG of cotton stalk, corncob, and tannery wastes. Their results showed that all four catalysts significantly enhanced H2 yields by favoring water–gas shift and reforming reactions. K2CO3 demonstrated the highest H2 yield with no CO detected in the gas products. However, the catalytic activity of Trona was analogous to that of K2CO3. Ferreira-Pinto et al. [43] investigated the effects of NaOH, KOH, and Na2CO3 catalysts in the SCWG of lactose. The increase in H2 yield was highest with NaOH, followed by Na2CO3 and KOH. All catalysts inhibited char formation and significantly reduced the total organic carbon content in the reactants, indicating high gasification efficiencies.
Alkali catalysts can also significantly reduce the sulfur content in gas products. High sulfur content in gas products is a serious issue as its combustion can release SOx. Sulfur can also deactivate and poison the catalysts as well as corrode pipelines. Feng et al. [34] used different homogeneous catalysts (e.g., KOH, K2CO3, NaOH, Na2CO3, and activated carbon or AC) in the SCWG of sewage sludge. K2CO3 showed the best desulfurization effect and limited the H2S and SO2 contents to around 140 ppm and 200 ppm, respectively. The order of desulfurization effects of catalysts was found to be: K2CO3 > Na2CO3 > NaOH > KOH > AC. KOH demonstrated the highest H2 yield and selectivity. Alkali catalysts converted the unstable sulfur compounds into stable sulfur compounds by promoting cyclization and oxidation reactions, thus preventing the migration of sulfur into gas and liquid products.
Zhong et al. [44] investigated the catalytic performance of KOH, K2CO3, KMnO4, and H2O2 on polycyclic aromatic hydrocarbons (PAHs) and gas formation during the SCWG of coking sludge. Their results showed that the PAH content decreased in the catalytic SCWG experiments. The catalytic action of KOH was attributed to its ability to promote free radical reactions during SCWG. These free radicals promote ring-opening reactions of PAHs, leading to their decomposition. KOH led to a higher H2 yield than K2CO3 because of an improved water–gas shift reaction through the formation of a formate intermediate and hydroxyl ions. These hydroxyl ions efficiently capture CO2 produced from the water–gas shift reaction. This shifted the equilibrium of the water–gas shift reaction towards the products side, thus producing more H2. Despite the numerous advantages of homogeneous catalysts, they can easily cause reactor plugging and corrosion in the reactor [45]. The recovery of homogeneous catalysts is also difficult compared to that of heterogeneous catalysts, which adds to overall process expenditures [46].

3. Heterogeneous Catalysts Used in Hydrothermal Gasification

Heterogeneous catalysts applied in the SCWG process can be broadly divided into two categories, namely metal oxides and transition metals. The recovery and recycling of heterogeneous catalysts are relatively easier compared to those of homogeneous catalysts [47]. Heterogeneous catalysts are more active, resulting in efficient and improved gasification efficiency [48]. They are also more selective for specific products by promoting desired reactions. A summary of promising studies on the use of heterogeneous catalysts in SCWG is presented in Table 2 [27,49,50,51,52,53,54,55].

3.1. Transition Metals

3.1.1. Nickel-Based Catalysts

Nickel-based catalysts are the most widely used heterogeneous catalysts in SCWG because of their high activity compared to other expensive transition metal catalysts. Ni-based catalysts require comparatively lower temperatures and promote biomass gasification with higher efficiency. However, Ni-based catalysts can also consume the produced H2, CO, and CO2 due to their high methanation activity, producing CH4 [56]. Furusawa et al. [57] used the Ni/MgO catalyst in the SCWG of lignin. They studied its regenerative capabilities by recovering and reusing the catalyst thrice. The catalyst showed satisfactory regenerative capability before suffering from deactivation due to the formation of carbon and Mg(OH)2.
Zhang et al. [58] studied the SCWG of glucose and compared the activities and H2 selectivities of Ni, Co, Ru, and Cu transition metals on γ-Al2O3, AC, and ZrO2 supports. Both 10%Ni/γ-Al2O3 and 10%Ru/Al2O3 demonstrated the highest catalytic activities and H2 selectivities. The order of activity of the supports for the Ni catalyst was: γ-Al2O3 > ZrO2 > AC. Due to satisfactory results with 10%Ni/γ-Al2O3, further enhancement with Na, K, Mg, and Ru promotors was also studied. The addition of the 0.5%K promoter on 10%Ni/γ-Al2O3 significantly increased the H2 yield by favoring the water–gas shift reaction.
Azadi et al. [28] studied the SCWG of various lignocellulosic feedstocks (e.g., glucose, fructose, cellulose, pulp, xylan, bark, and lignin) using five transition metals catalysts (e.g., Ni/Al2O3, Ru/C, Raney nickel, Ni/hydrotalcite, and Ru/Al2O3). The activities of Ni/Al2O3 and Ni/hydrotalcite catalysts for SCWG demonstrated the highest H2 selectivities. In contrast, Raney nickel showed the lowest H2 selectivity. Ni/α-Al2O3 and Ni/hydrotalcite also demonstrated low CH4 yields at high temperatures and longer reaction times. The high H2 selectivities of Ni/α-Al2O3 and Ni/hydrotalcite were attributed to the lower nickel dispersion and large crystallite sizes of Ni/α-Al2O3 and Ni/hydrotalcite catalysts compared to Raney nickel. The high nickel dispersion of Raney nickel strongly favored C–O bond cleavage compared to Ni/Al2O3 and Ni/hydrotalcite catalysts, thus explaining the low H2 selectivity of Raney nickel. The authors also reported that among all feedstocks, lignin was the most resistant to SCWG because of its branched polymeric structure. The lowest gas yield obtained from lignin was attributed to potential deactivation of the catalysts due to its sulfur content.
Azadi et al. [27] compared Ni catalysts on different support materials, including γ-Al2O3, α-Al2O3, activated carbon, carbon nanotubes (CNT), hydrotalcite, MgO, SiO2, silica gel, TiO2, ZrO2, and various zeolites in the SCWG of glucose. The 20%Ni/α-Al2O3 catalyst showed the highest H2 selectivity, and Ni/CNT demonstrated high H2 yields (17–24 mmol/g) and high stability with maximum carbon gasification efficiency. On the other hand, Ni/MgO demonstrated a better H2 yield (26 mmol/g) and satisfactory carbon gasification efficiency. Due to its low cost and high stability, the authors further investigated the Ni/α-Al2O3 catalyst by varying Ni loading and using promoters. Tin increased the H2 selectivity but decreased the catalytic activity, whereas alkali promoters increased the carbon gasification efficiency but decreased the H2 selectivity. Lu et al. [50] also studied Ni-based catalysts with various promoted Al2O3 supports (e.g., CeO2/Al2O3, MgO/Al2O3, La2O3/Al2O3, and ZrO2/Al2O3) in the SCWG of glucose. CeO2/Al2O3 showed the highest H2 yield, followed by La2O3/Al2O3, ZrO2/Al2O3, Al2O3, and MgO/Al2O3.
Onwudili and Williams [53] investigated the catalytic SCWG of various plastic wastes with Ru and Ni catalysts. By increasing RuO2 loading up to 5 wt% in the SCWG of low-density polyethylene, the H2 yield rose from 1 to 9.9 mol/kg at 450 °C in 1 h. However, the subsequent increase in RuO2 loading from 5 wt% to 20 wt% decreased the H2 yield to 4.9 mol/kg while increasing the hydrogen gasification and carbon gasification efficiency. By using a 20 wt% RuO2-γ-Al2O3 catalyst, polypropylene produced a high H2 yield and the highest carbon gasification efficiency of 99%. High- and low-density polyethylenes also showed similar gas yields, whereas polystyrene produced the lowest yields of C2-C4 gases. Low-density polyethylene demonstrated the highest H2 yield, followed by polystyrene, polypropylene, and high-density polyethylene.
Adamu et al. [59] studied Ce-mesoAl2O3 support impregnated with Ni in the SCWG of glucose (Figure 3). Ce-mesoAl2O3 had superior support properties compared to γ-Al2O3, such as moderate acidity, which helped to reduce coke formation and enabled high metal loading with low agglomeration. The Ni(20)/Ce-Al2O3 catalyst exhibited a very high H2 yield of 10.2 mol/mol of glucose. The meso-form led to the cracking of large intermediates such as tar compounds. Furthermore, Ce helped to improve the thermal stability of the alumina support.
Lu et al. [51] compared Ni, Cu, and Fe transition metals supported on MgO in the SCWG of wheat straw. The H2 yields varied with the application of different catalysts in the following order: Ni/MgO > Fe/MgO > Cu/MgO. Due to excellent H2 selectivity with Ni, the authors explored various supports, such as basic oxides (MgO and ZnO), acidic oxide (Al2O3), and amphoteric oxide (ZrO2). The H2 selectivities of Ni-supported catalysts varied in the order of Ni/MgO > Ni/ZnO > Ni/ Al2O3 > Ni/ZrO. Although the type of support had a minimal effect on H2 yield, a significant effect was observed on the decrease in CO yield. Basic oxide supports such as MgO and ZnO favored water–gas shift reactions, thus increasing H2 yields. The acidic support such as Al2O3 did not enhance the water–gas shift reaction. Hence, Ni/Al2O3 showed nearly double the CO yield as compared to the Ni/ZnO and Ni/MgO catalysts.
Okolie et al. [54] performed the SCWG of soybean straw using different Ni-based catalysts, catalyst supports, and promoters. ZrO2 and Al2O3 proved to be the most effective supports for Ni-based catalysts. Both 10%Ni-ZrO2 and 10%Ni-Al2O3 demonstrated higher H2 yields than other catalyst supports (e.g., CNT, SiO2/Al2O3, SiO2, and AC). Therefore, the authors further studied the effects of K, Na, and Ce promotors on Ni-based catalysts supported by ZrO2 and Al2O3. The 10%Ni-1%Ce/ZrO2 catalyst demonstrated the highest H2 yield of 10.9 mmol/g, followed by 10%Ni-1%K/ZrO2 and 10%Ni-1%Na/ZrO2. The relative increment in H2 yield and total gas yield without using any promoters was more substantial with the Ce and K promotors than with the Na promotor. However, the Na promotor showed the highest H2 yield with the Al2O3 support compared to the K and Ce promotors. The 10%Ni-1%Na/Al2O3 catalyst demonstrated the highest H2 yield (10.8 mmol/g) compared to 10%Ni-Ce/Al2O3 and 10%Ni-1%K/Al2O3. The 10%Ni-1%Ce/ZrO2 catalyst demonstrated an improved H2 yield and excellent catalytic performance. Further analysis revealed that the Ce promotor could store oxygen species and eliminate coke formation and sintering of the catalysts, resulting in its high performance.
Su et al. [60] investigated the effects of La2O3 in promoting the Ni-La2O3/θ-Al2O3 catalyst in the SCWG of food waste. La enhanced the water–gas shift reaction, resulting in a high H2 yield. La also inhibited the methanation reaction, which is a major limitation of Ni-based catalysts. La also improved the metal dispersion, which increased the catalytic activity. Chowdhury et al. [61] also reported that Ni/Al2O3 with an La promoter can lead to excellent catalytic activity in the SCWG of food waste. Ni/9%La-Al2O3 showed high H2 and gas yields. La improved the mesoporous structure and increased the dispersion of Ni, which enhanced the water–gas shift reaction and increased the H2 yield. Ni/9%La-Al2O3 also demonstrated high stability, which could be attributed to its better anti-carbon deposition property.
Mastuli et al. [62] compared doped and supported Zn and Ni catalysts on MgO support in the SCWG of oil palm frond. The doped catalysts had high surface areas, high stability, and high-activity basic sites, resulting in high H2 yields compared to supported catalysts. Zn-based catalysts showed higher H2 yields than Ni-based catalysts for both supported and doped catalysts. Mastuli et al. [63] further investigated the structural and catalytic effects of Mg1−xNixO nanomaterial as a catalyst. They synthesized Mg1−xNixO nanomaterial via a self-propagating combustion method in the SCWG of oil palm frond. As the Ni content increased, the cell volume decreased linearly. This increased the specific surface area and improved the basic properties of the catalyst. The Mg0.8Ni0.2O catalyst with the highest Ni content demonstrated the highest gas and H2 yields.
Li et al. [64] demonstrated that the formation of the char layer could be minimized using co-precipitated Ni-Mg-Al catalysts. They varied the Mg-Al molar ratio in the catalyst and investigated its effects in the SCWG of glucose. The catalysts favored H2 production, resulting in high H2 selectivity. Furthermore, Mg inhibited graphitic carbon formation because of its neutralizing action on alumina acidic sites, thus increasing the lifespan of the catalysts. However, the subsequent increase in Mg loading formed the MgNiO2 complex, which limited the activity of Ni metal.
Li et al. [65] also studied the stability and activities of various wet-impregnated Mg-promoted Ni catalysts on Al2O3 and CNT supports in the SCWG of glycerol. The stability studies showed the loss of Al, which resulted in deactivation of the Mg-promoted Ni-Al2O3 catalysts. Both the Ni/α-Al2O3 and Ni/γ-Al2O3 catalysts showed poorer stability and regenerability over repeated use than the Ni/CNT catalyst.
Li and Guo [66] compared the catalytic action of Mg-promoted Ni/Al2O3 catalysts synthesized via the co-precipitation and wet impregnation methods for a variety of feedstocks, such as glycerol, cellulose, glucose, poplar leaf, corncob, phenol, and sawdust. The results showed that the co-precipitated Ni-Mg-Al catalysts were more stable than the wet-impregnated Ni-Mg-Al catalysts. This was due to the growth of the crystal size of the wet-impregnated Ni-Mg-Al catalysts in SCW. Among different feedstocks, the co-precipitated Ni-Mg-Al catalysts were more active for the gasification of water-soluble organics as compared to real lignocellulosic biomasses.
Kang et al. [67] explored and proposed a detailed catalytic mechanism of Ni-Co supported on Mg-Al in the SCWG of lignin (Figure 4). The 2.6%Ni-5.2%Co/2.6%Mg-Al catalyst prepared via the co-precipitation method demonstrated high total gas and H2 yields due to significant improvement in its coke resistance ability. They also concluded that the co-precipitation method was more efficient than the wet-impregnated method. Norouzi et al. [68] showed that the addition of Ru on Fe-Ni/γ-Al2O3 could enhance gas yields while minimizing char formation. Another study by Lu et al. [50] showed that the addition of the Ce promoter on Ni/γ-Al2O3 was also capable of reducing coke and carbon deposition.
Catalysts synthesized in SCW have demonstrated high stability through their ability to reduce sintering. The supercritical water synthesis (SCWS) method for catalyst design provides better control over the size and shape of the nanoparticle without any requirement for organic solvents or precipitants. A few studies on SCWS synthesis of Ni-based catalysts on various supports (e.g., ZrO2, Ce-ZrO2, Al2O3, Mg-Al2O3, CNT and AC) have been reported for the SCWG of biomass [69,70]. SCWS-synthesized Ni/MgO-Al2O3 catalysts demonstrated the highest activities and stability. Despite their increased specific surface areas and pore volumes, SCWS-synthesized Ni/CeO2-ZrO2 catalysts showed no promotional effects when Ce was used. This was because of the low Ni particle dispersion in the Ni/CeO2-ZrO2 catalysts. However, as compared to sol-gel prepared catalysts, which have bigger bulk NiO particles, the SCWS-synthesized catalysts showed high dispersion and stable crystalline structures. After multiple use cycles, the SCWS-synthesized catalysts retained their high dispersion, whereas sol-gel-prepared catalysts experienced growth in size. This allowed the SCWS-prepared catalysts to maintain their high activities over repeated use, as opposed to catalysts prepared using conventional methods that may lose their activity over repeated use. Additionally, SCWS-synthesized catalysts are also synthesized in an environmentally friendly way as they do not require any organic solvents or robust chemical compounds.
Li et al. [71] studied and proposed a catalytic mechanism in the SCWG of dewatered sewage sludge and various model compounds using AlCl3 combined with Ni, KOH, or K2CO3 catalysts and oxidants (e.g., H2O2, K2S2O8, and CaO2). AlCl3-H2O2 demonstrated the highest gas yields, followed by AlCl3-K2S2O8. AlCl3 combined with Ni, KOH, CaO, or K2CO3 catalysts resulted in low H2 yields as compared to AlCl3 alone. However, using K2S2O8 or H2O2 alone decreased the H2 yield. The H2 yield decreased, and gasification efficiency increased with a rise in the addition of oxidants. Interestingly, AlCl3-H2O2 (8:2) showed the highest gas yield, followed by AlCl3-K2S2O8 (8:2) and AlCl3. For the AlCl3-catalyzed SCWG of the model compound, glycerol resulted in the highest H2 yield, followed by guaiacol, glucose, alanine, and humic acid. Al2Cl3-H2O2 increased the H2 yield of humic acid by 17% but decreased the H2 yields of glucose and glycerol by 20% and 12%, respectively, compared to the AlCl3 catalyst. The authors also proposed a catalytic mechanism in the SCWG of dewatered sewage sludge with an AlCl3-H2O2 catalyst. They proposed that AlCl3 promoted the cleavage of the C–C bond with Al3+ ions. The Al3+ ions increased the acidity of SCW by reacting with water and forming Al(OH)3 and H+ ions. Al(OH)3 further underwent dehydration to form AlO(OH), which formed precipitates in water. The H+ and Cl ions enhanced the gasification of intermediate compounds to produce H2, thus increasing the H2 yield. H2O2 further enhanced the gasification of benzene-containing monomers by favoring the steam reforming reaction. In the case of sewage sludge, H+ generated via Al3+ deposition further enhanced the ring-opening activity of H2O2 to promote the decomposition of benzene-containing monomers into small molecules. These small organic molecules were further gasified by the combined catalytic effects of Cl and H+ ions to increase H2 yields.
Although Ni-based catalysts demonstrate improvement in gasification efficiency, they suffer from deactivation mainly because of tar formation and coke deposition [72]. Despite the high activity of Ni/γ-Al2O3-based catalysts, they still suffer from various issues, such as sintering, formation of Ni/Al2O4 complexes, and transformation of the γ-Al2O3 phase to the α-Al2O3 phase. These issues significantly hamper the catalysts’ stability. This is a severe issue for alumina-supported catalysts due to the ready conversion of intermediate products adsorbed on the acidic site into carbon, which deactivates Ni-based catalysts. The addition of alkali promoters can suppress cracking and polymerization reactions. Alkali promoters can also neutralize the acidic sites of alumina supports. Thus, alkali promotors can significantly reduce carbon formation.

