Next Article in Journal
Study on the Deposition Uniformity of Triple-Target Magnetron Co-Sputtering System: Numerical Simulation and Experiment
Next Article in Special Issue
Nanosizing Approach—A Case Study on the Thermal Decomposition of Hydrazine Borane
Previous Article in Journal
The Analysis of Materials Strength Used in the Construction of the Flexible Underwater Bell—Batychron
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Perspectives on Nanomaterials and Nanotechnology for Sustainable Bioenergy Generation

by
Kalaimani Markandan
1,* and
Wai Siong Chai
2,*
1
Department of Chemical & Petroleum Engineering, Faculty of Engineering, Technology and Built Environment, UCSI University, Kuala Lumpur 56000, Malaysia
2
Department of Mechanical and Electro-Mechanical Engineering, National Sun Yat-sen University, Kaohsiung 80424, Taiwan
*
Authors to whom correspondence should be addressed.
Materials 2022, 15(21), 7769; https://doi.org/10.3390/ma15217769
Submission received: 18 September 2022 / Revised: 22 October 2022 / Accepted: 27 October 2022 / Published: 3 November 2022
(This article belongs to the Special Issue Technology and Applications of Green Energy-Based Materials)

Abstract

:
The issue of global warming calls for a greener energy production approach. To this end, bioenergy has significant greenhouse gas mitigation potential, since it makes use of biological products/wastes and can efficiently counter carbon dioxide emission. However, technologies for biomass processing remain limited due to the structure of biomass and difficulties such as high processing cost, development of harmful inhibitors and detoxification of produced inhibitors that hinder widespread usage. Additionally, cellulose pre-treatment is often required to be amenable for an enzymatic hydrolysis process. Nanotechnology (usage of nanomaterials, in this case) has been employed in recent years to improve bioenergy generation, especially in terms of catalyst and feedstock modification. This review starts with introducing the potential nanomaterials in bioenergy generation such as carbon nanotubes, metal oxides, silica and other novel materials. The role of nanotechnology to assist in bioenergy generation is discussed, particularly from the aspects of enzyme immobilization, biogas production and biohydrogen production. Future applications using nanotechnology to assist in bioenergy generation are also prospected.

Graphical Abstract

1. Introduction

Research on bioenergy generation has been growing at a relentless pace owing to the concerns that arose from dwindling fossil fuel reserves and the environmental pollution associated with the exploitation of these resources. Studies have reported that emission of greenhouse gases such as carbon dioxide (CO2) from the burning of fossil fuels has led to climate change issues [1,2]. Since the transportation sector accounts to 60% of the estimated oil demand for the year 2030 (116 million barrels per day), it is essential to replace fossil-fuel-based energy sources in vehicles with renewable energy sources [3,4,5]. In an attempt to gradually replace petroleum and coal resources, various renewable energy sources such as solar, wind, hydrothermal and biomass have been explored to date, with an increased interest in the latter, since lignocellulosic biomass is available abundantly from agriculture and forestry, with an estimated annual production of 2 × 1011 tonnes [6]. It was also reported that in 2011, 38 million tonnes of biomass were used for biofuels in the EU, out of 1.2 billion tonnes of biomass that were generated from various crops [7].
Despite the availability of lignocellulosic biomass in abundance, the technologies for biomass processing remain limited. For instance, the biomasses are recalcitrant in nature due to their cellulose crystallinity and non-reactive lignin, thus requiring pre-treatment techniques to allow the cellulose to be amenable for an enzymatic hydrolysis process [8]. This is required for the extraction of fermentable sugar for subsequent biofuel generation processes [9,10]. Figure 1 shows the cellulose and hemicellulose that accommodate 6 and 5 carbon sugars rigidly bounded with the lignin, where cellulose forms structure of the cell walls, while hemicellulose assists in the cross-linking between the non-cellulosic and cellulosic polymer via covalent bonding [11]. It is noteworthy that biomass resources currently deployed as feedstock in human or non-human food chains should not be used for chemical processing to ensure sustainability. For instance, wheat straw is a common fodder for raising animals and therefore is available in short supply due to increase in meat consumption; wheat straw should not be considered as a viable option for bioenergy generation [12].
Generally, biochemical or thermochemical routes can be employed for the conversion of lignocellulosic raw materials into bioenergy. The biochemical route uses microorganisms and various enzymes to convert components of the feedstock into sugars, followed by fermentation to produce ethanol [13]. Since the biochemical route uses various enzymes to convert feedstock to biofuel, the efficacy of the enzyme is important to improving the conversion process. On the other hand, the thermochemical route involves gasification technologies to produce simple sugars that are fermentable for biofuel production. In a study by Daystar et al. [14], it was highlighted that thermochemical conversion can use a wider range of feedstock compared to biochemical conversion and produces considerably higher alcohol yields. Additionally, unlike the biochemical process, the thermochemical process is not affected by the lignin in the biomass, although presence of moisture content can affect alcohol yield as well emissions in the thermochemical process. A general finding from most studies indicates that high cost and limited availability of existing infrastructure have posed limitations on obtaining high quality and yield of bioenergy.
However, with the emergence of nanotechnology, the limitations can be addressed efficiently since nanotechnology involves nanomaterials, which exhibit unprecedented characteristics and properties, all of which are highly useful in areas of bioenergy generation. For example, nanotechnology has proven to be advantageous in the bioenergy field for various applications such as feedstock modification and development of efficient catalysts for the hydrolysis of biomass to produce biofuels such as ethanol and biogas or for the catalysis of biodiesel production from oils and fats [15,16,17,18,19,20]. Within this context, various nanoparticles such as magnetic nanoparticles (MNPs), carbon nanotubes (CNTs), porous metal oxides, mesoporous silica and metallic organic frameworks (MOFs) have been commonly used to replace enzymes or immobilize them, which results in an efficient catalysis process for the production of bioenergy.
Whilst much of the emphasis has been intensifying on experimental studies of improving bioenergy yield of lignocellulosic biomass with the aid of nanoparticles, comprehensive review on how nanotechnology and nanomaterials affect bioenergy production is still limited. Herein, the authors make an attempt to gather information from the early to the most recent developments regarding the various nanomaterials and their role in bioenergy generation. More specifically, the paper is aimed at discussing three principle topics in depth: (i) the various nanomaterials in bioenergy generation (MNPs, CNTs, porous metal oxides, mesoporous silica and MOFs), (ii) the role of nanoparticles in bioenergy generation and (iii) the role and importance of nanotechnology in bioenergy generation. Lastly, we highlight the challenges of and prospects for future progress in bioenergy generation from lignocellulosic biomasses utilizing nanomaterials.

2. Potential Nanomaterials in Bioenergy Generation

The International Standardization Organization (ISO) has classified nanomaterials as any material with either its external, internal or surface structure having a dimensions in the nanoscale range of 1–100 nm [21]. In most cases, nanostructured materials are classified based on their dimensionality, such as zero-dimensional (e.g., graphene/carbon/inorganic quantum dots, fullerenes, magnetic nanoparticles), one-dimensional (e.g., nanotubes, nanofibers and nanowires), two-dimensional (e.g., nanosheets, nanoflakes of: graphene, boron nitride, MXenes, transition metal dichalcogenide) or three-dimensional (nanophase materials composed of equiaxed nanometer-sized grains) materials. The physical and chemical properties of nanoparticles (NPs) vary significantly from their macro counterparts, which have been widely exploited for use in bioenergy generation. For instance, the faster reaction rate (high catalytic activity) of NPs owing to the relatively fine particle size of NPs (i.e., large surface-area-to-volume ratio, which increases the number of active sites for various reactions to occur) is highly advantageous in the bioenergy generation process. Furthermore, other characteristics of NPs such as crystallinity, durability, stability, adsorption capability, and efficient storage have been reported to be highly desirable in the bioenergy generation field to enhance the productivity, hydrolysis and stability of cellulase enzymes [22,23].
NPs can be broadly classified as organic or inorganic nanoparticles. Examples of organic NPs are liposomes, polymersomes, polymer constructs and micelles, all of which have found great usage in imaging or drug and gene delivery techniques. On the other hand, inorganic NPs such as gold, quantum dots, carbon nanotubes and magnetic nanoparticles have drawn significant research and commercial interest in bioenergy production owing to their excellent physical properties (optical, magnetic) as well as their chemical properties (inertness, stability and ease of functionalization). Hybrid NPs, which combine both organic and inorganic NPs, exhibit the properties of both parent NPs and have shown improved performances (better catalyst recovery, higher selectivity) over the parent NPs in certain cases. Recently, single-atom catalysts (SACs) possessing single-atom centers decorated on support have drawn significant research interest among the scientific community due to biomass conversion attributed to its unique geometric configuration, electronic properties and ensemble effect [24]. In comparison to a traditional catalytic system, the SACs combine advantages of both homogenous and heterogenous catalysts by showcasing enhanced performance, increased thermal and chemical stability (i.e., due to stabilization of metal center on support) and easy recoverability. In fact, it has been reported that SACs displayed excellent catalytic activity and selectivity, i.e., 95% in most cases [25]. The following section will review the various NPs that have been used in bioenergy production.

2.1. Magnetic Nanoparticles (MNPs)

The magnetic properties of MNPs are a result of the intrinsic properties of the particles as well as the interaction between the particles. MNPs perform optimally when the particle size is below the critical value of 10–20 nm, since at this size range, NPs behave as a single magnetic domain, and with application of a magnetic field, they behave as giant paramagnetic atom [3,26]. Some advantages of MNPs include their high surface-area-to-volume ratio, excellent quantum properties and small size, which enables them to carry other compounds. Examples of materials that have been used to build MNPs include cobalt, iron, nickel and platinum [27,28]. One of the major advantages of MNPs that supersedes their counter NPs is the ability to be coated and used as catalyst through immobilization, which can later be easily removed (recoverability) via application of a suitable magnetic field [29,30]. However, one major limitation of MNPs is their intrinsic instability over a large duration of time, during which the small particles form agglomerates to reduce the energy associated with the high surface-area-to-volume ratio. Additionally, since the MNPs are chemically very active, they can easily lose their magnetic properties by oxidation and dispersion. As such, strategies such as coating or modification with surfactants or inorganic layers of silica and carbon can be essential to protect the MNPs against chemical destruction during and after the synthesis stages [31]. The functionalized MNPs can be useful in bioenergy production, since they will be highly stable (i.e., highly dispersible), reactive and easily separable at later stages. A schematic illustration describing the use of MNPs immobilized cellulase enzyme in hydrolysis of lignocellulosic biomass for the production of biofuel is shown in Figure 2.