3.1.2. Ruthenium-Based Catalysts

Ru-based catalysts with promising metal dispersion are more reactive at low temperatures than Ni-based catalysts [73]. Ru-based catalysts have higher surface areas and distribution than Ni-based catalysts. Therefore, high surface area and more metal distribution can be achieved with relatively low Ru metal loading on the support material. Nguyen et al. [74] also confirmed that Ru-based catalysts show higher catalytic activities per metallic mass than Ni-based catalysts. Additionally, Ru-based catalysts are highly resistant to oxidation and hydrothermal conditions compared to Ni-based catalysts. Ru-based catalysts have higher activities toward hydrogenation and C–C bond cleavage [75]. When compared to other expensive transition metals, Ru-based catalysts exhibit the highest activity and H2 selectivity.
As opposed to Ni-based catalysts, Ru-based catalysts are more susceptible to deactivation by sulfur poisoning [76]. To overcome sulfur sintering, a sacrificial agent with a relatively high affinity towards sulfur can be used to protect Ru from sulfur sintering. Peng et al. [77] used ZnO as a sacrificial agent with Ru/C catalysts to study the SCWG of microalgae (Chlorella vulgaris). ZnO showed high mechanical stability and sulfur adoption performance, which minimized Ru metal sintering. Despite Ru-based catalysts having high surface areas, high dispersion, and high catalytic performance, the relatively low cost of Ni-based catalysts makes them preferable for large-scale industrial applications over Ru-based catalysts.
Kang et al. [29] also observed that Ru/Al2O3 showed the highest metal dispersion compared to Ni-based catalysts. They concluded that 5%Ru/Al2O3 demonstrated a higher H2 yield than the 5%Ni/Al2O3 catalyst in the SCWG of cellulose and lignin. Therefore, for the same metal loading, Ru-based catalysts had higher H2 yields than Ni-based catalysts. Nanda et al. [55] compared Ru/Al2O3 with Ni/Si-Al2O3, K2CO3, and Na2CO3 catalysts in the SCWG of waste cooking oil. The order of H2 yield was Ru/Al2O3 > Ni/Si-Al2O3 > K2CO3 > Na2CO3. The effects of metal loading showed that 5 wt% Ru/Al2O3 resulted in the maximum H2 yield.
The superior catalytic performance of Ru/Al2O3 catalysts has also been reported in the SCWG of glucose and guaiacol [75,78]. In the SCWG of glucose, the Ru/Al2O3 catalyst inhibited the production of furfural and 5-hydroxymethylfurfural while favoring the degradation of intermediates such as phenols, ketones, acids, and arenes [75]. Enhanced gasification of intermediates improved process efficiency and increased total gas and H2 yields while preventing the formation of char. During the SCWG of guaiacol, Ru/Al2O3 catalysts enhanced the conversion of phenol to cyclohexanol by favoring the hydrogenation reaction and the conversion of cyclohexanol to hexanone or hexenol by favoring ring-opening reactions [78]. Hexanone and hexenol can further decompose into small gaseous molecules, including H2. Thus, Ru/Al2O3 improved H2 and total gas yields while minimizing char and tar formation.
Zhang et al. [58] observed the effects of Ni and Ru bimetallic catalysts supported on γ-Al2O3. They recommended the use of Ni and Ru bimetallic catalysts supported on γ-Al2O3 in the SCWG of glucose to achieve high activity and H2 selectivity. Hossain et al. [52] further investigated various bimetallic Ni-Ru/Al2O3-supported aerogel catalysts. Ni-Ru/Al2O3 aerogel catalysts demonstrated 1.3- and 1.6-times higher H2 yields than mesoporous and wet-impregnated synthesized Ni-Ru/Al2O3 catalysts for the same amount of metal loading. The aerogel catalysts showed high and uniform metal particle dispersion with strong interaction between the support and active metal. The high catalytic performance of the aerogel catalysts was due to the supercritical CO2 drying step during aerogel catalyst synthesis, which improved the surface area and reactant diffusivity. A significant decrease in coke formation was also observed with the aerogel catalysts due to their low acidity. This resulted in high stability and activities of the aerogel catalysts.
Tushar et al. [79] confirmed the catalytic effects of Ni and Ru catalysts. They investigated ten different combinations of Ni and Ru catalysts on various supports, such as γ-Al2O3 and ZrO2. Overall, Ni-Ru/γ-Al2O3-ZrO2 demonstrated the maximum H2 yields and high carbon gasification efficiency. Ni-Ru/γ-Al2O3-ZrO2 also demonstrated high stability and activities over repeated use. In another study, dual-component catalysts having equal amounts of Ru/C-Ru/C demonstrated better catalytic activities than single-component catalysts [80].
Yang et al. [81] investigated the kinetics and intermediate products of Ni-Ru/Al2O3 bimetallic catalysts for the SCWG of phenol. They proposed that phenol converted into an enol intermediate via a partial hydrogenation reaction. Furthermore, enol rapidly formed cyclohexanone. This observation was different from the mechanism proposed by Zhu et al. [78] where cyclohexanone was considered as an intermediate product for the formation of cyclohexanol. The kinetic study revealed that phenol was more difficult to gasify than the intermediate compounds. Interestingly, steam reforming of cyclohexanone was not the main contributor to H2 production due to its lower concentration than phenol.

3.1.3. Other Heterogeneous Catalysts

Apart from Ni and Ru, other transition metals such as Pt, Co, and Rh (supported or unsupported) are also used as heterogeneous catalysts in the SCWG process. Karakuş et al. [49] investigated Pt/Al2O3 and Ru/Al2O3 catalysts in the SCWG of 2-propanol. Their results showed that the H2 selectivity of Pt/Al2O3 was relatively higher than that of Ru/Al2O3 due to enhancement of the methanation reaction, which produced CH4 at the expense of H2. Pairojpiriyakul et al. [82] used Co-based catalysts on a variety of supports, such as α-Al2O3, ZrO2, γ-Al2O3, La2O3, and yttria-stabilized zirconia (YSZ), in the SCWG of glycerol. The highest H2 yield was obtained with Co/YSZ. In addition, increasing the Co loading up to 10% improved the gasification efficiency of glycerol and H2 production. However, a further increase in the Co loading decreased both H2 yield and glycerol conversion.
Deactivation, sintering, and poisoning of heterogeneous catalysts by sulfur or coke is still a major challenge. Additionally, heterogeneous catalysts oxidize the elemental sulfur and chlorine in biomass to acids. Retention of these acids in the liquid products of SCWG poses a serious challenge for its disposal and/or recycling. The non-polar nature of SCW dissolves the organic compounds during hydrothermal gasification but the inorganic components, including the active metal (catalyst) and mineral matter (catalyst support), can precipitate and form agglomerates in the reactor if not removed properly. The gradual deposition of these precipitates and agglomerates can corrode the reactor during high-temperature and high-pressure operations [83]. Nevertheless, more advancements are needed to address these challenges to synthesize suitable heterogeneous catalysts with high activity, regenerability, and stability, with resistance to sintering and deactivation.

3.2. Metal Oxide Catalysts

Metal oxide catalysts are rarely used in the SCWG process and very little literature is available on their catalytic performance in SCWG processes. They are generally used as supports to improve the stability and activities of metal-supported catalysts. The most common metal oxides used in SCWG processes are RuO2, ZrO2, and CeO2. Cao et al. [84] compared different metal oxides catalysts such as V2O5, MnO2, Cr2O3, Fe2O3, CuO, Co2O3, ZnO, MoO3, ZrO2, SnO2, CeO2, and WO3 in SCWG of glucose. Among all metal oxide catalysts, Cr2O3, CuO, and WO3 showed high gasification efficiencies compared to Fe2O3, ZnO, and ZrO2. The H2 yields decreased with almost all metal oxide catalysts, except Cr2O3, which improved the H2 yield.
Various co-precipitated binary metal oxide catalysts, such as CeO2-ZrO2, CuO-ZnO, and Fe2O3-Cr2O3, have demonstrated high catalytic performance in SCWG [85,86]. Cao et al. [85] showed that in the SCWG of lignin, the CuO-ZnO catalyst demonstrated high catalytic performance with a high H2 yield and better gasification efficiency, followed by Fe2O3-Cr2O3 and CeO2-ZrO2. However, in the SCWG of cellulose, Fe2O3-Cr2O3 showed a greater H2 yield and high carbon gasification efficiency, followed by CuO-ZnO and CeO2-ZrO2. This was due to the higher oxygen content of cellulose compared to lignin. Thus, oxygen released by metal oxide catalysts had less pronounced effects in the SCWG of cellulose. Additionally, the H2 yield from cellulose was less than that from lignin, which also decreased the reducibility of the reaction medium. The catalytic mechanism of binary metal oxide catalysts showed that CeO2 was the main active component in the CeO2-ZrO2 catalyst [86]. CeO2 distributed on ZrO2 released active oxygen via redox reactions to enhance the SCWG process. ZrO2 also absorbed active H2 and small intermediates to increase contact between the intermediates and CeO2 for improved catalytic performance. In CuO-ZnO, Cu was the main active component, which released oxygen species (Figure 5). ZnO acted as a structural stabilizer, promotor and absorbent for sulfur in the CuO-ZnO supported catalyst.
Onwudili [87] studied the detailed catalytic mechanism of RuO2/γ-Al2O3 in the SCWG of municipal solid waste. RuO2/γ-Al2O3 drastically increased H2, CH4, and CO2 yields while significantly improving gasification efficiency. The high yield of H2 was due to enhancement of the water–gas shift reaction by the catalytic action of RuO2/γ-Al2O3. In addition, the enhancement of methanation of CO or CO2 and hydrogenolysis of C–C hydrocarbons resulted in a high CH4 yield. Improvement in the yields of the reduction product (CH4) and oxidation product (CO2) indicated the involvement of the RuO2/γ-Al2O3 catalyst in Ru(IV) and Ru(0) cyclic redox reactions. Reduction of Ru(IV) into Ru(0) was essential for the SCWG process, whereas oxidation of Ru(0) into Ru(IV) was necessary for the catalytic process. The primary synergetic effects were due to the improvement of the dispersion of RuO2 on γ-Al2O3, which resulted in enhanced carbon gasification efficiency.
Samiee-Zafarghandi et al. [88] compared MnO2/SiO2 and NiO/SiO2 catalysts in the SCWG of microalgae Chlorella. MnO2/SiO2 demonstrated the highest H2 yield (1.1 mmol/g) compared to NiO/SiO2 (0.6 mmol/g) and non-catalytic SCWG (0.2 mmol/g). Therefore, NiO/SiO2 was less active than the supported MnO2/SiO2. Borges et al. [89] investigated the Ni/Fe2O4 catalyst in the SCWG of Eucalyptus wood chips. Ni/Fe2O4 enhanced the H2 yield and decreased the char yield. Further investigation showed that Ni/Fe2O4 favored the water–gas shift and steam reforming reactions, thus increasing H2 yield and decreasing CH4 yield. It also demonstrated good stability and recyclability despite the coke deposit [90].