2.2. Carbon Nanotubes (CNTs)

CNTs consist of graphite sheets rolled up into a cylindrical shape with a nanometric diameter, and they have excellent biocompatibility and mechanical strength. Additionally, CNTs have a large surface area, which enables a high capacity to load enzymes and low diffusion resistance [33]. However, the biocompatibility of CNT composite is affected by CNT amount and length, as well as the blended substance, which might render it carcinogenic [34]. All the excellent properties of CNTs are appealing for use in enzyme immobilization. The ability of CNTs to attach themselves to the rooted sites of enzymes for direct electron transfer and the 3D electro-active area of CNTs, which increases enzyme concentration and other redox compounds on its surface, have made CNTs a very important material in biofuel applications [35,36]. Specifically, multi-walled CNTs (MWCNTs) perform comparatively better than single-walled CNTs (SWCNTs) since the immobilization of enzymes is highly compatible with the structural arrangement of MWCNTs, which further enhances the catalytic activities of immobilized enzymes. For example, Ahmad et al. [37] and Mubarak et al. [38] reported that MWCNTs outperformed hydrolysis of cellulose from Aspergillus niger, where an efficiency of 85% to 97% was attained, while retaining recyclability by 52% to 75% after six cycles of a hydrolysis process. Surface functionalization of CNTs has also been reported to further enhance the catalytic activity of the immobilized enzymes. For example, in a study by Pavlidis et al., the immobilization of lipase on the functionalized MWCNTs supports using the cross-liner glutaraldehyde enabled the biomaterial to retain >55% of its activity after 6 months at 4 °C and 25% of their initial activity after 30 days of incubation in hexane at 60 °C [39].

2.3. Porous Metal Oxides

Mesoporosity in metal oxides can be achieved by using various templates such as block copolymers [40], amines [41] or surfactants [42]. Additionally, some studies have also reported that mesopores can be introduced on the surface of metal oxide particles using soft templates such as aspartic acid and salicylic acid to convert glucose and fructose to 5- hydroxymethylfurfural (HMF). For instance, in a study by Dutta et al. [43], using sodium salicylate as a template, direct conversion of carbohydrates into HMF was investigated over a self-assembled mesoporous TiO2 nanoparticles (NPs) catalyst. It was reported that the high surface area for the self-assembled mesoporous TiO2 nanomaterials along with the Lewis acidity enabled high catalytic activity, which was able to produce 34.3% and 54.1% HMF in water and dimethyl sulfoxide (DMSO), respectively. Similarly, in another study by De at al. [44], the biopolymer sodium alginate templated porous TiO2 nanocatalyst with large pores in nanoscale dimension at the material surface assisted in the high catalytic activity, as under hydrothermal conditions, the material was able to catalyze the transformation of unutilized sugar derivates (i.e., D-mannose, D-galactose and lactose) to HMF under microwave irradiation at 140 °C, producing up to 44% yield. On the other hand, the TiO2-H was capable of retaining its catalytic activity for four cycles, suggesting its high importance in biomass conversion. Other studies have used reduced copper porous metal oxides for the supercritical methanol depolymerization and hydrodeoxygenation of biomass [45], mesoporous Al2O3 for the dehydration reaction for conversion of glucose to HMF [46] and highly ordered mesoporous sulfated zirconium for biodiesel production from oleic acid and jatropha oil [47]. Additionally, Sn or Zr metals have been the more preferred option compared to Ti metals, since they have low electronegativity, and hence the metal protons can be released easily; thus, SO42−/SnO2 or SO42−/ZrO2 show stronger acidity compared to SO42−/TiO2, while the low surface area of the sulfated nonporous oxide also restricts reactants from accessing the active acid sites [48].

2.4. Mesoporous Silica

The properties of mesoporous silica, such as large surface area and pore volume with tunable pore diameter, simple surface functionalization, hydrothermal stability and biocompatibility, have been reported to be advantageous in immobilization of a range of enzymes [49,50]. For instance, the mesoporous silica material SBA-15 has been commonly used in comparison to MCM-41, owing to its large pore diameter and its mechanical and hydrothermal stability. In particular, the pores attached to organic groups such as -SO3H and -COOH with uniformly distributed pore size are highly essential within the class of mesoporous materials, since they play an important role in the field of heterogenous catalysis. For example, the silica-based materials, which have a high concentration of silanol groups, are important in covalent immobilization of active sites, which also allows for additional organic functionality to control the surface hydrophobicity, all of which are characteristics highly desirable to enhance aspects of catalytic performance such as durability, activity and selectivity in acid-catalyzed chemical reactions [51].
In many studies, active sites of functionalized mesoporous silica have been reported to be appealing for biomass conversion. For example, in a study by Dufaud and Davis, it was shown that mesoporous silica (SBA-15) with active sites such as sulfonated acid using organic–inorganic hybrid silane showed enhanced catalytic activity and selectivity [52]. In another study by Poorakbar et al. [53], magnetic gold mesoporous silica nanoparticle core shells were fabricated as a support for cellulase immobilization via covalent bonding, and it was reported that the immobilized enzyme maintained 58% of its initial catalytic activity even after 9 h. Additionally, the support for cellulase immobilization exhibited enhanced thermal stability, applicable for a broader temperature and pH range, while showcasing its capability for long-term storage as well as easy separability using an external magnet (Figure 3). A study by Lee et al. [54] used Fe3O4 encapsulated mesoporous silica nanoparticles for cellulosic biomass conversion, and it was reported that a maximum fructose yield of 51% was achievable under optimized reaction temperature (50 °C), time (24 h) and pH values (phosphate buffer with pH 4.8). Authors have highlighted the effectiveness of Fe3O4 -enzyme immobilized mesoporous silica nanoparticles for the deconstruction of glycoside bonds of cellulose at a pH of 4.8 and the possibility of integrating enzymatic and chemocatalytic biomass processing for various catalytic applications. Other studies have reported using molybdenum supported on KIT-5 mesoporous silica catalyst for the production of biofuels and value-added chemicals (furans and phenols) from biomass through catalytic fast pyrolysis of pine [55], organically modified biogenic mesoporous silica nanoparticles (RKIT-6) for xylanase immobilization for lignocellulosic biomass hydrolysis [56] and nickel nanoparticles supported on 2D (COK-12) and 3D (KIT-6 and SBA-16) mesoporous silicas for the hydrocyclization of biomass-derived levulinic acid [57].

2.5. Metal Organic Frameworks (MOF)

Metal organic frameworks (MOF) are highly useful in heterogenous catalysts, since they have high surface area with modifiable metal–organic frameworks. Researchers have utilized the merits of MOF to develop MOF-derived solid acid as a heterogenous catalyst for the biotransformation of carbohydrates to HMF, and it has been reported that the high porosity and ordered structure of the catalyst enables more substrate transfer within the MOF catalyst. For instance, the first example of MOF application in carbohydrate dehydration was reported in a study by Zhang et al. [58] using MIL-101, a chromium-based MOF that is well known for its large pore size, large surface area and good stability. Phosphotungstic acid (H3PW12O40, PTA) encapsulated in MIL-101 was used as the potential catalyst for the selective dehydration of fructose and glucose to HMF, where under optimal conditions HMF yield of 79% was attained from fructose after 2.5 h (Figure 4). In another study by Chen at al., a series of sulfonic-acid-functionalized MOF, MIL-101 (Cr)-SO3H, was used as the catalyst for the conversion of fructose into HMF, and it was reported that an HMF yield of 90% was obtained with full fructose conversion at 120 °C for 60 min in DMSO. MIL-101(Cr)-SO3H behaved as the heterogenous catalyst and has easy recoverability and reusability, all of which highlights the good prospects of MOF-derived solid acid catalyst for biomass carbohydrate valorization [59].
Cellulose and cellobiose were also selectively converted to sorbitol over phosphotungstic acid (PTA)-MOF hybrid supported ruthenium catalysts Ru-PTA/MIL-100(Cr) under aqueous hydrogenation conditions, where under optimal conditions, 63.2% yield in hexitol with sorbitol selectivity of 57.9% at complete conversion of cellulose and 97.1% yield in hexitol and 95.1% selectivity of sorbitol at complete conversion of cellobiose were obtained [60]. This study provided a breakthrough in perspectives for the rational design of acid–metal bifunctional catalysts for efficient biomass conversion. Other studies have reported that post-synthetic modification of the MOF functional group, which contains an organic ligand component, is more advantageous than conventional carbon materials or inorganic solids [61,62].

2.6. Zeolites

Zeolites have strong Brønsted and Lewis acid sites with well-arranged micropores and excellent thermal stability owing to their crystalline inorganic framework and have been commonly used as a principle heterogenous catalyst in petrochemical and fine industrial plants. However, it is noteworthy that limitations of zeolites such as sensitivity to hot water have limited their usage in aqueous phase processes such as biomass conversion. A schematic diagram illustrating the synthesis of an iron-oxide-nanoparticle–zeolite-based hybrid MNP is shown in Figure 5.
Among the numerous zeolites, ZSM-5 based catalysts have been extensively researched, and in a comprehensive review written by Stocker et al., it was reported that the active acidic sites of zeolite (i.e., H-ZSM) that behaves as carbonium ion can be effective in the pyrolysis process of wood biomass [63]. The acidity and porosity of ZSM-5 based catalysts can be tuned by altering the content of Si or Al in zeolite. For example, in a study by Srinivasan et al., SiO2/Al2O3 ratios of zeolite (ZSM-5) catalyst in aromatic hydrocarbon production through catalytic pyrolysis of cellulose were investigated, and it was shown that acidity of catalysts plays an essential role in eliminating anhydro sugars and other oxygenated compounds to produce more aromatics [64]. Maximum aromatic yield (≈25 wt.%) was obtained using ZSM-5 with the highest acidity (SiO2/Al2O3 = 30), while the lowest yield (≈7.99 wt.%) was obtained with the least acidic catalyst (SiO2/Al2O3 = 280).
Additionally, doping of other metal elements such as Mn, Co, Ni, Ga, Ce or Pt into ZSM-5 based catalysts can be efficient to regulate the strength and density of acidic sites in zeolite. For instance, in a study by Vichapund et al., it was reported that catalytic pyrolysis of jatropha residues (cellulose: 59.2%, hemicellulose: 18%, lignin: 22.8%) with 3 wt.% Ni-ZSM-5 catalyst provided an ideal environment for the oligomerization, cyclization and dehydrogenation of small olefins, all of which improved the yield of aromatic compounds [65]. In another study, Zhou et al. used silicon carbide foam supported on ZSM-5 catalyst for the microwave-assisted pyrolysis of biomass and reported that increasing the catalyst-to-biomass ratio (0 to ½) improved the quality of bio-oil at the cost of its yield (yield reduced from 40.2% to 33.5%) [66]. The lower bio-oil yield with higher gas yield is attributed to the enhanced catalytic conversion by ZSM-5 catalyst; while primary pyrolytic vapor enters the pore of ZSM-5, oxygen is removed in the forms of CO, CO2, and H2O through decarbonylation, decarboxylation and dehydration reactions. It was also reported that the catalyst was capable of retaining its activity up to a catalyst-to-biomass ratio of 1/10, which outperformed ZSM-5 catalyst configurations reported in other studies while also allowing its regeneration for reuse with preserved material and catalytic properties for up to seven reaction–regeneration cycles. In another study by Jorge et al. [67], a series of new catalysts consisting of magnetically recoverable β-zeolites exchanged with transition metals showed significant potential for catalytic valorization of hemicellulosic biomass compounds. In particular, Pd-exchanged catalyst showed enhanced efficiency for conversion of both furfural and furfuryl alcohol to isopropyl levulinate.
Figure 5. Schematic diagram describing the synthesis of iron-oxide-nanoparticle–zeolite-based hybrid MNP, where ferric trichloride was initially added to eucalyptus leaf extract and stirred for an hour at 70 °C before being filtered. In the second stage, zeolite was added to the eucalyptus–MNP solution, and in the final stage, the eucalyptus–MNP–zeolite solid was filtered and washed with anhydrous ethanol and deionized water before drying in vacuum oven at 45 °C for 12 h. Reproduced from [68] with permission from Elsevier.
Figure 5. Schematic diagram describing the synthesis of iron-oxide-nanoparticle–zeolite-based hybrid MNP, where ferric trichloride was initially added to eucalyptus leaf extract and stirred for an hour at 70 °C before being filtered. In the second stage, zeolite was added to the eucalyptus–MNP solution, and in the final stage, the eucalyptus–MNP–zeolite solid was filtered and washed with anhydrous ethanol and deionized water before drying in vacuum oven at 45 °C for 12 h. Reproduced from [68] with permission from Elsevier.
Materials 15 07769 g005

3. Role of Nanotechnology in Bioenergy Production

3.1. Enzyme Immobilization

Enzyme immobilization on support materials is useful to enhance enzyme characteristics such as high activity at most pH values, reusability and selectivity and reduce their inhibition. Some studies have reported that immobilization of cellulase enzyme makes it more resistant to structural alterations which may be caused by increase in the temperature [29]. An excellent enzyme immobilization on solid support ensures good distribution of the catalyst with minimal agglomeration. For instance, covalent binding between the support and enzyme have been reported to increase the enzyme activity [53,69]. Other than covalent binding, approaches such as adsorption, ionic bonding, entrapment and encapsulation have been reported for enzyme immobilization. In comparison to other NPs, MNPs have gained significant interest as promising support carriers in enzyme immobilization due to factors highlighted in Section 2.1 of this work. Most importantly, in enzyme immobilization, the large surface-area-to-volume ratio of MNPs is highly advantageous to loading a large number of enzymes on their surface, which in turn increases the catalytic activity of the enzymes. Several techniques can be used for immobilization of enzymes on the MNP support. For example, MNP functionalization for cellulase immobilization can be on silica-functionalized MNP, amino-functionalized MNPs, composite-functionalized MNPs, chitosan-functionalized MNPs or carrier-free functionalized MNPs.
For instance, sulfonated magnetic carbonaceous nanoparticles, which were used for the hydrolysis of various lignocellulosic biomass, showed significant levels of glucose yields of jatropha, bagasse and plukenetia hulls of 35.6%, 58.3% and 35.8%, respectively, with the capability of recycling the catalyst at least seven times with a high catalyst recovery rate of 92.8% [70]. In another study by Goh et al., enzyme immobilization was performed on magnetic single-walled carbon nanotubes for application in biofuel production. It was reported that the enzyme was recyclable, which reduced the cost of biofuel production while also retaining their activity for a month in acetate buffer at 4°C [71]. In another study, 93% binding efficiency and 50% activity retainment after 16th cycle was achieved with a nanobiocatalyst system for biofuel production using functionalized magnetic nanoparticles immobilized with β-glucosidase isolated from fungus [72]. In biodiesel production, immobilization of lipase on polydopamine-coated magnetic nanoparticles enhanced the pH and thermal stability while achieving 73.9% binding efficiency, more than 70% of activity even after 21 repeated cycles and easy recoverability from reaction mixture [73]. Similar findings were also reported in a study by Hu et al., where immobilization on magnetic nanoparticles increased biodiesel production by 95% when a reaction was performed at 65 °C for a duration of 3 h. Additionally, the catalyst was reusable 14 times while allowing recoverability of more than 90% [74]. To date, various MNPs have been explored as support carriers in enzyme immobilization. A summary of the various MNPS that have been used as the support carriers in enzyme immobilization and their effect on bioenergy production is summarized in Table 1.

3.2. Biogas Production

In the biogas production process, nanoparticles have proven to show promising results in the anaerobic processes (biochemical processes to convert various organic materials into biogas and its constituents), in particular as electron donors, acceptors and cofactors of key enzymes such as [Fe]- and [Ni-Fe]-hydrogenase [84]. For example, nanoparticles increase the hydrolysis of organic matter due to their large surface-area-to-volume ratio, which enables microorganisms to bind onto the active sites of the molecule. This subsequently stimulates the biochemical process for the activity of hydrogenase enzymes and ferredoxins. Various nanoparticles have been used in the anaerobic digestion process such as zero-valent metals, metal oxides and carbon-based nanomaterials.
Generally, the anaerobic digestion process involves four important steps: (i) hydrolysis, (ii) acidogenesis, (iii) acetogenesis and (iv) methanogenesis. Nanoparticles such as carbon-based materials have been reported to be useful in all the four stages. For example, in a study by Velimirovic et al. [85], carbon-based nanomaterials in the hydrolysis stage, when available as electron donors, can lower the oxygen reduction potential of the water environment, which is highly beneficial in the hydrolysis reaction. In the acidogenesis and acetogenesis processes, studies have reported that the addition of 12.5 g/L biochar enhanced the biohydrogen production from 22.6 to 96.3 mL/g [86], the addition of 8.3 g/L of biochar increased the biohydrogen yield from 750.4 to 944.5 mL/L from food waste [87], and the addition of 600 mg/L of sawdust biochar improved biohydrogen production from 31.5 to 36.5 mL [88]. In all the aforementioned studies, it was highlighted that addition of carbon-based nanomaterials enhanced the performance of acid formation processes, which resulted in improved yield of hydrogen. Similar enhancement in biogas production were also reported in other studies using zero-valent ions (ZVI); i.e., 0.1 wt.% ZVI using waste-activated sludge enhanced biogas production by 30.4% [89], whereas the highest methane formation rate (0.310 mmol CH4 formed/mol Fe0. day) was reported using the finest grade iron ZVI [90].
Studies have also been performed to study the effect of nanoparticles on microbial communities in the anaerobic digestion process. For example, Wang et al. used four types of nanoparticles, i.e., ZVI, Ag, Fe2O3 and MgO, to investigate their effect on biogas production from waste-activated sludge, and it was reported that ZVI (10 mg/g) and Fe2O3 (100 mg/g TSS) increased biogas production by 120% and 117%, respectively, which suggests that ZVI and Fe2O3 nanoparticles improve the activity of the methanogenic bacteria [91]. This was also supported in another study by Yang et al., which reported that addition of ZVI nanoparticles increases the population of methanogens in an anaerobic digester [92].

3.3. Biohydrogen Production

Biohydrogen production is performed by anaerobic bacteria via metabolic routes to generate molecular hydrogen. It has been reported that the activity of these microorganisms can be enhanced with the use of nanoparticles to increase electron transfers and kinetics in metabolic processes to produce biohydrogen due to their capability to react faster with electron donors. The role of nanoparticles in various biohydrogen process is as described below.

3.3.1. Dark Fermentative Biohydrogen Process

Nanoparticles have been reported to improve the dark fermentation process. For instance, 5 nm gold nanoparticles improved substrate utilization and biohydrogen yield by 56% and 46%, respectively, owing to the large surface-area-to-volume ratio of gold nanoparticles, which provided a stimulatory effect to produce biohydrogen [93]. It has also been highlighted in other studies that gold nanoparticles enhance the activity of biohydrogen-producing enzymes, which mediate the transfer of electrons such as [Fe-Fe]- and [Ni-Fe]-hydrogenases and ferredoxins [94,95]. In other studies, zero-valent iron (Fe0) nanoparticles were used for dark fermentation of grass, and it was shown that Fe0 nanoparticles stimulated activity of a hydrogenase enzyme to produce a higher yield of biohydrogen (i.e., maximum hydrogen yield of 64.7 mL/g dry grass; 73% higher than control experiments) [96]. Other studies have also reported the use of metallic nanoparticles such as Pb, Ag and Cu along with FeO nanoparticles immobilized on porous silica (SiO2), and highest yield and production rate was achieved using FeO nanoparticles, i.e., 38% and 58%, respectively, in comparison to control experiments [97]. Ni-graphene-based nanoparticles have also been reported to enhance dark fermentative biohydrogen production, in a study where maximum yield of 41.3 H2/g COD with 105% increase in H2 yield was obtained [98].

3.3.2. Photo-Fermentative Biohydrogen Production

Nanoparticles have also drawn significant interest as means to improve the biomass growth, photosynthetic activity, nitrogen metabolism and protein level of microalgal species to produce biohydrogen. The key roles of nanoparticles in photo-fermentative biohydrogen production are as summarized below:
  • Nanoparticles behave as the catalytic agents to generate metabolic pathways to promote synthesis of chlorophyll a, chlorophyll b, carotenoids and anthocyanin, lipid production and nitrogen metabolism [99,100].
  • Nanoparticles enhance production of carbohydrates, which results in increased growth of algal cells. For example, silica nanoparticles enhanced growth of microalgal cells (measured from chlorophyll concentration), because silica nanoparticles scattered light within a reactor to ensure uniform light distribution during the photosynthetic process, which in turn promoted growth of microalgal cells [101]. Similar findings were also reported using zero-valent iron (Fe0) [102] and TiO2 nanoparticles, which increased chlorophyll and carotenoid pigments [103].
  • Nanoparticles enhance activity of key enzymes for metabolism of microalgal species such as glutamate dehydrogenase, glutamate–pyruvate transaminase, glutamine synthase and nitrate reductase [104,105]. The capability of nanoparticles to maintain the pH of a medium and to promote the activity of hydrogenase enzymes and substrate hydrolysis may promote higher biohydrogen yield by enhancing biohydrogen-producing metabolic pathways such as acetate and butyrate reactions [32,103,106].

3.3.3. Photocatalytic Hydrogen Production

Photocatalytic biohydrogen production involves the breaking of water molecules into H2 and O2 in the presence of an illuminating source. A photocatalyst such as TiO2 has been commonly preferred owing to its high photocatalytic performance, non-toxicity, chemical stability and cost effectiveness [107,108]. For instance, nanoparticles such as Pt-TiO2-activated carbon generated a high hydrogen production rate of 7490 µmol/h/g photocatalyst, which is 75 times higher than conventional Aeroxide TiO2 P25 catalyst [107]. Other studies have reported the combination of TiO2-graphene has higher light absorption efficiency than TiO2 alone, which resulted in higher charge separation efficiency to yield higher hydrogen [109]. The effect of various nanoparticles on the hydrogen yield rate is summarized in Table 2.