4. Novel Carbon-Based Catalysts Used in Hydrothermal Gasification

Carbon-based supports can also be used with transition metals in the SCWG of biomass. Their high surface areas along with the renewable and biodegradable nature of activated carbon and other carbon-based supports make them sustainable catalytic materials. Table 3 summarizes some notable studies on the use of carbon-based catalysts for SCWG processes [65,71,91,92,93,94]. Taylor et al. [95] compared Ni/AC and Ru/AC with other catalysts such as KOH, Trona, dolomite, and Borax in the SCWG of wood chips. Both Ni/AC and Ru/AC demonstrated higher H2 yields because of improved water–gas shift compared to other non-carbonaceous catalysts.
Yamaguchi et al. [96] investigated various metals (e.g., Ru, Ni, Pt, Rh and Pd) supported on activated carbon in the SCWG of woody biomass. The Ru/AC catalysts demonstrated the highest gas yields, followed by Rh/AC, Pt/AC, Pd/AC, and Ni/AC. Ru/AC showed the highest activity for lignin gasification. However, it showed an inferior H2 yield, which was due to enhancement of the methanation reaction, which consumed H2. Interestingly, the Pd/AC catalyst demonstrated the highest H2 yield, followed by Ru/AC, Pt/AC, Rh/AC, and Ni/AC. Thus, Pd/AC showed the best H2 yield but poor gas yields, whereas Ni/AC showed the lowest gas and H2 yields. Activated carbon also improved the H2 yield over a wide range of reaction temperatures.
Osada et al. [94] investigated TiO2 and activated carbon as supports for Ru catalysts in the SCWG of lignin, cellulose, and sugarcane bagasse. Ru/AC demonstrated the highest gasification efficiency with near-complete gasification of sugarcane bagasse in 15 min. For the same amount of Ru metal, Ru/AC showed slightly higher activity as compared to Ru/TiO2 catalysts. This was due to the high Ru metal dispersion of 51% in the Ru/AC catalyst as compared to 27% metal dispersion in Ru/TiO2. However, the gas yield and composition of both catalysts were the same when 100% carbon conversion was achieved. This indicated that the equilibrium gas yield and composition did not have any correlation with metal dispersion. For the Ru/AC catalysts, repetitive use increased H2 selectivity but decreased CH4 selectivity due to disintegration of the active sites for the methanation reaction. However, Ru/AC suffered from deactivation since its activity decreased significantly after repetitive use. Therefore, more active and durable AC-based catalysts need to be developed to overcome these challenges. Yamaguchi et al. [96] reported that Ru/γ-Al2O3 demonstrated high gasification activity but low stability as the crystallographic phase of γ-Al2O3 transformed into α-Al2O3.
CNT is another carbon-based support that has a large surface area, high heat conductivity, excellent chemical and physical stability, and a tunable porous structure. Among SCWS-prepared metal-impregnated carbon catalysts, CNT-based catalysts showed higher activities and stability than active carbon and Al2O3 supported catalysts [97]. At reaction conditions of 480 °C, 25 MPa, and 10–50 h, Ni/CNT resulted in the highest H2, CO, CH4, and total gas yields, followed by Ni/AC, Ni/Al2O3, and Ni catalysts. Ni/CNT maintained its high activity even at a longer reaction time of 50 h, whereas Ni/AC and Ni/Al2O3 significantly dropped their activities after 30 h of use. This was primarily due to the leaching of active Ni metal in the Ni/AC and Ni/Al2O3 catalysts.
Rashidi and Tavasoli [98] evaluated the effects of a copper promoter on Ni/CNT catalysts in the SCWG of sugarcane bagasse. Cu-promoted Ni/CNT was found to increase the H2 and total gas yields but decreased the CH4 yield. Thus, Cu-promoted Ni/CNT catalysts overcome the methanation tendency of Ni, which is a major limiting factor of Ni-based catalysts. Azadi et al. [28] reported that Ni-Cu/CNT showed a nearly ten-fold increase in H2 yield and 40 times less CH4 yield with a significant reduction in CO2 yield. Thus, Cu-promoted Ni/CNT catalysts have high H2 selectivity and low CH4 and CO2 selectivities. Li et al. [65] confirmed the high catalytic stability over repeated use of Ni/CNT catalysts in the SCWG of glycerol. de Vlieger et al. [99] also showed the high stability of Pd/CNT catalysts in the SCWG of ethylene glycerol. Pt/CNT exhibited no mass loss with no change in the size and distribution of Pt particles on CNT during SCWG.
Carbonaceous materials such as hydrochar and biochar are also potential materials for the development of catalysts. Safari et al. [91] investigated the performance of catalysts developed from the hydochars of green algae (Cladophora glomerata) and wheat straw in the SCWG of almond shells. The high amounts of alkali and alkaline earth metals in the hydrochar samples enhanced the cracking of biopolymers and favored the water–gas shift reaction, thus increasing the H2 yield. The total gas yield and H2 fraction were selectively improved from 26.7 mmol/g and 41% in the non-catalytic run to 29.2 mmol/g and 58%, respectively, when wheat straw hydrochar was used as the catalyst. The total gas yield and H2 concentration also increased to 31.1 mmol/g and 60%, respectively, when green algae hydrochar was used as the catalyst in the SCWG of almond shells.
Another novel method for catalytic SCWG is the in-situ impregnation of metal nanoparticles in biomass feedstock. This approach can overcome the issue of deactivation encountered by conventional catalysts and help to reduce the cost of catalyst preparation. Nanda et al. [93] carried out the SCWG of pinewood and wheat straw impregnated with Ni-nanoparticles. Ni-impregnated biomasses demonstrated high H2, CO2, and CH4 yields compared to the raw feedstocks. Huang et al. [100] also used in-situ-generated Ni particles using nickel acetate as a precursor for the gasification of glucose in SCW. In-situ-generated Ni catalysts demonstrated superior catalytic performance compared to nickel wire catalysts. They also proved the role of in-situ generated Ni particles from nickel acetate in enabling the catalytic production of H2 during SCWG of glucose.
Kumar and Reddy [101] investigated the SCWG of in-situ Ni-impregnated sugarcane bagasse and lemon peels and compared the results with the Raney nickel catalyst. They used nickel nitrate hexahydrate salt as a precursor for the in-situ generation of nickel nanoparticles. Both Ni-impregnated biomasses demonstrated significantly higher gas yields, H2 yields, and carbon gasification efficiencies than Raney nickel. Ni-impregnated sugarcane bagasse achieved higher carbon gasification efficiency, gas yield, and H2 yield than Ni-lemon peel. Kumar and Reddy [92] also performed the SCWG of banana pseudo-stem using impregnation of Ni, Ru, and Fe metals onto the biomass as the support material. The H2 yields and gasification efficiencies of the metals were in order of Ni > Ru > Fe. The superior performance of Ni to act as an in-situ nanocatalyst is due to its ability to effectively cleave C–H and C–C bonds for improved reforming reactions [102]. However, very little literature is available on the development of in-situ nanocatalysts impregnated onto biomass for proper assessment of their robustness, stability, regeneration, and post-gasification compared to commercially available homogeneous and heterogeneous catalysts. One of the limitations in the design of such novel catalysts can be the presence of lignin and other mineral matter in the biomass [103,104], which can hinder the penetration of catalytic nanoparticles within the cell wall. Therefore, more research is needed for a better understanding of such catalysts and to address these limitations.

5. Conclusions and Perspectives

SCWG is a promising technology for the sustainable production of H2 due to its various advantages over other thermochemical processes. SCWG has shown its potential for converting a wide variety of low-value biomasses into high-value H2-rich gas products. This can serve as a green alternative to the steam methane reforming process due to the renewable and clean nature of biomass sources compared to fossil fuels. However, SCWG requires high energy input to achieve supercritical conditions. Nonetheless, catalysts are used to achieve high gas yields and process efficiencies even at near-critical conditions.
Various homogeneous and heterogeneous catalysts have been studied to achieve high H2 yields at low temperatures in SCWG processes. Although homogeneous catalysts are suitable compared to heterogeneous catalysts, they suffer from recovery issues. This also increases the cost of the process and hinders its use in large-scale industrial applications. On the other hand, heterogeneous catalysts are relatively easier to recover, but they can suffer from deactivation. Deactivation of heterogeneous catalysts can occur for various reasons, such as fouling, poising, sintering, and char formation. Transition metal catalysts (e.g., Ni, Cu, Co, and Ru) have demonstrated excellent performance in enhancing SCWG reactions. Ru- and Ni-based catalysts are the most widely used catalysts owing to their superior performance in SCWG processes, especially in water–gas shift, hydrogenation, and methanation reactions. Novel catalysts such as activated carbon, char, CNT, and lignocellulosic biomass impregnated with catalytic nanoparticles have demonstrated promising potential to achieve comparable catalytic performance and renewability in SCWG reactions.
It cannot be denied that SCWG is an innovative and viable technology for producing combustible gases with higher selectivity to individual gas components using catalysts. However, a detailed study of the economic viability and technical feasibility of these catalysts is needed. New developments in the field of catalysts can facilitate the commercialization of SCWG technology. Extensive research strategies are required to tackle the unique challenges faced by SCWG technology that prevents its scalability and commercialization. Some of the common challenges are associated with reactor corrosion, plugging due to salt and mineral precipitation, the requirement of special reactor set-up resistant to high temperatures, high pressures, and corrosion, coking of catalyst supports, as well as catalyst poisoning, sintering, and deactivation. The techno-economic, environmental, and lifecycle viability of SCWG technology on a commercial scale is also contingent on the efficient conversion of feedstocks, catalyst recovery, regeneration and reuse, effective separation of gas, liquid, and solid products, as well as upgrading and applications of main products and co-products. Nonetheless, SCWG remains an appealing technology with many benefits in the use of water as a source of aqueous reaction media to valorize complex feedstocks and pollutants under environmentally benign conditions while addressing the issues of waste management and clean energy recovery.

Author Contributions

Conceptualization, K.K., P.B. and S.N.; validation, K.K., P.B. and S.N.; investigation, K.K., P.B. and S.N.; resources, A.K.D.; data curation, K.K., P.B. and S.N.; writing—original draft preparation, K.K., P.B. and S.N.; writing—review and editing, K.K., P.B., S.N. and A.K.D.; visualization, K.K., P.B. and S.N.; supervision, A.K.D.; project administration, A.K.D.; funding acquisition, A.K.D. All authors have read and agreed to the published version of the manuscript.

Funding

The authors would like to thank the Natural Sciences and Engineering Research Council of Canada (NSERC) and the Canada Research Chairs (CRC) program for funding this work.