3.4. Bioethanol Production

Generally, lignocellulosic materials are processed for bioethanol production via three major operations, which include (i) pretreatment for delignification to liberate the cellulose and hemicellulose, (ii) hydrolysis to produce fermentable sugars such as glucose, xylose, arabinose, galactose or mannose and (iii) fermentation of reducing sugars. In a very recent study, Saeed et al. [124] used graphitic carbon nitride (g-C3N4) nanomaterials (Figure 6A) and laser irradiation (Figure 6B) to increase bioethanol production from potato waste. Authors reported that the control sample without laser irradiation or g-C3N4 showed only 4% yield of bioethanol, while the addition of g-C3N4 coupled with laser irradiation increased bioethanol yield to 56.8% (Figure 6D).
In another study by Gupta et al. [125], zinc oxide nanoparticles were used to enhance bioethanol production from rice straw, and a maximum ethanol yield of 0.0359 g/g dry weight-based plant biomass was attained at 200 mg/L concentration of ZnO nanoparticles (Figure 6C). Additionally, the possibility for reusability and recovery of the nanoparticles makes the entire process more economical. Similarly, in another study by Ivanova et al. [126], it was reported that bioethanol fermentation was enhanced with the use of alginate magnetic nanoparticles entrapped with yeast cells, with the productivity rate reaching up to 264 g/Lh at 70% particle loading. It was also interesting to note that the magnetic particles with fixed yeast cells were stable for more than a month at 4 °C in saline condition. In a similar vein, Kim et al. [127] reported enhanced bioethanol production in syngas fermentation using methyl-functionalized silica nanoparticles. After 9 h of cultivation time, it was reported that ethanol concentrations without and with nanoparticles were 0.1150 and 0.3060 g/L,w respectively, which indicates that ethanol production was enhanced by 166.1% by the use of nanoparticles. When producing liquid fuels through fermentation, it is expected that nanomaterials will influence the biochemical conversion process by affecting the enzymatic activity or the mass transfer rate.

4. Concluding Remarks and Future Prospects

Development and usage of bioenergy is of great importance towards a greener future. Performance enhancement such as enhancement of catalytic and microbe activities, and reduction in inhibition formation have been shown to result from using nanomaterials. For example, nanoparticles have been used as additives to improve the working performance of biodiesel [128].
Generally, these nanomaterials offer large surface area, highly modifiable framework, and high tunability in their performance. All of them have shown good recoverability after several runs. MNPs can be coated and immobilized as catalyst during hydrolysis, as well as possessing high recoverability. Coated MNPs are more resistant against chemical destruction during synthesis stages. Enzyme immobilization is highly compatible with the structural arrangement of MWCNTs, as well as mesoporosity in metal oxides and silica, further enhancing the catalytic activities. Biocompatibility of silica is advantageous in enzyme immobilization, whereas the silanol group on silica can help to control the surface hydrophobicity and enhance catalytic performance such as durability, activity and selectivity in acid-catalyzed chemical reactions. MOF is employed as a heterogenous catalyst owing to its high porosity and ordered structure, enabling more substrate transfer. Zeolites usage in aqueous phase is limited as they are sensitive to hot water.
Studies on employing nanotechnology in bioenergy generation have been conducted since the late 2000s. The incorporation of nanotechnology can be highly advantageous, in particular when used to design catalysts at the atomic level to increase selectivity, efficiency, operational temperature range, and many important properties of the catalyst. However, despite several advantages of employing nanotechnology in such a field, further understanding its interaction and relevant mechanism is required, as nanomaterials exhibit their own characteristics when interacting with biomass. For example, enzyme immobilization, improved catalytic activity and biohydrogen production are the important results when employing nanomaterials for bioenergy generation purposes. Different nanomaterials have been reviewed in this paper, and the relevant interactions are outlined.
Enzyme immobilization can be considered as the main role of nanomaterials in terms of bioenergy production owing to the large surface-area-to-volume ratio of nanoparticles which increases the enzyme activity at the same time. Enhancement of enzyme activity at most pH values, reusability, selectivity and inhibition reduction are the key criteria in enzyme immobilization. Nanomaterials acting as support carriers are successful in obtaining improved catalytic activity and stability. In terms of biogas production, microorganisms are able to bind to the active site of nanoparticles, increasing the organic matter hydrolysis rate and thereby improving the anaerobic digestion process. Carbon-based nanomaterials have been shown to improve all four important steps in the anaerobic digestion process. Additionally, the nanoparticles aid the population increase in methanogens in an anaerobic digester. For biohydrogen production, nanoparticles catalyze the metabolic pathway for chlorophyll synthesis, as well as carbohydrates.
Despite the advantages offered by nanomaterials for bioenergy generation, there are still several issues which hinders their widescale usage. From an economic perspective, nanomaterials are relatively expensive, which inhibits widescale usage of nanomaterials for industrial scale. However, carbon-based nanomaterials are becoming cheaper currently, owing to the efforts from various industrial producers. Considerable efforts are still required for further reduction in cost of other nanomaterials. Besides, recoverability issue of nanomaterials leads to loss of profit. With further research, the recoverability issue can be resolved in the near future. Additionally, although toxicity induced by nanoparticles has hitherto not been understood, metal nanocatalyst or nanoparticles emitted as exhaust from vehicles or industries may be deposited in lung tissue via respiration, disrupt normal functioning of human cells or cause respiratory ailments such as asthma or bronchitis. Thus, safety assessment is of utmost importance to minimize such concerns when utilizing nanomaterials for bioenergy generation.

Author Contributions

Conceptualization, writing—original draft preparation, data curation, investigation, K.M. and W.S.C.; writing—review and editing, supervision, project administration, W.S.C. and K.M. All authors have read and agreed to the published version of the manuscript.

Funding

Authors would like to thank CERVIE, UCSI University, for partially supporting APC for publication.

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Acknowledgments

The authors acknowledge the copyright permission (Creative Commons Attribution License) for reproducing Figure 1 from Suitability of magnetic nanoparticle immobilised cellulases in enhancing enzymatic saccharification of pretreated hemp biomass by R.E. Abraham et al., published 11 June 2014 in Biotechnology for Biofuels and Bioproducts. The rights for Figure 5A–C belong to Springer Nature.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Erickson, P.; Lazarus, M.; Piggot, G. Limiting fossil fuel production as the next big step in climate policy. Nat. Clim. Chang. 2018, 8, 1037–1043. [Google Scholar] [CrossRef]
  2. Johnsson, F.; Kjärstad, J.; Rootzén, J. The threat to climate change mitigation posed by the abundance of fossil fuels. Clim. Policy 2019, 19, 258–274. [Google Scholar] [CrossRef] [Green Version]
  3. Antunes, F.A.F.; Gaikwad, S.; Ingle, A.P.; Pandit, R.; dos Santos, J.C.; Rai, M.; da Silva, S.S. Bioenergy and Biofuels: Nanotechnological Solutions for Sustainable Production; Springer: Berlin/Heidelberg, Germany, 2017; ISBN 9783319454597. [Google Scholar]
  4. Chai, W.S.; Bao, Y.; Jin, P.; Tang, G.; Zhou, L. A review on ammonia, ammonia-hydrogen and ammonia-methane fuels. Renew. Sustain. Energy Rev. 2021, 147, 111254. [Google Scholar] [CrossRef]
  5. Du, H.; Shen, P.; Chai, W.S.; Nie, D.; Shan, C.; Zhou, L. Perspective and analysis of ammonia-based distributed energy system (DES) for achieving low carbon community in China. iScience 2022, 25, 105120. [Google Scholar] [CrossRef] [PubMed]
  6. Elumalai, S.; Agarwal, B.; Runge, T.M.; Sangwan, R.S. Advances in Transformation of Lignocellulosic Biomass to Carbohydrate-Derived Fuel Precursors. In Biorefining of Biomass to Biofuels; Springer: Berlin/Heidelberg, Germany, 2018; pp. 87–116. ISBN 9783319676784. [Google Scholar]
  7. Scarlat, N.; Dallemand, J.-F.; Monforti-Ferrario, F.; Nita, V. The role of biomass and bioenergy in a future bioeconomy: Policies and facts. Environ. Dev. 2015, 15, 3–34. [Google Scholar] [CrossRef]
  8. Den, W.; Sharma, V.K.; Lee, M.; Nadadur, G.; Varma, R.S. Lignocellulosic biomass transformations via greener oxidative pretreatment processes: Access to energy and value added chemicals. Front. Chem. 2018, 6, 141. [Google Scholar] [CrossRef] [Green Version]
  9. Sindhu, R.; Binod, P.; Pandey, A. Biological pretreatment of lignocellulosic biomass—An overview. Bioresour. Technol. 2016, 199, 76–82. [Google Scholar] [CrossRef]
  10. Singhvi, M.; Kim, B.S. Current developments in lignocellulosic biomass conversion into biofuels using nanobiotechology approach. Energies 2020, 13, 5300. [Google Scholar] [CrossRef]
  11. Sankaran, R.; Markandan, K.; Khoo, K.S.; Cheng, C.K.; Ashokkumar, V.; Deepanraj, B.; Show, P.L. The Expansion of Lignocellulose Biomass Conversion Into Bioenergy via Nanobiotechnology. Front. Nanotechnol. 2021, 3, 96. [Google Scholar] [CrossRef]
  12. Varma, R.S. Biomass-Derived Renewable Carbonaceous Materials for Sustainable Chemical and Environmental Applications. ACS Sustain. Chem. Eng. 2019, 7, 6458–6470. [Google Scholar] [CrossRef] [Green Version]
  13. Rai, A.K.; Al Makishah, N.H.; Wen, Z.; Gupta, G.; Pandit, S.; Prasad, R. Recent Developments in Lignocellulosic Biofuels, a Renewable Source of Bioenergy. Fermentation 2022, 8, 161. [Google Scholar] [CrossRef]
  14. Daystar, J.; Treasure, T.; Gonzalez, R.; Reeb, C.; Venditti, R.; Kelley, S. The NREL biochemical and thermochemical ethanol conversion processes: Financial and environmental analysis comparison. BioResources 2015, 10, 5083–5095. [Google Scholar] [CrossRef]
  15. Arora, A.; Nandal, P.; Singh, J.; Verma, M.L. Nanobiotechnological advancements in lignocellulosic biomass pretreatment. Mater. Sci. Energy Technol. 2020, 3, 308–318. [Google Scholar] [CrossRef]
  16. Rai, M.; Ingle, A.P.; Pandit, R.; Paralikar, P.; Biswas, J.K.; da Silva, S.S. Emerging role of nanobiocatalysts in hydrolysis of lignocellulosic biomass leading to sustainable bioethanol production. Catal. Rev. 2019, 61, 1–26. [Google Scholar] [CrossRef]
  17. Ali, S.; Shafique, O.; Mahmood, S.; Mahmood, T.; Khan, B.A.; Ahmad, I. Biofuels production from weed biomass using nanocatalyst technology. Biomass Bioenergy 2020, 139, 105595. [Google Scholar] [CrossRef]
  18. Srivastava, N.; Srivastava, M.; Mishra, P.K.; Singh, P.; Ramteke, P.W. Application of Cellulases in Biofuels Industries: An Overview. J. Biofuels Bioenergy 2015, 1, 55. [Google Scholar] [CrossRef]
  19. Rozina; Ahmad, M.; Asif, S.; Klemeš, J.J.; Mubashir, M.; Bokhari, A.; Sultana, S.; Mukhtar, A.; Zafar, M.; Bazmi, A.A.; et al. Conversion of the toxic and hazardous Zanthoxylum armatum seed oil into methyl ester using green and recyclable silver oxide nanoparticles. Fuel 2022, 310, 122296. [Google Scholar] [CrossRef]
  20. Rozina; Chia, S.R.; Ahmad, M.; Sultana, S.; Zafar, M.; Asif, S.; Bokhari, A.; Nomanbhay, S.; Mubashir, M.; Khoo, K.S.; et al. Green synthesis of biodiesel from Citrus medica seed oil using green nanoparticles of copper oxide. Fuel 2022, 323, 124285. [Google Scholar] [CrossRef]
  21. ISO/TS 27687:2008; Nanotechnologies—Terminology and Definitions for Nano-Objects—Nanoparticle, Nanofibre and Nanoplate. ISO: Geneva, Switzerland, 2009.
  22. Hamawand, I.; Seneweera, S.; Kumarasinghe, P.; Bundschuh, J. Nano-Structures & Nano-Objects Nanoparticle technology for separation of cellulose, hemicellulose and lignin nanoparticles from lignocellulose biomass: A short review. Nano-Struct. Nano-Objects 2020, 24, 100601. [Google Scholar] [CrossRef]
  23. Zhang, X.; Yan, S.; Tyagi, R.D.; Surampalli, R.Y. Synthesis of nanoparticles by microorganisms and their application in enhancing microbiological reaction rates. Chemosphere 2011, 82, 489–494. [Google Scholar] [CrossRef]
  24. Kumar, P.; Khan, M.A.; Hu, J.; Kibria, M.G. Single-atom catalysts for biomass-derived drop-in chemicals. In Advanced Catalysis for Drop-in Chemicals; Elsevier: Amsterdam, The Netherlands, 2022; pp. 63–100. [Google Scholar]
  25. Vasconcelos, S.C.; Marchini, L.; Lima, C.G.S.; Madriaga, V.G.C.; Ribeiro, R.S.A.; Rossa, V.; Ferreira, L.E.M.; de Carvalho da Silva, F.; Ferreira, V.F.; Passos, F.B.; et al. Single-atom catalysts for the upgrading of biomass-derived molecules: An overview of their preparation, properties and applications. Green Chem. 2022, 24, 2722–2751. [Google Scholar] [CrossRef]
  26. Singh, N.; Dhanya, B.S.; Verma, M.L. Nano-immobilized biocatalysts and their potential biotechnological applications in bioenergy production. Mater. Sci. Energy Technol. 2020, 3, 808–824. [Google Scholar] [CrossRef]
  27. Lai, H.K.; Chou, Y.Z.; Lee, M.H.; Lin, K.Y.A. Coordination polymer-derived cobalt nanoparticle-embedded carbon nanocomposite as a magnetic multi-functional catalyst for energy generation and biomass conversion. Chem. Eng. J. 2018, 332, 717–726. [Google Scholar] [CrossRef]
  28. Zanuso, E.; Ruiz, H.A.; Domingues, L.; Teixeira, J.A. Magnetic Nanoparticles as Support for Cellulase Immobilization Strategy for Enzymatic Hydrolysis Using Hydrothermally Pretreated Corn Cob Biomass. Bioenergy Res. 2022. [Google Scholar] [CrossRef]
  29. Baskar, G.; Naveen Kumar, R.; Heronimus Melvin, X.; Aiswarya, R.; Soumya, S. Sesbania aculeate biomass hydrolysis using magnetic nanobiocomposite of cellulase for bioethanol production. Renew. Energy 2016, 98, 23–28. [Google Scholar] [CrossRef]
  30. Wang, X.; Dou, P.; Zhao, P.; Zhao, C.; Ding, Y.; Xu, P. Immobilization of lipases onto magnetic Fe3O4 nanoparticles for application in biodiesel production. ChemSusChem 2009, 2, 947–950. [Google Scholar] [CrossRef]
  31. Zhong, R.; Sels, B.F. Sulfonated mesoporous carbon and silica-carbon nanocomposites for biomass conversion. Appl. Catal. B Environ. 2018, 236, 518–545. [Google Scholar] [CrossRef]
  32. Abraham, R.E.; Verma, M.L.; Barrow, C.J.; Puri, M. Suitability of magnetic nanoparticle immobilised cellulases in enhancing enzymatic saccharification of pretreated hemp biomass. Biotechnol. Biofuels 2014, 7, 90. [Google Scholar] [CrossRef] [Green Version]
  33. Bianco, A.; Kostarelos, K.; Prato, M. Making carbon nanotubes biocompatible and biodegradable. Chem. Commun. 2011, 47, 10182–10188. [Google Scholar] [CrossRef]
  34. Aoki, K.; Saito, N. Biocompatibility and Carcinogenicity of Carbon Nanotubes as Biomaterials. Nanomaterials 2020, 10, 264. [Google Scholar] [CrossRef]
  35. Rai, M.; César, J.; Soler, M.F.; Ricardo, P.; Marcelino, F.; Brumano, L.P.; Ingle, A.P. Strategic role of nanotechnology for production of bioethanol and biodiesel. Nanotechnol. Rev. 2016, 5, 231–250. [Google Scholar] [CrossRef]
  36. Lau, Z.L.; Low, S.S.; Ezeigwe, E.R.; Chew, K.W.; Chai, W.S.; Bhatnagar, A.; Yap, Y.J.; Show, P.L. A review on the diverse interactions between microalgae and nanomaterials: Growth variation, photosynthetic performance and toxicity. Bioresour. Technol. 2022, 351, 127048. [Google Scholar] [CrossRef]
  37. Ahmad, R.; Khare, S.K. Immobilization of Aspergillus niger cellulase on multiwall carbon nanotubes for cellulose hydrolysis. Bioresour. Technol. 2018, 252, 72–75. [Google Scholar] [CrossRef]
  38. Mubarak, N.M.; Wong, J.R.; Tan, K.W.; Sahu, J.N.; Abdullah, E.C.; Jayakumar, N.S.; Ganesan, P. Enzymatic Immobilization of cellulase enzyme on functionalized multiwall carbon nanotubes. J. Mol. Catal. B Enzym. 2014, 107, 124–131. [Google Scholar] [CrossRef]
  39. Pavlidis, B.I.V.; Tsoufis, T.; Enotiadis, A.; Gournis, D.; Stamatis, H. Functionalized Multi-Wall Carbon Nanotubes for Lipase Immobilization. Adv. Eng. Mater. 2010, 1, 179–183. [Google Scholar] [CrossRef]
  40. Jagtap, S.; Yenkie, M.K.N.; Labhsetwar, N.; Rayalu, S. Defluoridation of drinking water using chitosan based mesoporous alumina. Microporous Mesoporous Mater. 2011, 142, 454–463. [Google Scholar] [CrossRef]
  41. Chandra, D.; Mukherjee, N.; Mondal, A.; Bhaumik, A. Design and synthesis of nanostructured porous SnO2 with high surface areas and their optical and dielectric properties. J. Phys. Chem. C 2008, 112, 8668–8674. [Google Scholar] [CrossRef]