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

Not applicable.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Nanda, S.; Reddy, S.N.; Mitra, S.K.; Kozinski, J.A. The progressive routes for carbon capture and sequestration. Energy Sci. Eng. 2016, 4, 99–122. [Google Scholar] [CrossRef] [Green Version]
  2. Our World in Data. CO2 Emissions by Fuel or Industry, World. Available online: https://ourworldindata.org/grapher/co2-emissions-by-fuel-line?facet=none (accessed on 11 May 2023).
  3. Jha, S.; Nanda, S.; Acharya, B.; Dalai, A.K. A review of thermochemical conversion of waste biomass to biofuels. Energies 2022, 15, 6352. [Google Scholar] [CrossRef]
  4. Okolie, J.A.; Nanda, S.; Dalai, A.K.; Kozinski, J.A. Chemistry and specialty industrial applications of lignocellulosic biomass. Waste Biomass Valor. 2021, 12, 2145–2169. [Google Scholar] [CrossRef]
  5. Okolie, J.A.; Patra, B.R.; Mukherjee, A.; Nanda, S.; Dalai, A.K.; Kozinski, J.A. Futuristic applications of hydrogen in energy, biorefining, aerospace, pharmaceuticals and metallurgy. Int. J. Hydrogen Energy 2021, 46, 8885–8905. [Google Scholar] [CrossRef]
  6. Fan, L.; Tu, Z.; Chan, S.H. Recent development of hydrogen and fuel cell technologies: A review. Energy Rep. 2021, 7, 8421–8446. [Google Scholar] [CrossRef]
  7. Nanda, S.; Rana, R.; Zheng, Y.; Kozinski, J.A.; Dalai, A.K. Insights on pathways for hydrogen generation from ethanol. Sustain. Energy Fuels 2017, 1, 1232–1245. [Google Scholar] [CrossRef]
  8. Zang, G.; Sun, P.; Elgowainy, A.; Wang, M. Technoeconomic and life cycle analysis of synthetic methanol production from hydrogen and industrial byproduct CO2. Environ. Sci. Technol. 2021, 55, 5248–5257. [Google Scholar] [CrossRef]
  9. Kim, S.; Kwon, E.E.; Kim, Y.T.; Jung, S.; Kim, H.J.; Hubere, G.W.; Lee, J. Recent advances in hydrodeoxygenation of biomass-derived oxygenates over heterogeneous catalysts. Green Chem. 2019, 21, 3715–3743. [Google Scholar] [CrossRef]
  10. Bello, S.S.; Wang, C.; Zhang, M.; Gao, H.; Han, Z.; Shi, L.; Su, F.; Xu, G. A review on the reaction mechanism of hydrodesulfurization and hydrodenitrogenation in heavy oil upgrading. Energy Fuels 2021, 35, 10998–11016. [Google Scholar] [CrossRef]
  11. Nguyen, T.H.; Nguyen, Q.A.; Cao, A.N.T.; Ernest, T.; Nguyen, T.B.; Pham, P.H.T.; Nguyen, T.M. Hydrodemetallization of heavy oil: Recent progress, challenge, and future prospects. J. Petrol. Sci. Eng. 2022, 216, 110762. [Google Scholar] [CrossRef]
  12. National Grid. The Hydrogen Colour Spectrum. Available online: https://www.nationalgrid.com/stories/energy-explained/hydrogen-colour-spectrum (accessed on 11 May 2023).
  13. Finke, C.E.; Leandri, H.F.; Karumb, E.T.; Zheng, D.; Hoffmann, M.R.; Fromer, N.A. Economically advantageous pathways for reducing greenhouse gas emissions from industrial hydrogen under common, current economic conditions. Energy Environ. Sci. 2021, 14, 1517–1529. [Google Scholar] [CrossRef]
  14. Rapier, R. Estimating the Carbon Footprint of Hydrogen Production. Forbes. Available online: https://www.forbes.com/sites/rrapier/2020/06/06/estimating-the-carbon-footprint-of-hydrogen-production/?sh=5cfba0b424bd (accessed on 11 May 2023).
  15. Sarangi, P.K.; Nanda, S. Biohydrogen production through dark fermentation. Chem. Eng. Technol. 2020, 43, 601–612. [Google Scholar] [CrossRef]
  16. Reddy, S.N.; Nanda, S.; Dalai, A.K.; Kozinski, J.A. Supercritical water gasification of biomass for hydrogen production. Int. J. Hydrogen Energy 2014, 39, 6912–6926. [Google Scholar] [CrossRef]
  17. Okolie, J.A.; Nanda, S.; Dalai, A.K.; Berruti, F.; Kozinski, J.A. A review on subcritical and supercritical water gasification of biogenic, polymeric and petroleum wastes to hydrogen-rich synthesis gas. Renew. Sustain. Energy Rev. 2020, 119, 109546. [Google Scholar] [CrossRef]
  18. Okolie, J.A.; Rana, R.; Nanda, S.; Dalai, A.K.; Kozinski, J.A. Supercritical water gasification of biomass: A state-of-the-art review of process parameters, reaction mechanisms and catalysis. Sustain. Energy Fuels 2019, 3, 578–598. [Google Scholar] [CrossRef]
  19. Schienbein, P.; Marx, D. Assessing the properties of supercritical water in terms of structural dynamics and electronic polarization effects. Phys. Chem. Chem. Phys. 2020, 22, 10462–10479. [Google Scholar] [CrossRef]
  20. Correa, C.R.; Kruse, A. Supercritical water gasification of biomass for hydrogen production–Review. J. Supercrit. Fluids 2018, 133, 573–590. [Google Scholar] [CrossRef]
  21. Masoumi, S.; Borugadda, V.B.; Nanda, S.; Dalai, A.K. Hydrochar: A review on its production technologies and applications. Catalysts 2021, 11, 939. [Google Scholar] [CrossRef]
  22. Nanda, S.; Dalai, A.K.; Berruti, F.; Kozinski, J.A. Biochar as an exceptional bioresource for energy, agronomy, carbon sequestration, activated carbon and specialty materials. Waste Biomass Valor. 2016, 7, 201–235. [Google Scholar] [CrossRef]
  23. Patra, B.R.; Mukherjee, A.; Nanda, S.; Dalai, A.K. Biochar production, activation and adsorptive applications: A review. Environ. Chem. Lett. 2021, 19, 2237–2259. [Google Scholar] [CrossRef]
  24. Kang, K.; Nanda, S.; Hu, Y. Current trends in biochar application for catalytic conversion of biomass to biofuels. Catal. Today 2022, 404, 3–18. [Google Scholar] [CrossRef]
  25. Khorasani, R.; Khodaparasti, M.S.; Tavakoli, O. Hydrogen production from dairy wastewater using catalytic supercritical water gasification: Mechanism and reaction pathway. Int. J. Hydrogen Energy 2021, 46, 22368–22384. [Google Scholar] [CrossRef]
  26. Ghavami, N.; Özdenkçi, K.; Salierno, G.; Björklund-Sänkiaho, M.; De Blasio, C. Analysis of operational issues in hydrothermal liquefaction and supercritical water gasification processes: A review. Biomass Convers. Bioref. 2021. [Google Scholar] [CrossRef]
  27. Azadi, P.; Afif, E.; Azadi, F.; Farnood, R. Screening of nickel catalysts for selective hydrogen production using supercritical water gasification of glucose. Green Chem. 2012, 14, 1766–1777. [Google Scholar] [CrossRef]
  28. Azadi, P.; Afif, E.; Foroughi, H.; Dai, T.; Azadi, F.; Farnood, R. Catalytic reforming of activated sludge model compounds in supercritical water using nickel and ruthenium catalysts. Appl. Catal. B Environ. 2013, 134–135, 265–273. [Google Scholar] [CrossRef]
  29. Kang, K.; Azargohar, R.; Dalai, A.K.; Wang, H. Hydrogen production from lignin, cellulose and waste biomass via supercritical water gasification: Catalyst activity and process optimization study. Energy Convers. Manag. 2016, 117, 528–537. [Google Scholar] [CrossRef]
  30. Madenoğlu, T.G.; Sağlam, M.; Yüksel, M.; Ballice, L. Hydrothermal gasification of biomass model compounds (cellulose and lignin alkali) and model mixtures. J. Supercrit. Fluids 2016, 115, 79–85. [Google Scholar] [CrossRef]
  31. García-Jarana, M.B.; Portela, J.R.; Sánchez-Oneto, J.; de la Ossa, E.J.M.; Al-Duri, B. Analysis of the supercritical water gasification of cellulose in a continuous system using short residence times. Appl. Sci. 2020, 10, 5185. [Google Scholar] [CrossRef]
  32. Sınaǧ, A.; Kruse, A.; Rathert, J. Influence of the heating rate and the type of catalyst on the formation of key intermediates and on the generation of gases during hydropyrolysis of glucose in supercritical water in a batch reactor. Ind. Eng. Chem. Res. 2004, 43, 502–508. [Google Scholar] [CrossRef]
  33. Rönnlund, I.; Myréen, L.; Lundqvist, K.; Ahlbeck, J.; Westerlund, T. Waste to energy by industrially integrated supercritical water gasification–Effects of alkali salts in residual by-products from the pulp and paper industry. Energy 2011, 36, 2151–2163. [Google Scholar] [CrossRef]
  34. Feng, H.; Zhou, Z.; Hantoko, D.; Zhong, L.; Rahim, D.A.; Fang, W.; Yan, M. Effect of alkali additives on desulfurization of syngas during supercritical water gasification of sewage sludge. Waste Manag. 2021, 131, 394–402. [Google Scholar] [CrossRef] [PubMed]
  35. Nanda, S.; Dalai, A.K.; Kozinski, J.A. Supercritical water gasification of timothy grass as an energy crop in the presence of alkali carbonate and hydroxide catalysts. Biomass Bioenergy 2016, 95, 378–387. [Google Scholar] [CrossRef]
  36. Kruse, A.; Bernolle, P.; Dahmen, N.; Dinjus, E.; Maniam, P. Hydrothermal gasification of biomass: Consecutive reactions to long-living intermediates. Energy Environ. Sci. 2010, 3, 136–143. [Google Scholar] [CrossRef]
  37. Su, W.; Zhao, M.; Xing, Y.; Ma, H.; Liu, P.; Li, X.; Zhang, H.; Wu, Y.; Xia, C. Supercritical water gasification of hyperaccumulators for hydrogen production and heavy metal immobilization with alkali metal catalysts. Environ. Res. 2022, 214, 114093. [Google Scholar] [CrossRef]
  38. Watanabe, M.; Sue, K.; Adschiri, T.; Inomata, H.; Smith, R.L.; Arai, K. Control of methanol oxidation by ionic behavior in supercritical water. Chem. Commun. 2001, 1, 2270–2271. [Google Scholar] [CrossRef] [PubMed]
  39. Ge, Z.; Jin, H.; Guo, L. Hydrogen production by catalytic gasification of coal in supercritical water with alkaline catalysts: Explore the way to complete gasification of coal. Int. J. Hydrogen Energy 2014, 39, 19583–19592. [Google Scholar] [CrossRef]
  40. Sınaǧ, A.; Kruse, A.; Maniam, P. Hydrothermal conversion of biomass and different model compounds. J. Supercrit. Fluids 2012, 71, 80–85. [Google Scholar] [CrossRef]
  41. Nanda, S.; Reddy, S.N.; Hunter, H.N.; Dalai, A.K.; Kozinski, J.A. Supercritical water gasification of fructose as a model compound for waste fruits and vegetables. J. Supercrit. Fluids 2015, 104, 112–121. [Google Scholar] [CrossRef]
  42. Yanik, J.; Ebale, S.; Kruse, A.; Saglam, M.; Yüksel, M. Biomass gasification in supercritical water: II. Effect of catalyst. Int. J. Hydrogen Energy 2008, 33, 4520–4526. [Google Scholar] [CrossRef]
  43. Ferreira-Pinto, L.; Feirhrmann, A.C.; Corazza, M.L.; Fernandes-Machado, N.R.C.; dos Reis Coimbra, J.S.; Saldaña, M.D.A.; Cardozo-Filho, L. Hydrogen production and TOC reduction from gasification of lactose by supercritical water. Int. J. Hydrogen Energy 2015, 40, 12162–12168. [Google Scholar] [CrossRef]
  44. Zhong, J.; Zhu, W.; Wang, C.; Mu, B.; Lin, N.; Chen, S.; Li, Z. Transformation mechanism of polycyclic aromatic hydrocarbons and hydrogen production during the gasification of coking sludge in supercritical water. Chemosphere 2022, 300, 134467. [Google Scholar] [CrossRef]
  45. Leite, M.J.L.; Marques, I.R.; Proner, M.C.; Araújo, P.H.H.; Ambrosi, A.; Luccio, M.D. Catalytically active membranes for esterification: A review. Chin. J. Chem. Eng. 2023, 53, 142–154. [Google Scholar] [CrossRef]
  46. Dreimann, J.; Lutze, P.; Zagajewski, M.; Behr, A.; Górak, A.; Vorholt, A.J. Highly integrated reactor–separator systems for the recycling of homogeneous catalysts. Chem. Eng. Process. Process Intens. 2016, 99, 124–131. [Google Scholar] [CrossRef]
  47. Fadhel, A.J.; Pollet, P.; Liotta, C.L.; Eckert, C.A. Combining the benefits of homogeneous and heterogeneous catalysis with tunable solvents and nearcritical water. Molecules 2010, 15, 8400–8424. [Google Scholar] [CrossRef] [Green Version]
  48. Zaera, F. Designing sites in heterogeneous catalysis: Are we reaching selectivities competitive with those of homogeneous catalysts? Chem. Rev. 2022, 122, 8594–8757. [Google Scholar] [CrossRef]
  49. Karakuş, Y.; Aynacı, F.; Kıpçak, E.; Akgün, M. Hydrogen production from 2-propanol over Pt/Al2O3 and Ru/Al2O3 catalysts in supercritical water. Int. J. Hydrogen Energy 2013, 38, 7298–7306. [Google Scholar] [CrossRef]
  50. Lu, Y.; Zhu, Y.; Li, S.; Zhang, X.; Guo, L. Behavior of nickel catalysts in supercritical water gasification of glucose: Influence of support. Biomass Bioenergy 2014, 67, 125–136. [Google Scholar] [CrossRef]
  51. Lu, Y.; Jin, H.; Zhang, R. Evaluation of stability and catalytic activity of Ni catalysts for hydrogen production by biomass gasification in supercritical water. Carbon Resour. Convers. 2019, 2, 95–101. [Google Scholar] [CrossRef]
  52. Hossain, M.Z.; Chowdhury, M.B.I.; Jhawar, A.K.; Charpentier, P.A. Supercritical water gasification of glucose using bimetallic aerogel Ru-Ni-Al2O3 catalyst for H2 production. Biomass Bioenergy 2017, 107, 39–51. [Google Scholar] [CrossRef]
  53. Onwudili, J.A.; Williams, P.T. Catalytic Supercritical Water Gasification of Plastics with Supported RuO2: A Potential Solution to Hydrocarbons–Water Pollution Problem. Process Saf. Environ. Prot. 2016, 102, 140–149. [Google Scholar] [CrossRef]
  54. Okolie, J.A.; Mukherjee, A.; Nanda, S.; Dalai, A.K.; Kozinski, J.A. Catalytic supercritical water gasification of soybean straw: Effects of catalyst supports and promoters. Ind. Eng. Chem. Res. 2021, 60, 5770–5782. [Google Scholar] [CrossRef]
  55. Nanda, S.; Rana, R.; Hunter, H.N.; Fang, Z.; Dalai, A.K.; Kozinski, J.A. Hydrothermal catalytic processing of waste cooking oil for hydrogen-rich syngas production. Chem. Eng. Sci. 2019, 195, 935–945. [Google Scholar] [CrossRef]
  56. Shen, L.; Xu, J.; Zhu, M.; Han, Y.F. Essential role of the support for nickel-based CO2 methanation catalysts. ACS Catal. 2020, 10, 14581–14591. [Google Scholar] [CrossRef]
  57. Furusawa, T.; Sato, T.; Saito, M.; Ishiyama, Y.; Sato, M.; Itoh, N.; Suzuki, N. The evaluation of the stability of Ni/MgO catalysts for the gasification of lignin in supercritical water. Appl. Catal. A Gen. 2007, 327, 300–310. [Google Scholar] [CrossRef]
  58. Zhang, L.; Champagne, P.; Xu, C.C. Screening of supported transition metal catalysts for hydrogen production from glucose via catalytic supercritical water gasification. Int. J. Hydrogen Energy 2011, 36, 9591–9601. [Google Scholar] [CrossRef]
  59. Adamu, S.; Razzak, S.A.; Hossain, M.M. Fluidizable Ni/Ce-Meso-Al2O3 for gasification of glucose: Effect of catalyst reduction on hydrogen selectivity. J. Ind. Eng. Chem. 2018, 64, 467–477. [Google Scholar] [CrossRef]
  60. Su, H.; Kanchanatip, E.; Wang, D.; Zhang, H.; Antoni; Mubeen, I.; Huang, Z.; Yan, M. Catalytic gasification of food waste in supercritical water over la promoted Ni/Al2O3 catalysts for enhancing H2 production. Int. J. Hydrogen Energy 2020, 45, 553–564. [Google Scholar] [CrossRef]
  61. Chowdhury, M.B.I.; Hossain, M.Z.; Mazumder, J.; Jhawar, A.K.; Charpentier, P.A. La-based catalysts to enhance hydrogen production during supercritical water gasification of glucose. Fuel 2018, 217, 166–174. [Google Scholar] [CrossRef]
  62. Mastuli, M.S.; Kamarulzaman, N.; Kasim, M.F.; Zainal, Z.; Matsumura, Y.; Taufiq-Yap, Y.H. Comparative study between supported and doped MgO catalysts in supercritical water gasification for hydrogen production. Int. J. Hydrogen Energy 2019, 44, 3690–3701. [Google Scholar] [CrossRef]
  63. Mastuli, M.S.; Kasim, M.F.; Mahat, A.M.; Asikin-Mijan, N.; Sivasangar, S.; Taufiq-Yap, Y.H. Structural and catalytic studies of Mg1-xNixO nanomaterials for gasification of biomass in supercritical water for H2-rich syngas production. Int. J. Hydrogen Energy 2020, 45, 33218–33234. [Google Scholar] [CrossRef]
  64. Li, S.; Guo, L.; Zhu, C.; Lu, Y. Co-precipitated Ni–Mg–Al catalysts for hydrogen production by supercritical water gasification of glucose. Int. J. Hydrogen Energy 2013, 38, 9688–9700. [Google Scholar] [CrossRef]
  65. Li, S.; Savage, P.E.; Guo, L. Stability and activity maintenance of Al2O3- and carbon nanotube-supported Ni catalysts during continuous gasification of glycerol in supercritical water. J. Supercrit. Fluids 2018, 135, 188–197. [Google Scholar] [CrossRef]
  66. Li, S.; Guo, L. Stability and activity of a co-precipitated Mg promoted Ni/Al2O3 catalyst for supercritical water gasification of biomass. Int. J. Hydrogen Energy 2019, 44, 15842–15852. [Google Scholar] [CrossRef]
  67. Kang, K.; Azargohar, R.; Dalai, A.K.; Wang, H. Hydrogen generation via supercritical water gasification of lignin using Ni-Co/Mg-Al catalysts. Int. J. Energy Res. 2017, 41, 1835–1846. [Google Scholar] [CrossRef]
  68. Norouzi, O.; Safari, F.; Jafarian, S.; Tavasoli, A.; Karimi, A. Hydrothermal gasification performance of Enteromorpha intestinalis as an algal biomass for hydrogen-rich gas production using Ru promoted Fe–Ni/γ-Al2O3 nanocatalysts. Energy Convers. Manag. 2017, 141, 63–71. [Google Scholar] [CrossRef]
  69. Zhu, B.; Li, S.; Wang, W.; Zhang, H. Supercritical water synthesized Ni/ZrO2 catalyst for hydrogen production from supercritical water gasification of glycerol. Int. J. Hydrogen Energy 2019, 44, 30917–30926. [Google Scholar] [CrossRef]
  70. Li, S.; Zhu, B.; Wang, W.; Zhang, H.; Li, Q. Efficient and stable supercritical-water-synthesized Ni-based catalysts for supercritical water gasification. J. Supercrit. Fluids 2020, 160, 104790. [Google Scholar] [CrossRef]
  71. Li, Z.; Gong, M.; Wang, M.; Feng, A.; Wang, L.; Ma, P.; Yuan, S. Influence of AlCl3 and oxidant catalysts on hydrogen production from the supercritical water gasification of dewatered sewage sludge and model compounds. Int. J. Hydrogen Energy 2021, 46, 31262–31274. [Google Scholar] [CrossRef]
  72. Pattnaik, F.; Patra, B.R.; Okolie, J.A.; Nanda, S.; Dalai, A.K.; Naik, S. A review of thermocatalytic conversion of biogenic wastes into crude biofuels and biochemical precursors. Fuel 2022, 320, 123857. [Google Scholar] [CrossRef]
  73. Al-Doghachi, F.A.J.; Islam, A.; Zainal, Z.; Saiman, M.I.; Embong, Z.; Taufiq-Yap, Y.H. High coke-resistance Pt/Mg1-xNixO catalyst for dry reforming of methane. PLoS ONE 2016, 11, e0146862. [Google Scholar] [CrossRef] [Green Version]
  74. Nguyen, H.T.; Yoda, E.; Komiyama, M. Catalytic supercritical water gasification of proteinaceous biomass: Catalyst performances in gasification of ethanol fermentation stillage with batch and flow reactors. Chem. Eng. Sci. 2014, 109, 197–203. [Google Scholar] [CrossRef]
  75. Zhu, C.; Guo, L.; Jin, H.; Huang, J.; Li, S.; Lian, X. Effects of reaction time and catalyst on gasification of glucose in supercritical water: Detailed reaction pathway and mechanisms. Int. J. Hydrogen Energy 2016, 41, 6630–6639. [Google Scholar] [CrossRef]
  76. Hunston, C.; Baudouin, D.; Tarik, M.; Kröcher, O.; Vogel, F. Investigating active phase loss from supported ruthenium catalysts during supercritical water gasification. Catal. Sci. Technol. 2021, 11, 7431–7444. [Google Scholar] [CrossRef] [PubMed]
  77. Peng, G.; Ludwig, C.; Vogel, F. Catalytic supercritical water gasification: Interaction of sulfur with ZnO and the ruthenium catalyst. Appl. Catal. B Environ. 2017, 202, 262–268. [Google Scholar] [CrossRef]
  78. Zhu, C.; Guo, L.; Jin, H.; Ou, Z.; Wei, W.; Huang, J. Gasification of guaiacol in supercritical water: Detailed reaction pathway and mechanisms. Int. J. Hydrogen Energy 2018, 43, 14078–14086. [Google Scholar] [CrossRef]
  79. Tushar, M.S.H.K.; Dutta, A.; Xu, C.C. Catalytic supercritical gasification of biocrude from hydrothermal liquefaction of cattle manure. Appl. Catal. B Environ. 2016, 189, 119–132. [Google Scholar] [CrossRef]