  42. Parida, K.M.; Sahu, N.; Mohapatra, P.; Scurrell, M.S. Low temperature CO oxidation over gold supported mesoporous Fe–TiO2. J. Mol. Catal. A Chem. 2010, 319, 92–97. [Google Scholar] [CrossRef]
  43. Dutta, S.; De, S.; Patra, A.K.; Sasidharan, M.; Bhaumik, A.; Saha, B. Applied Catalysis A: General Microwave assisted rapid conversion of carbohydrates into 5-hydroxymethylfurfural catalyzed by mesoporous TiO2 nanoparticles. Appl. Catal. A Gen. 2011, 409–410, 133–139. [Google Scholar] [CrossRef]
  44. De, S.; Dutta, S.; Patra, A.K.; Rana, B.S.; Sinha, A.K.; Saha, B.; Bhaumik, A. Biopolymer templated porous TiO2: An efficient catalyst for the conversion of unutilized sugars derived from hemicellulose. Appl. Catal. A Gen. 2012, 435–436, 197–203. [Google Scholar] [CrossRef]
  45. McClelland, D.J.; Galebach, P.H.; Motagamwala, A.H.; Wittrig, A.M.; Karlen, S.D.; Buchanan, J.S.; Dumesic, J.A.; Huber, G.W. Supercritical methanol depolymerization and hydrodeoxygenation of lignin and biomass over reduced copper porous metal oxides. Green Chem. 2019, 21, 2988–3005. [Google Scholar] [CrossRef]
  46. Liu, J.; Li, H.; Liu, Y.C.; Lu, Y.M.; He, J.; Liu, X.F.; Wu, Z.B.; Yang, S. Catalytic conversion of glucose to 5-hydroxymethylfurfural over nano-sized mesoporous Al2O3-B2O3 solid acids. Catal. Commun. 2015, 62, 19–23. [Google Scholar] [CrossRef]
  47. Gopinath, S.; Kumar, P.S.M.; Arafath, K.A.Y.; Thiruvengadaravi, K.V.; Sivanesan, S.; Baskaralingam, P. Efficient mesoporous SO42−/Zr-KIT-6 solid acid catalyst for green diesel production from esterification of oleic acid. Fuel 2017, 203, 488–500. [Google Scholar] [CrossRef]
  48. Luo, H.Y.; Lewis, J.D.; Román-Leshkov, Y. Lewis Acid Zeolites for Biomass Conversion: Perspectives and Challenges on Reactivity, Synthesis, and Stability. Annu. Rev. Chem. Biomol. Eng. 2016, 7, 663–692. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  49. Tang, F.; Li, L.; Chen, D. Mesoporous silica nanoparticles: Synthesis, biocompatibility and drug delivery. Adv. Mater. 2012, 24, 1504–1534. [Google Scholar] [CrossRef]
  50. Trisunaryanti, W.; Triyono; Fatmawati, D.A. Synthesis of Co-NH2/mesoporous silica bifunctional catalyst using sidoarjo mud and bovine bone gelatin template for conversion of used cooking oil into biofuel. Rasayan J. Chem. 2020, 13, 723–732. [Google Scholar] [CrossRef]
  51. Bhanja, P.; Bhaumik, A. Porous nanomaterials as green catalyst for the conversion of biomass to bioenergy. Fuel 2016, 185, 432–441. [Google Scholar] [CrossRef]
  52. Dufaud, V.; Davis, M.E. Design of heterogeneous catalysts via multiple active site positioning in organic-inorganic hybrid materials. J. Am. Chem. Soc. 2003, 125, 9403–9413. [Google Scholar] [CrossRef] [Green Version]
  53. Poorakbar, E.; Shafiee, A.; Saboury, A.A.; Rad, B.L.; Khoshnevisan, K.; Ma’mani, L.; Derakhshankhah, H.; Ganjali, M.R.; Hosseini, M. Synthesis of magnetic gold mesoporous silica nanoparticles core shell for cellulase enzyme immobilization: Improvement of enzymatic activity and thermal stability. Process Biochem. 2018, 71, 92–100. [Google Scholar] [CrossRef]
  54. Lee, Y.C.; Dutta, S.; Wu, K.C.W. Integrated, cascading enzyme-/chemocatalytic cellulose conversion using catalysts based on mesoporous silica nanoparticles. ChemSusChem 2014, 7, 3241–3246. [Google Scholar] [CrossRef]
  55. Budhi, S.; Mukarakate, C.; Iisa, K.; Pylypenko, S.; Ciesielski, P.N.; Yung, M.M.; Donohoe, B.S.; Katahira, R.; Nimlos, M.R.; Trewyn, B.G. Molybdenum incorporated mesoporous silica catalyst for production of biofuels and value-added chemicals via catalytic fast pyrolysis. Green Chem. 2015, 17, 3035–3046. [Google Scholar] [CrossRef]
  56. Ariaeenejad, S.; Jokar, F.; Hadian, P.; Ma’mani, L.; Gharaghani, S.; Fereidoonnezhad, M.; Salekdeh, G.H. An efficient nano-biocatalyst for lignocellulosic biomass hydrolysis: Xylanase immobilization on organically modified biogenic mesoporous silica nanoparticles. Int. J. Biol. Macromol. 2020, 164, 3462–3473. [Google Scholar] [CrossRef]
  57. Varkolu, M.; Velpula, V.; Ganji, S.; Burri, D.R.; Rao Kamaraju, S.R. Ni nanoparticles supported on mesoporous silica (2D, 3D) architectures: Highly efficient catalysts for the hydrocyclization of biomass-derived levulinic acid. RSC Adv. 2015, 5, 57201–57210. [Google Scholar] [CrossRef]
  58. Zhang, Y.; Degirmenci, V.; Li, C.; Hensen, E.J.M. Phosphotungstic acid encapsulated in metal-organic framework as catalysts for carbohydrate dehydration to 5-hydroxymethylfurfural. ChemSusChem 2011, 4, 59–64. [Google Scholar] [CrossRef]
  59. Chen, J.; Li, K.; Chen, L.; Liu, R.; Huang, X.; Ye, D. Conversion of fructose into 5-hydroxymethylfurfural catalyzed by recyclable sulfonic acid-functionalized metal–organic frameworks. Green Chem. 2014, 16, 2490–2499. [Google Scholar] [CrossRef]
  60. Chen, J.; Wang, S.; Huang, J.; Chen, L.; Ma, L.; Huang, X. Conversion of cellulose and cellobiose into sorbitol catalyzed by ruthenium supported on a polyoxometalate/metal-organic framework hybrid. ChemSusChem 2013, 6, 1545–1555. [Google Scholar] [CrossRef]
  61. Taylor, J.M.; Dawson, K.W.; Shimizu, G.K.H. A Water-Stable Metal–Organic Framework with Highly Acidic Pores for Proton-Conducting Applications. J. Am. Chem. Soc. 2013, 135, 1193–1196. [Google Scholar] [CrossRef]
  62. Corma, A.; García, H.; Llabrés i Xamena, F.X. Engineering Metal Organic Frameworks for Heterogeneous Catalysis. Chem. Rev. 2010, 110, 4606–4655. [Google Scholar] [CrossRef]
  63. Stöcker, M. Biofuels and biomass-to-liquid fuels in the biorefinery: Catalytic conversion of lignocellulosic biomass using porous materials. Angew. Chem. Int. Ed. 2008, 47, 9200–9211. [Google Scholar] [CrossRef] [PubMed]
  64. Srinivasan, V.; Adhikari, S.; Chattanathan, S.A.; Tu, M.; Park, S. Catalytic Pyrolysis of Raw and Thermally Treated Cellulose Using Different Acidic Zeolites. Bioenergy Res. 2014, 7, 867–875. [Google Scholar] [CrossRef]
  65. Vichaphund, S.; Aht-Ong, D.; Sricharoenchaikul, V.; Atong, D. Production of aromatic compounds from catalytic fast pyrolysis of Jatropha residues using metal/HZSM-5 prepared by ion-exchange and impregnation methods. Renew. Energy 2015, 79, 28–37. [Google Scholar] [CrossRef] [Green Version]
  66. Zhou, N.; Liu, S.; Zhang, Y.; Fan, L.; Cheng, Y.; Wang, Y.; Liu, Y.; Chen, P.; Ruan, R. Silicon carbide foam supported ZSM-5 composite catalyst for microwave-assisted pyrolysis of biomass. Bioresour. Technol. 2018, 267, 257–264. [Google Scholar] [CrossRef] [PubMed]
  67. Jorge, E.Y.C.; Lima, T.D.M.; Lima, C.G.S.; Marchini, L.; Castelblanco, W.N.; Rivera, D.G.; Urquieta-González, E.A.; Varma, R.S.; Paixão, M.W. Metal-exchanged magnetic β-zeolites: Valorization of lignocellulosic biomass-derived compounds to platform chemicals. Green Chem. 2017, 19, 3856–3868. [Google Scholar] [CrossRef]
  68. Xu, Q.; Li, W.; Ma, L.; Cao, D.; Owens, G.; Chen, Z. Simultaneous removal of ammonia and phosphate using green synthesized iron oxide nanoparticles dispersed onto zeolite. Sci. Total Environ. 2020, 703, 135002. [Google Scholar] [CrossRef]
  69. Misson, M.; Zhang, H.; Jin, B. Nanobiocatalyst advancements and bioprocessing applications. J. R. Soc. Interface 2015, 12, 20140891. [Google Scholar] [CrossRef] [Green Version]
  70. Su, T.C.; Fang, Z.; Zhang, F.; Luo, J.; Li, X.K. Hydrolysis of Selected Tropical Plant Wastes Catalyzed by a Magnetic Carbonaceous Acid with Microwave. Sci. Rep. 2015, 5, 17538. [Google Scholar] [CrossRef] [Green Version]
  71. Goh, W.J.; Makam, V.S.; Hu, J.; Kang, L.; Zheng, M.; Yoong, S.L.; Udalagama, C.N.B.; Pastorin, G. Iron oxide filled magnetic carbon nanotube-enzyme conjugates for recycling of amyloglucosidase: Toward useful applications in biofuel production process. Langmuir 2012, 28, 16864–16873. [Google Scholar] [CrossRef]
  72. Verma, M.L.; Chaudhary, R.; Tsuzuki, T.; Barrow, C.J.; Puri, M. Immobilization of β-glucosidase on a magnetic nanoparticle improves thermostability: Application in cellobiose hydrolysis. Bioresour. Technol. 2013, 135, 2–6. [Google Scholar] [CrossRef]
  73. Ren, Y.; Rivera, J.G.; He, L.; Kulkarni, H.; Lee, D.K.; Messersmith, P.B. Facile, high efficiency immobilization of lipase enzyme on magnetic iron oxide nanoparticles via a biomimetic coating. BMC Biotechnol. 2011, 11, 63. [Google Scholar] [CrossRef] [Green Version]
  74. Hu, S.; Guan, Y.; Wang, Y.; Han, H. Nano-magnetic catalyst KF/CaO-Fe3O4 for biodiesel production. Appl. Energy 2011, 88, 2685–2690. [Google Scholar] [CrossRef]
  75. Sanusi, I.A.; Suinyuy, T.N.; Kana, G.E.B. Impact of nanoparticle inclusion on bioethanol production process kinetic and inhibitor profile. Biotechnol. Rep. 2021, 29, e00585. [Google Scholar] [CrossRef]
  76. Vijayalakshmi, S.; Govindarajan, M.; Al-Mulahim, N.; Ahmed, Z.; Mahboob, S. Cellulase immobilized magnetic nanoparticles for green energy production from Allamanda schottii L: Sustainability research in waste recycling. Saudi J. Biol. Sci. 2021, 28, 901–910. [Google Scholar] [CrossRef]
  77. Cherian, E.; Dharmendirakumar, M.; Baskar, G. Immobilization of cellulase onto MnO2 nanoparticles for bioethanol production by enhanced hydrolysis of agricultural waste. Cuihua Xuebao/Chin. J. Catal. 2015, 36, 1223–1229. [Google Scholar] [CrossRef]
  78. Banerjee, S.; Rout, S.; Banerjee, S.; Atta, A.; Das, D. Fe2O3 nanocatalyst aided transesterification for biodiesel production from lipid-intact wet microalgal biomass: A biorefinery approach. Energy Convers. Manag. 2019, 195, 844–853. [Google Scholar] [CrossRef]
  79. Xie, W.; Ma, N. Immobilized Lipase on Fe 3 O 4 Nanoparticles as Biocatalyst for Biodiesel Production. Energy Fuels 2009, 23, 1347–1353. [Google Scholar] [CrossRef]
  80. Wang, X.; Liu, X.; Zhao, C.; Ding, Y.; Xu, P. Biodiesel production in packed-bed reactors using lipase-nanoparticle biocomposite. Bioresour. Technol. 2011, 102, 6352–6355. [Google Scholar] [CrossRef]
  81. Chiang, Y.D.; Dutta, S.; Chen, C.T.; Huang, Y.T.; Lin, K.S.; Wu, J.C.S.; Suzuki, N.; Yamauchi, Y.; Wu, K.C.W. Functionalized Fe3O4@Silica Core-Shell Nanoparticles as Microalgae Harvester and Catalyst for Biodiesel Production. ChemSusChem 2015, 8, 789–794. [Google Scholar] [CrossRef]
  82. Zhou, Z.W.; Cai, C.X.; Xing, X.; Li, J.; Hu, Z.E.; Xie, Z.B.; Wang, N.; Yu, X.Q. Magnetic COFs as satisfied support for lipase immobilization and recovery to effectively achieve the production of biodiesel by great maintenance of enzyme activity. Biotechnol. Biofuels 2021, 14, 156. [Google Scholar] [CrossRef]
  83. Zhang, H.; Liu, T.; Zhu, Y.; Hong, L.; Li, T.; Wang, X.; Fu, Y. Lipases immobilized on the modified polyporous magnetic cellulose support as an efficient and recyclable catalyst for biodiesel production from Yellow horn seed oil. Renew. Energy 2020, 145, 1246–1254. [Google Scholar] [CrossRef]
  84. Abdelsalam, E.; Samer, M.; Attia, Y.A.; Abdel-Hadi, M.A.; Hassan, H.E.; Badr, Y. Influence of zero valent iron nanoparticles and magnetic iron oxide nanoparticles on biogas and methane production from anaerobic digestion of manure. Energy 2017, 120, 842–853. [Google Scholar] [CrossRef]
  85. Velimirovic, M.; Schmid, D.; Wagner, S.; Micić, V.; von der Kammer, F.; Hofmann, T. Agar agar-stabilized milled zerovalent iron particles for in situ groundwater remediation. Sci. Total Environ. 2016, 563–564, 713–723. [Google Scholar] [CrossRef] [PubMed]
  86. Sharma, P.; Melkania, U. Biochar-enhanced hydrogen production from organic fraction of municipal solid waste using co-culture of Enterobacter aerogenes and E. coli. Int. J. Hydrogen Energy 2017, 42, 18865–18874. [Google Scholar] [CrossRef]
  87. Sunyoto, N.M.S.; Zhu, M.; Zhang, Z.; Zhang, D. Effect of biochar addition on hydrogen and methane production in two-phase anaerobic digestion of aqueous carbohydrates food waste. Bioresour. Technol. 2016, 219, 29–36. [Google Scholar] [CrossRef] [PubMed]
  88. Yang, G.; Wang, J. Synergistic enhancement of biohydrogen production from grass fermentation using biochar combined with zero-valent iron nanoparticles. Fuel 2019, 251, 420–427. [Google Scholar] [CrossRef]
  89. Su, L.; Shi, X.; Guo, G.; Zhao, A.; Zhao, Y. Stabilization of sewage sludge in the presence of nanoscale zero-valent iron (nZVI): Abatement of odor and improvement of biogas production. J. Mater. Cycles Waste Manag. 2013, 15, 461–468. [Google Scholar] [CrossRef]
  90. Karri, S.; Sierra-Alvarez, R.; Field, J.A. Zero valent iron as an electron-donor for methanogenesis and sulfate reduction in anaerobic sludge. Biotechnol. Bioeng. 2005, 92, 810–819. [Google Scholar] [CrossRef]
  91. Wang, T.; Zhang, D.; Dai, L.; Chen, Y.; Dai, X. Effects of metal nanoparticles on methane production from waste-activated sludge and microorganism community shift in anaerobic granular sludge. Sci. Rep. 2016, 6, 25857. [Google Scholar] [CrossRef]
  92. Yang, Y.; Guo, J.; Hu, Z. Impact of nano zero valent iron (NZVI) on methanogenic activity and population dynamics in anaerobic digestion. Water Res. 2013, 47, 6790–6800. [Google Scholar] [CrossRef]
  93. Zhang, Y.; Shen, J. Enhancement effect of gold nanoparticles on biohydrogen production from artificial wastewater. Int. J. Hydrogen Energy 2007, 32, 17–23. [Google Scholar] [CrossRef]
  94. Ramsurn, H.; Gupta, R.B. Nanotechnology in solar and biofuels. ACS Sustain. Chem. Eng. 2013, 1, 779–797. [Google Scholar] [CrossRef]
  95. Gordon, R.; Seckbach, J. (Eds.) The Science of Algal Fuels: Phycology, Geology, Biophotonics, Genomics and Nanotechnology; Springer Science & Business Media: Berlin/Heidelberg, Germany, 2012; ISBN 978-94-007-5110-1. [Google Scholar]
  96. Yang, G.; Wang, J. Improving mechanisms of biohydrogen production from grass using zero- valent iron nanoparticles. Bioresour. Technol. 2018, 266, 413–420. [Google Scholar] [CrossRef]
  97. Beckers, L.; Hiligsmann, S.; Lambert, S.D.; Heinrichs, B.; Thonart, P. Improving effect of metal and oxide nanoparticles encapsulated in porous silica on fermentative biohydrogen production by Clostridium butyricum. Bioresour. Technol. 2013, 133, 109–117. [Google Scholar] [CrossRef]
  98. Elreedy, A.; Ibrahim, E.; Hassan, N.; El-Dissouky, A.; Fujii, M.; Yoshimura, C.; Tawfik, A. Nickel-graphene nanocomposite as a novel supplement for enhancement of biohydrogen production from industrial wastewater containing mono-ethylene glycol. Energy Convers. Manag. 2017, 140, 133–144. [Google Scholar] [CrossRef]
  99. Raliya, R.; Biswas, P.; Tarafdar, J.C. TiO2 nanoparticle biosynthesis and its physiological effect on mung bean (Vigna radiata L.). Biotechnol. Rep. 2015, 5, 22–26. [Google Scholar] [CrossRef] [Green Version]
  100. Pádrová, K.; Lukavský, J.; Nedbalová, L.; Čejková, A.; Cajthaml, T.; Sigler, K.; Vítová, M.; Řezanka, T. Trace concentrations of iron nanoparticles cause overproduction of biomass and lipids during cultivation of cyanobacteria and microalgae. J. Appl. Phycol. 2015, 27, 1443–1451. [Google Scholar] [CrossRef]
  101. Giannelli, L.; Torzillo, G. Hydrogen production with the microalga Chlamydomonas reinhardtii grown in a compact tubular photobioreactor immersed in a scattering light nanoparticle suspension. Int. J. Hydrogen Energy 2012, 37, 16951–16961. [Google Scholar] [CrossRef]
  102. Eroglu, E.; Eggers, P.K.; Winslade, M.; Smith, S.M.; Raston, C.L. Enhanced accumulation of microalgal pigments using metal nanoparticle solutions as light filtering devices. Green Chem. 2013, 15, 3155–3159. [Google Scholar] [CrossRef]
  103. Pandey, A.; Gupta, K.; Pandey, A. Effect of nanosized TiO2 on photofermentation by Rhodobacter sphaeroides NMBL-02. Biomass Bioenergy 2015, 72, 273–279. [Google Scholar] [CrossRef]
  104. Yang, F.; Hong, F.; You, W.; Liu, C.; Gao, F.; Wu, C.; Yang, P. Influences of nano-anatase TiO2 on the nitrogen metabolism of growing spinach. Biol. Trace Elem. Res. 2006, 110, 179–190. [Google Scholar] [CrossRef]
  105. Mishra, A.; Kumari, M.; Pandey, S.; Chaudhry, V.; Gupta, K.C.; Nautiyal, C.S. Biocatalytic and antimicrobial activities of gold nanoparticles synthesized by Trichoderma sp. Bioresour. Technol. 2014, 166, 235–242. [Google Scholar] [CrossRef]
  106. Yates, M.D.; Cusick, R.D.; Logan, B.E. Extracellular palladium nanoparticle production using geobacter sulfurreducens. ACS Sustain. Chem. Eng. 2013, 1, 1165–1171. [Google Scholar] [CrossRef]
  107. Hakamizadeh, M.; Afshar, S.; Tadjarodi, A.; Khajavian, R.; Fadaie, M.R.; Bozorgi, B. Improving hydrogen production via water splitting over Pt/TiO2/activated carbon nanocomposite. Int. J. Hydrogen Energy 2014, 39, 7262–7269. [Google Scholar] [CrossRef]
  108. Markowska-Szczupak, A.; Wang, K.; Rokicka, P.; Endo, M.; Wei, Z.; Ohtani, B.; Morawski, A.W.; Kowalska, E. The effect of anatase and rutile crystallites isolated from titania P25 photocatalyst on growth of selected mould fungi. J. Photochem. Photobiol. B Biol. 2015, 151, 54–62. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  109. Cheng, P.; Yang, Z.; Wang, H.; Cheng, W.; Chen, M.; Shangguan, W.; Ding, G. TiO2-graphene nanocomposites for photocatalytic hydrogen production from splitting water. Int. J. Hydrogen Energy 2012, 37, 2224–2230. [Google Scholar] [CrossRef]
  110. Sekoai, P.T.; Ouma, C.N.M.; du Preez, S.P.; Modisha, P.; Engelbrecht, N.; Bessarabov, D.G.; Ghimire, A. Application of nanoparticles in biofuels: An overview. Fuel 2019, 237, 380–397. [Google Scholar] [CrossRef]
  111. Zhao, W.; Zhang, Y.; Du, B.; Wei, D.; Wei, Q.; Zhao, Y. Enhancement effect of silver nanoparticles on fermentative biohydrogen production using mixed bacteria. Bioresour. Technol. 2013, 142, 240–245. [Google Scholar] [CrossRef]
  112. Mohanraj, S.; Anbalagan, K.; Kodhaiyolii, S.; Pugalenthi, V. Comparative evaluation of fermentative hydrogen production using Enterobacter cloacae and mixed culture: Effect of Pd (II) ion and phytogenic palladium nanoparticles. J. Biotechnol. 2014, 192, 87–95. [Google Scholar] [CrossRef]
  113. Mullai, P.; Yogeswari, M.K.; Sridevi, K. Optimisation and enhancement of biohydrogen production using nickel nanoparticles—A novel approach. Bioresour. Technol. 2013, 141, 212–219. [Google Scholar] [CrossRef]
  114. Mohanraj, S.; Anbalagan, K.; Rajaguru, P.; Pugalenthi, V. Effects of phytogenic copper nanoparticles on fermentative hydrogen production by Enterobacter cloacae and Clostridium acetobutylicum. Int. J. Hydrogen Energy 2016, 41, 10639–10645. [Google Scholar] [CrossRef]
  115. Nath, D.; Manhar, A.K.; Gupta, K.; Saikia, D.; Das, S.K.; Mandal, M. Phytosynthesized iron nanoparticles: Effects on fermentative hydrogen production by Enterobacter cloacae DH-89. Bull. Mater. Sci. 2015, 38, 1533–1538. [Google Scholar] [CrossRef]
  116. Dolly, S.; Pandey, A.; Pandey, B.K.; Gopal, R. Process parameter optimization and enhancement of photo-biohydrogen production by mixed culture of Rhodobacter sphaeroides NMBL-02 and Escherichia coli NMBL-04 using Fe-nanoparticle. Int. J. Hydrogen Energy 2015, 40, 16010–16020. [Google Scholar] [CrossRef]
  117. Lin, R.; Cheng, J.; Ding, L.; Song, W.; Liu, M.; Zhou, J.; Cen, K. Enhanced dark hydrogen fermentation by addition of ferric oxide nanoparticles using Enterobacter aerogenes. Bioresour. Technol. 2016, 207, 213–219. [Google Scholar] [CrossRef]
  118. Malik, S.N.; Pugalenthi, V.; Vaidya, A.N.; Ghosh, P.C.; Mudliar, S.N. Kinetics of nano-catalysed dark fermentative hydrogen production from distillery wastewater. Energy Procedia 2014, 54, 417–430. [Google Scholar] [CrossRef] [Green Version]
  119. Mohanraj, S.; Kodhaiyolii, S.; Rengasamy, M.; Pugalenthi, V. Green synthesized iron oxide nanoparticles effect on fermentative hydrogen production by Clostridium acetobutylicum. Appl. Biochem. Biotechnol. 2014, 173, 318–331. [Google Scholar] [CrossRef]
  120. Engliman, N.S.; Abdul, P.M.; Wu, S.Y.; Jahim, J.M. Influence of iron (II) oxide nanoparticle on biohydrogen production in thermophilic mixed fermentation. Int. J. Hydrogen Energy 2017, 42, 27482–27493. [Google Scholar] [CrossRef]