  80. Duan, P.G.; Yang, S.K.; Xu, Y.P.; Wang, F.; Zhao, D.; Weng, Y.J.; Shi, X.L. Integration of hydrothermal liquefaction and supercritical water gasification for improvement of energy recovery from algal biomass. Energy 2018, 155, 734–745. [Google Scholar] [CrossRef]
  81. Yang, M.; Zhang, J.; Guo, Y. Supercritical water gasification of phenol over Ni-Ru bimetallic catalyst: Intermediates and kinetics. J. Supercrit. Fluids 2020, 160, 104810. [Google Scholar] [CrossRef]
  82. Pairojpiriyakul, T.; Croiset, E.; Kiatkittipong, W.; Kiatkittipong, K.; Arpornwichanop, A.; Assabumrungrat, S. Hydrogen production from catalytic supercritical water reforming of glycerol with cobalt-based catalysts. Int. J. Hydrogen Energy 2013, 38, 4368–4379. [Google Scholar] [CrossRef]
  83. Lee, C.S.; Conradie, A.V.; Lester, E. Review of supercritical water gasification with lignocellulosic real biomass as the feedstocks: Process parameters, biomass composition, catalyst development, reactor design and its challenges. Chem. Eng. J. 2021, 415, 128837. [Google Scholar] [CrossRef]
  84. Cao, C.; Zhang, Y.; Cao, W.; Jin, H.; Guo, L.; Huo, Z. Transition metal oxides as catalysts for hydrogen production from supercritical water gasification of glucose. Catal. Lett. 2017, 147, 828–836. [Google Scholar] [CrossRef]
  85. Cao, C.; Xie, Y.; Chen, Y.; Lu, J.; Shi, J.; Jin, H.; Wang, S.; Zhang, L. Hydrogen production from supercritical water gasification of lignin and cellulose with coprecipitated CuO–ZnO and Fe2O3–Cr2O3. Ind. Eng. Chem. Res. 2021, 60, 7033–7042. [Google Scholar] [CrossRef]
  86. Cao, C.; Xie, Y.; Li, L.; Wei, W.; Jin, H.; Wang, S.; Li, W. Supercritical water gasification of lignin and cellulose catalyzed with co-precipitated CeO2-ZrO2. Energy Fuels 2021, 35, 6030–6039. [Google Scholar] [CrossRef]
  87. Onwudili, J.A. Supercritical water gasification of RDF and its components over RuO2/γ-Al2O3 catalyst: New insights into RuO2 catalytic reaction mechanisms. Fuel 2016, 181, 157–169. [Google Scholar] [CrossRef] [Green Version]
  88. Samiee-Zafarghandi, R.; Karimi-Sabet, J.; Abdoli, M.A.; Karbassi, A. Supercritical water gasification of microalga Chlorella PTCC 6010 for hydrogen production: Box-Behnken optimization and evaluating catalytic effect of MnO2/SiO2 and NiO/SiO2. Renew. Energy 2018, 126, 189–201. [Google Scholar] [CrossRef]
  89. Borges, A.C.P.; Onwudili, J.A.; Andrade, H.M.C.; Alves, C.T.; Ingram, A.; Vieira de Melo, S.A.B.; Torres, E.A. Catalytic supercritical water gasification of eucalyptus wood chips in a batch reactor. Fuel 2019, 255, 115804. [Google Scholar] [CrossRef]
  90. Borges, A.C.P.; Onwudili, J.A.; Andrade, H.; Alves, C.; Ingram, A.; Vieira de Melo, S.; Torres, E. Catalytic properties and recycling of NiFe2O4 catalyst for hydrogen production by supercritical water gasification of eucalyptus wood chips. Energies 2020, 13, 4553. [Google Scholar] [CrossRef]
  91. Safari, F.; Javani, N.; Yumurtaci, Z. Hydrogen production via supercritical water gasification of almond shell over algal and agricultural hydrochars as catalysts. Int. J. Hydrogen Energy 2018, 43, 1071–1080. [Google Scholar] [CrossRef]
  92. Kumar, A.; Reddy, S.N. Subcritical and supercritical water in-situ gasification of metal (Ni/Ru/Fe) impregnated banana pseudo-stem for hydrogen rich fuel gas mixture. Int. J. Hydrogen Energy 2020, 45, 18348–18362. [Google Scholar] [CrossRef]
  93. Nanda, S.; Reddy, S.N.; Dalai, A.K.; Kozinski, J.A. Subcritical and supercritical water gasification of lignocellulosic biomass impregnated with nickel nanocatalyst for hydrogen production. Int. J. Hydrogen Energy 2016, 41, 4907–4921. [Google Scholar] [CrossRef]
  94. Osada, M.; Yamaguchi, A.; Hiyoshi, N.; Sato, O.; Shirai, M. Gasification of sugarcane bagasse over supported ruthenium catalysts in supercritical water. Energy Fuels 2012, 26, 3179–3186. [Google Scholar] [CrossRef]
  95. Taylor, A.D.; DiLeo, G.J.; Sun, K. Hydrogen production and performance of nickel based catalysts synthesized using supercritical fluids for the gasification of biomass. Appl. Catal. B Environ. 2009, 93, 126–133. [Google Scholar] [CrossRef]
  96. Yamaguchi, A.; Hiyoshi, N.; Sato, O.; Bando, K.K.; Osada, M.; Shirai, M. Hydrogen production from woody biomass over supported metal catalysts in supercritical water. Catal. Today 2009, 146, 192–195. [Google Scholar] [CrossRef]
  97. Wang, Y.; Zhu, Y.; Liu, Z.; Wang, L.; Xu, D.; Fang, C.; Wang, S. Catalytic performances of Ni-based catalysts on supercritical water gasification of phenol solution and coal-gasification wastewater. Int. J. Hydrogen Energy 2019, 44, 3470–3480. [Google Scholar] [CrossRef]
  98. Rashidi, M.; Tavasoli, A. Hydrogen rich gas production via supercritical water gasification of sugarcane bagasse using unpromoted and copper promoted Ni/CNT nanocatalysts. J. Supercrit. Fluids 2015, 98, 111–118. [Google Scholar] [CrossRef]
  99. de Vlieger, D.J.M.; Thakur, D.B.; Lefferts, L.; Seshan, K. Carbon nanotubes: A promising catalyst support material for supercritical water gasification of biomass waste. ChemCatChem 2012, 4, 2068–2074. [Google Scholar] [CrossRef]
  100. Huang, J.; Zhu, C.; Lian, X.; Feng, H.; Sun, J.; Wang, L.; Jin, H. Catalytic supercritical water gasification of glucose with in-situ generated nickel nanoparticles for hydrogen production. Int. J. Hydrogen Energy 2019, 44, 21020–21029. [Google Scholar] [CrossRef]
  101. Kumar, A.; Reddy, S.N. In situ sub- and supercritical water gasification of nano-nickel (Ni2+) impregnated biomass for H2 production. Ind. Eng. Chem. Res. 2019, 58, 4780–4793. [Google Scholar] [CrossRef]
  102. Shen, Y.; Zhao, P.; Shao, Q.; Takahashi, F.; Yoshikawa, K. In situ catalytic conversion of tar using rice husk char/ash supported nickel–iron catalysts for biomass pyrolytic gasification combined with the mixing-simulation in fluidized-bed gasifier. Appl. Energy 2015, 160, 808–819. [Google Scholar] [CrossRef]
  103. Fougere, D.; Nanda, S.; Clarke, K.; Kozinski, J.A.; Li, K. Effect of acidic pretreatment on the chemistry and distribution of lignin in aspen wood and wheat straw substrates. Biomass Bioenergy 2016, 91, 56–68. [Google Scholar] [CrossRef]
  104. Sun, J.; Xu, L.; Dong, G.H.; Nanda, S.; Li, H.; Fang, Z.; Kozinski, J.A.; Dalai, A.K. Subcritical water gasification of lignocellulosic wastes for hydrogen production with Co modified Ni/Al2O3 catalyst. J. Supercrit. Fluids 2020, 162, 104863. [Google Scholar] [CrossRef]
Figure 1. Different color shades of the hydrogen spectrum.
Figure 1. Different color shades of the hydrogen spectrum.
Molecules 28 05137 g001
Figure 2. Catalytic mechanism of potassium in SCWG of biomass (adapted with permission from Ge et al. [39]).
Figure 2. Catalytic mechanism of potassium in SCWG of biomass (adapted with permission from Ge et al. [39]).
Molecules 28 05137 g002
Figure 3. Catalytic mechanism of Ni/Ce-Al2O3 in SCWG of glucose (adapted with permission from Adamu et al. [59]).
Figure 3. Catalytic mechanism of Ni/Ce-Al2O3 in SCWG of glucose (adapted with permission from Adamu et al. [59]).
Molecules 28 05137 g003
Figure 4. Catalytic mechanism of Ni-Co/Mg-Al in SCWG of lignin (adapted with permission from Kang et al. [67]).
Figure 4. Catalytic mechanism of Ni-Co/Mg-Al in SCWG of lignin (adapted with permission from Kang et al. [67]).
Molecules 28 05137 g004
Figure 5. Catalytic mechanism of CuO-ZnO in SCWG of biomass (adapted with permission from Cao et al. [85]).
Figure 5. Catalytic mechanism of CuO-ZnO in SCWG of biomass (adapted with permission from Cao et al. [85]).
Molecules 28 05137 g005
Table 1. Notable studies on SCWG of waste biomass assisted by homogeneous catalysts.
Table 1. Notable studies on SCWG of waste biomass assisted by homogeneous catalysts.
FeedstockCatalystOperating ConditionsMain FindingsReference
Cellulose and ligninK2CO3
  • Temperature: 300–600 °C
  • Reaction time: 1 h
  • Feed concentration: 7.4 wt% (0.45 M)
  • Pressure: 9–41 MPa
  • Reactor: Batch
  • K2CO3 enhanced gasification efficiency and limited char formation.
  • The highest H2 and CH4 yields were obtained with K2CO3 at 600 °C.
Kang et al. [29]
Cellulose and ligninK2CO3
  • Temperature: 300–600 °C
  • Reaction time: 1 h
  • Pressure: 9–41 MPa
  • Reactor: Batch
  • K2CO3 enhanced the water–gas shift reaction, leading to a high H2 yield.
  • The highest H2 yield of 28 mmol/g was achieved with K2CO3 at 600 °C from SCWG of glucose.
  • Total organic carbon levels decreased with catalyst loading, indicating efficient gasification of the feedstock.
Madenoğlu et al. [30]
GlucoseKOH
  • Temperature: 450–560 °C
  • Reaction time: 6–10 s
  • Feed concentration: 0.2–2 wt%
  • Pressure: 25 MPa
  • Reactor: Continuous
  • The highest heating value of 113% was achieved with KOH catalysts at optimized gasification conditions.
Garcia-Jarana et al. [31]
GlucoseRaney nickel and K2CO3
  • Temperature: 500 °C
  • Reaction time: 1 h
  • Feed concentration: 5 wt%
  • Pressure: 30 MPa
  • Reactor: Batch
  • 0.5 wt% K2CO3 demonstrated better catalytic activity than 1 wt% Raney nickel.
Sınaǧ et al. [32]
Paper sludge and black liquorKOH, K2CO3, and NaOH
  • Temperature: 500–650 °C
  • Reaction time: 2 min
  • Feed concentration: 2–3 wt%
  • Pressure: 25 MPa
  • Reactor: Semi-continuous
  • The highest H2 yield of 25 mmol/g was obtained with K2CO3 at 600 °C from SCWG of paper sludge.
Rönnlund et al. [33]
Sewage sludgeKOH, K2CO3, NaOH, Na2CO3, and AC
  • Temperature: 450 °C
  • Reaction time: 1 h
  • Pressure: 23–26 MPa
  • Reactor: Batch
  • KOH increased gas yield to 12.2 mmol/g from 11.3 mmol/g in the non-catalytic run.
  • K2CO3 demonstrated the highest desulfurization effect followed by Na2CO3, NaOH, KOH, and AC.
Feng et al. [34]
Timothy grassKOH, K2CO3, NaOH, and Na2CO3
  • Temperature: 650 °C
  • Reaction time: 45 min
  • Biomass-to-water ratio: 1:8
  • Pressure: 23–25 MPa
  • Reactor: Batch
  • KOH demonstrated the highest H2 yield (8.9 mmol/g) followed by K2CO3 (7.8 mmol/g), NaOH (6.7 mmol/g), and Na2CO3 (6.3 mmol/g).
  • KOH enhanced the water–gas shift reaction to maximize the H2 yield.
  • NaOH enhanced the methanation reaction favoring CH4 formation at the expense of H2.
Nanda et al. [35]
Table 2. Notable studies on SCWG of waste biomass assisted by heterogeneous catalysts.
Table 2. Notable studies on SCWG of waste biomass assisted by heterogeneous catalysts.
FeedstockCatalystOperating ConditionsMain FindingsReference
2-PropanolPt/Al2O3 and Ru/Al2O3
  • Temperature: 400–550 °C
  • Reaction time: 10–30 s
  • Feed concentration: 0.5 M
  • Pressure: 25 MPa
  • Reactor: Continuous
  • Pt/Al2O3 showed high H2 selectivity at lower temperatures than Ru/Al2O3.
  • Ru/Al2O3 showed 10 mol% H2 compared to 96 mol% H2 in the case of Pt/Al2O3.
  • The low H2 selectivity of Ru/Al2O3 was due to enhancement of the methanation reaction, which led to CH4 yields.
Karakuş et al. [49]
GlucoseNi/Al2O3 and Ni/CeO2-Al2O3
  • Temperature: 400 °C
  • Feed concentration: 9.1 wt%
  • Pressure: 24.5 MPa
  • Reactor: Batch
  • Both catalysts significantly improved the H2 yield and selectivity.
  • Ni/CeO2-Al2O3 showed superior catalytic activity than Ni/Al2O3 with higher yields of total gases and H2.
  • The high activity of Ni/CeO2-Al2O3 was attributed to the inhibition of coke formation and sintering by Ce metal in catalysts.
  • Further addition of Ce improved the H2 yield and selectivity, attaining maxima at 8.5 wt% loading.
Lu et al. [50]; Lu et al. [51]
GlucoseNi/Al2O3 and
Ru-Ni/Al2O3
  • Temperature: 400–500 °C
  • Feed concentration: 45 kg/m3
  • Pressure: 25–35 MPa
  • Reactor: Batch
  • Aerogel-synthesized catalysts showed high H2 yields compared to mesoporous and wet-impregnated catalysts.
  • The supercritical CO2 drying step in the aerogel synthesis method enhanced the surface area and reactant diffusivity to improve catalytic performance.
  • Ru-Ni/Al2O3 demonstrated the highest H2 yield of 4.9 mmol/g.
  • The high H2 yield and stability of Ru-Ni/Al2O3 were due to the inhibition of graphite coke formation by Ru metal.
Hossain et al. [52]
Glucose, cellulose, fructose, xylan, pulp, lignin, and barkNi/Al2O3, Ni/hydrotalcite, Raney nickel,
Ru/C, and Ru/Al2O3
  • Temperature: 380 °C
  • Reaction time: 15 min
  • Feed concentration: 2 wt%
  • Pressure: 25 MPa
  • Reactor: Batch
  • Ni/Al2O3 demonstrated the highest H2 selectivity.
  • Ni/hydrotalcite showed the highest H2 yield for all the feedstocks followed by Ni/Al2O3.
  • The high H2 yield of Ni/hydrotalcite and Ni/Al2O3 was attributed to the poor dispersion of Ni metal.
Azadi et al. [27]
Plastic wastesNiO/γ-Al2O3, RuO2/γ-Al2O3, and bimetallic catalysts
  • Temperature: 450 °C
  • Reaction time: 1 h
  • Feed concentration: 20 wt%
  • Pressure: 25 MPa
  • Reactor: Batch
  • The highest carbon gasification efficiency of 99% was achieved with polypropylene followed by high-density polyethylene, low-density polyethylene, and polystyrene.
  • The highest H2 yield in the non-catalytic run was achieved with low-density polyethylene followed by polystyrene, polypropylene, and high-density polyethylene.
  • Compared to only using NiO, the bimetallic catalyst with RuO2 increased the H2 yield and reduced C2–C4 gas yields.
Onwudili and Williams [53]
Soyabean strawNi supported on carbon nanotubes (CNT), ZrO2, Al2O3, SiO2, and Al2O3-SiO2, and promoted by K, Ce, and Na.
  • Temperature: 500 °C
  • Reaction time: 45 min
  • Biomass-to-water ratio: 1:10
  • Pressure: 23–25 MPa
  • Reactor: Batch
  • Ni supported on ZrO2 and Al2O3 demonstrated superior performance compared to other supports.
  • 10%Ni-1%Ce/ZrO2 showed the highest H2 yield of 10.9 mmol/g and excellent catalytic performance.
  • This was attributed to the high oxygen storage and mobility capabilities of Ce promotors for high reduction and oxidation performance.
Okolie et al. [54]
Waste cooking oilRu/Al2O3, Ni/Si-Al2O3, K2CO3, and Na2CO3
  • Temperature: 375–675 °C
  • Reaction time: 15–60 min
  • Feed concentration: 25–40 wt%
  • Pressure: 23–25 MPa
  • Reactor: Batch
  • The order of catalytic performance in enhancing the H2 yield was Ru/Al2O3 (10.2 mmol/g) > Ni/Si-Al2O3 (9.3 mmol/g) > K2CO3 (8.1 mmol/g) > Na2CO3 (7.5 mmol/g).
  • Ru enhanced the water–gas shift reaction to improve H2 yields.
Nanda et al. [55]
Wheat strawNi/MgO, Fe/MgO, Cu/MgO, Ni/ZnO, Ni/Al2O3, and Ni/ZrO2
  • Temperature: 450 °C
  • Reaction time: 20 min
  • Feed concentration: 7.4 wt%
  • Pressure: 23–28 MPa
  • Reactor: Batch
  • The order of H2 yield was Ni/MgO (11.6 mmol/g) > Fe/MgO (9.2 mmol/g) > Cu/MgO (8.1 mmol/g).
  • Among Ni-based supported catalysts, Ni/MgO demonstrated the highest H2 yields.
  • Basic supports favored water–gas shift reactions, leading to high H2 yields.
Lu et al. [51]
Table 3. Notable studies on SCWG of waste biomass assisted by novel carbonaceous catalysts.
Table 3. Notable studies on SCWG of waste biomass assisted by novel carbonaceous catalysts.
FeedstockCatalystOperating ConditionsMain FindingsReference
Almond shellHydrochar generated from SCWG of wheat straw and algae (Cladophora glomerata)
  • Temperature: 460 °C
  • Reaction time: 10 min
  • Feed-to-water ratio: 0.01
  • Pressure: 25 MPa
  • Reactor: Batch
  • Hydrochar from algal and wheat straw demonstrated H2 yields of 11.6 and 10.8 mmol/g, respectively.
  • Algal hydrochar showed higher H2 yield, H2 selectivity, and total gas yield due to the presence of alkali and alkaline earth metals, which enhanced the water–gas shift reaction.
Safari et al. [91]
Banana pseudo-stemIn-situ impregnated biomass with Fe, Ru, and Ni
  • Temperature: 600 °C
  • Reaction time: 1 h
  • Pressure: 22–25 MPa
  • Reactor: Batch
  • Impregnated metals dramatically improved gasification efficiency and H2 yields.
  • Ni-impregnated biomass showed the highest H2 yield (11.1 mmol/g) followed by Ru (8.8 mmol/g) and Fe (4.2 mmol/g).
  • Ni nanoparticles enhanced cracking and reforming reactions.
  • Ru nanoparticles favored the methanation reaction.
  • Fe nanoparticles formed an oxide layer and promoted H2 yields.
Kumar et al. [92]
Dewater sewage sludge and model compoundsAlCl3 with Ni, KOH, and K2CO3 catalysts with H2O2, K2S2O8, and CaO2 oxidants
  • Temperature: 400 °C
  • Reaction time: 30 min
  • Feed concentration: 5 wt%
  • Pressure: 24 MPa
  • Reactor: Batch
  • AlCl3-K2CO3 led to the highest H2 yield (0.85 mmol/g) followed by Ni (0.8 mmol/g) and KOH (0.7 mmol/g), compared to using AlCl3 alone (7.8 mmol/g).
  • Oxidants performed better than catalysts.
  • AlCl3 combined with H2O2 showed the highest H2 yield of 8.9 mmol/g. However, oxidants alone decreased the H2 yield.
  • Glycerol demonstrated the highest H2 yield followed by guaiacol, glucose, alanine, and humic acid with AlCl3.
Li et al. [71]
GlycerolNi/MgAl2O4-Al2O3, Ni/Al2O3, and Ni/CNT
  • Temperature: 425 °C
  • Reaction time: 54 s
  • Feed concentration: 5 wt%
  • Pressure: 25.2 MPa
  • Reactor: Continuous
  • Ni/CNT demonstrated high stability with superior carbon gasification efficiency.
  • Ni/CNT improved the H2 yield by 2.7 times compared to the non-catalytic SCWG.
  • Ni/MgAl2O4 showed the highest H2 yield (1.5 mol/mol) followed by Ni/Al2O3 (1.0 mol/mol) and Ni/CNT (0.8 mol/mol).
Li et al. [65]
Pinewood and wheat strawIn-situ Ni impregnation
  • Temperature: 500 °C
  • Reaction time: 45 min
  • Pressure: 23–25 MPa
  • Reactor: Batch
  • In-situ Ni-impregnated pinewood showed improvements of 59%, 40%, and 34% in H2 yield, total gas yield, and carbon gasification efficiency, respectively, compared to non-catalytic run.
  • Ni impregnation resulted in an H2 yield of 5.8 mmol/g from wheat straw and 2.8 mmol/g with pinewood.
  • High H2 yield with wheat straw was attributed to a better distribution of Ni nanoparticles in wheat straw.
Nanda et al. [93]
Sugarcane bagasseRu/AC and Ru/TiO2
  • Temperature: 400 °C
  • Reaction time: 15 min
  • Pressure: 22–25 MPa
  • Reactor: Batch
  • Ru/AC demonstrated superior activity compared to Ru/TiO2 with near-complete gasification of sugarcane bagasse in 15 min compared to 30 min using Ru/TiO2.
  • H2 yield increased with repeated use of Ru/AC catalyst but it suffered from deactivation leading to a significant reduction in its activity.
Osada et al. [94]
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Khandelwal, K.; Boahene, P.; Nanda, S.; Dalai, A.K. A Review of the Design and Performance of Catalysts for Hydrothermal Gasification of Biomass to Produce Hydrogen-Rich Gas Fuel. Molecules 2023, 28, 5137. https://doi.org/10.3390/molecules28135137

AMA Style

Khandelwal K, Boahene P, Nanda S, Dalai AK. A Review of the Design and Performance of Catalysts for Hydrothermal Gasification of Biomass to Produce Hydrogen-Rich Gas Fuel. Molecules. 2023; 28(13):5137. https://doi.org/10.3390/molecules28135137

Chicago/Turabian Style

Khandelwal, Kapil, Philip Boahene, Sonil Nanda, and Ajay K. Dalai. 2023. "A Review of the Design and Performance of Catalysts for Hydrothermal Gasification of Biomass to Produce Hydrogen-Rich Gas Fuel" Molecules 28, no. 13: 5137. https://doi.org/10.3390/molecules28135137

Article Metrics

Back to TopTop