  121. Zhao, Y.; Chen, Y. Nano-TiO2 enhanced photofermentative hydrogen produced from the dark fermentation liquid of waste activated sludge. Environ. Sci. Technol. 2011, 45, 8589–8595. [Google Scholar] [CrossRef]
  122. Liu, Z.; Lv, F.; Zheng, H.; Zhang, C.; Wei, F.; Xing, X.H. Enhanced hydrogen production in a UASB reactor by retaining microbial consortium onto carbon nanotubes (CNTs). Int. J. Hydrogen Energy 2012, 37, 10619–10626. [Google Scholar] [CrossRef]
  123. Hsieh, P.H.; Lai, Y.C.; Chen, K.Y.; Hung, C.H. Explore the possible effect of TiO2 and magnetic hematite nanoparticle addition on biohydrogen production by Clostridium pasteurianum based on gene expression measurements. Int. J. Hydrogen Energy 2016, 41, 21685–21691. [Google Scholar] [CrossRef]
  124. Saeed, S.; Samer, M.; Mohamed, M.S.M.; Abdelsalam, E.; Mohamed, Y.M.A.; Abdel-Hafez, S.H.; Attia, Y.A. Implementation of graphitic carbon nitride nanomaterials and laser irradiation for increasing bioethanol production from potato processing wastes. Environ. Sci. Pollut. Res. 2022, 29, 34887–34897. [Google Scholar] [CrossRef]
  125. Gupta, K.; Chundawat, T.S. Zinc oxide nanoparticles synthesized using Fusarium oxysporum to enhance bioethanol production from rice-straw. Biomass Bioenergy 2020, 143, 105840. [Google Scholar] [CrossRef]
  126. Ivanova, V.; Petrova, P.; Hristov, J. Application in the Ethanol Fermentation of Immobilized Yeast Cells in Matrix of Alginate/Magnetic Nanoparticles, on Chitosan-Magnetite Microparticles and Cellulose-coated Magnetic Nanoparticles. Int. Rev. Chem. Eng. 2011, 3, 289–299. [Google Scholar] [CrossRef]
  127. Kim, Y.K.; Park, S.E.; Lee, H.; Yun, J.Y. Enhancement of bioethanol production in syngas fermentation with Clostridium ljungdahlii using nanoparticles. Bioresour. Technol. 2014, 159, 446–450. [Google Scholar] [CrossRef]
  128. Prabu, A. Nanoparticles as additive in biodiesel on the working characteristics of a DI diesel engine. Ain Shams Eng. J. 2018, 9, 2343–2349. [Google Scholar] [CrossRef]
Figure 1. Lignocellulosic biomass as bioenergy crop; plant cell wall is made of lignocellulose containing carbohydrate polymers (cellulose, hemicellulose and aromatic polymer lignin). The cellulose and hemicellulose accommodate 6 and 5 carbon sugars, which are rigidly bounded with the lignin.
Figure 1. Lignocellulosic biomass as bioenergy crop; plant cell wall is made of lignocellulose containing carbohydrate polymers (cellulose, hemicellulose and aromatic polymer lignin). The cellulose and hemicellulose accommodate 6 and 5 carbon sugars, which are rigidly bounded with the lignin.
Materials 15 07769 g001
Figure 2. Schematic illustration describing the use of magnetic nanoparticles (MNPs) immobilized cellulase enzyme in hydrolysis process of lignocellulosic biomass for biofuel production. Reproduced from [32] with permission from Springer Nature.
Figure 2. Schematic illustration describing the use of magnetic nanoparticles (MNPs) immobilized cellulase enzyme in hydrolysis process of lignocellulosic biomass for biofuel production. Reproduced from [32] with permission from Springer Nature.
Materials 15 07769 g002
Figure 3. (A,B) Schematic illustration of immobilized enzyme; cellulase (CEL) immobilized glutamic-acid-functionalized magnetic gold mesoporous silica nanoparticle core shells (CEL@Glu@PEGylated mAu@PSN). (C,D) TEM/SEM images of Glu@PEGylated mAu@PSNs. (E) Change in enzymatic activity of free and immobilized cellulase with time at pH = 4.8 and 50 °C. Reproduced from [53] with permission from Elsevier.
Figure 3. (A,B) Schematic illustration of immobilized enzyme; cellulase (CEL) immobilized glutamic-acid-functionalized magnetic gold mesoporous silica nanoparticle core shells (CEL@Glu@PEGylated mAu@PSN). (C,D) TEM/SEM images of Glu@PEGylated mAu@PSNs. (E) Change in enzymatic activity of free and immobilized cellulase with time at pH = 4.8 and 50 °C. Reproduced from [53] with permission from Elsevier.
Materials 15 07769 g003
Figure 4. (AC) Structure of sulfonic-acid-functionalized MOF as catalyst: MIL-101(Cr)-SO3H for conversion of fructose into HMF, the fructose conversion over time and HMF yield against time. (D) SEM image of catalyst prepared by adsorbing phosphotungstic acid (PTA) in MIL-101 MOF, (E) yield and selectivity over time for fructose dehydration using (PTA)-MIL-101 in 1-ethyl-3-methlyimidazolium chloride at 80 °C; square, diamond and triangular points indicate fructose retention, HMF yield and HMF selectivity, respectively (F) yield and selectivity over time for fructose dehydration using (PTA)-MIL-101 in DMSO at 130;square, circular and triangular points indicate fructose retention, HMF yield and HMF selectivity, respectively. Reproduced from [58,59] with permission from John Wiley and Sons and the Royal Society of Chemistry.
Figure 4. (AC) Structure of sulfonic-acid-functionalized MOF as catalyst: MIL-101(Cr)-SO3H for conversion of fructose into HMF, the fructose conversion over time and HMF yield against time. (D) SEM image of catalyst prepared by adsorbing phosphotungstic acid (PTA) in MIL-101 MOF, (E) yield and selectivity over time for fructose dehydration using (PTA)-MIL-101 in 1-ethyl-3-methlyimidazolium chloride at 80 °C; square, diamond and triangular points indicate fructose retention, HMF yield and HMF selectivity, respectively (F) yield and selectivity over time for fructose dehydration using (PTA)-MIL-101 in DMSO at 130;square, circular and triangular points indicate fructose retention, HMF yield and HMF selectivity, respectively. Reproduced from [58,59] with permission from John Wiley and Sons and the Royal Society of Chemistry.
Materials 15 07769 g004
Figure 6. (A) SEM image of prepared g-C3N4 nanoparticles used to increase bioethanol production. (B) Irradiation using blue and red laser to increase bioethanol production from potato waste. (C) Effect of ZnO nanoparticles on bioethanol yield from rice straw. (D) Effect of various processing conditions of g-C3N4 nanoparticles on the production of bioethanol. Reproduced with permission from [124] Springer Nature and [125] Elsevier.
Figure 6. (A) SEM image of prepared g-C3N4 nanoparticles used to increase bioethanol production. (B) Irradiation using blue and red laser to increase bioethanol production from potato waste. (C) Effect of ZnO nanoparticles on bioethanol yield from rice straw. (D) Effect of various processing conditions of g-C3N4 nanoparticles on the production of bioethanol. Reproduced with permission from [124] Springer Nature and [125] Elsevier.
Materials 15 07769 g006
Table 1. Summary of various MNPs explored as support carriers in enzyme immobilization for bioenergy generation.
Table 1. Summary of various MNPs explored as support carriers in enzyme immobilization for bioenergy generation.
BioenergyMNPsFeedstockEnzymeSummaryRef.
BioethanolFe3O4Aspergillus fumigatus AA001CellulaseThermal stability for 8 h at 70 °C[18]
Fe3O4Potato peelsAmylase/AmyloglucosidaseBioethanol yield: 93%[75]
Fe3O4Allamanda schottii L.CellulaseBioethanol yield (free enzymes): 182 g/L; (immobilized enzymes): 252 g/L[76]
FeCl3Sesbania aculeataCellulaseBioethanol yield: 5.31 g/L[29]
MnO2Aspergillus fumigatus JCFCellulaseCellulase binding efficiency of 75%
Bioethanol yield: 21.96 g/L
[77]
BiodieselFe2O3Neochloris oleoabundansLipaseMaximum biodiesel yield: 81%[78]
Fe3O4Soybean oilLipaseBiodiesel yield: 90% with 60% immobilized lipase[79]
Fe3O4Soybean oilLipaseConversion rate: 47.60% after 24 h; >30% enzyme activity even after 10 cycles[30]
Fe3O4Soybean oilLipaseConversion rate: 88% after 192 h and 75% after 240 h[80]
Fe3O4Microalgae (Chlorella vulgaris)LipaseMaximum biodiesel yield: 97.1%[81]
Fe3O4Jatropha curcas oilRhizomucor miehei lipaseMaximum biodiesel yield: 70%[82]
Polyporous magnetic cellulose beads (PMCBs)Yellow horn seed oilCandida antarctica lipase BMaximum biodiesel yield: 92.3%; catalyst easily removed by magnet and can be recycled at least 5 times.[83]
Table 2. Effect of various nanoparticles on the H2 yield rate. Modified from [110].
Table 2. Effect of various nanoparticles on the H2 yield rate. Modified from [110].
NPFeedstockMicroorganismH2 Yield RateH2 Yield Increase (%)Ref.
AuArtificial wastewaterClostridium butyricum4.48 mol H2/mol sucrose61.7[93]
AgInorganic saltsClostridium butyricum2.48 mol H2/mol glucose67.5[111]
PdGlucoseEnterobacter cloacae + mixed culture2.48 mol H2/mol glucose6.4[112]
NiInorganic saltsGranular sludge2.54 mol H2/mol glucose22.7[113]
CuGlucoseClostridium acetobutylicum1.74 mol H2/mol glucoseN/A[114]
CuGlucoseEnterobacter clocae1.44 mol H2/mol glucoseN/A[114]
FeInorganic saltsEnterobacter clocae1.9 mol H2/mol glucose68.4[115]
FeGrowth mediumRhodobacter sphaeroides + Escherichia coli3.1 mol H2/mol malateN/A[116]
Fe2O3Casava starchEnterobacter aerogenes192.4 mL H2/g casava starch17[117]
Fe2O3Distillery wastewaterMixed culture44.28 mL H2/g CODN/A[118]
Fe2O3Growth mediumClostridium acetobutylicum2.33 mol H2/mol glucose52[119]
FeOGrowth mediumMixed culture1.92 mol H2/mol glucose7.9[120]
TiO2Growth mediumRhodopseudomonas palustrisN/A46.1[121]
SiO2Air: CO2 (97:3)Chlamydomonas reinhardtii CC1243121.5 H2/L/h45.2[101]
BiocharMunicipal solid wasteEnterobacter aerogenes + Escherichia coli96.3 mL/gN/A[86]
BiocharFood wasteN/A944.5 mL/L31[87]
Biochar + ZV iron NPGrass biomassN/A50.6 mL/g dry grass89.8[88]
Ni + grapheneIndustrial wastewaterMixed culture41.3 mL/gCOD105[98]
CNTGlucoseAnaerobic sludge2.45 mol/mol substrateN/A[122]
Hematite + TiO2GlucoseClostridium pasteurianum CH52.20 mol/mol substrate5[123]
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Markandan, K.; Chai, W.S. Perspectives on Nanomaterials and Nanotechnology for Sustainable Bioenergy Generation. Materials 2022, 15, 7769. https://doi.org/10.3390/ma15217769

AMA Style

Markandan K, Chai WS. Perspectives on Nanomaterials and Nanotechnology for Sustainable Bioenergy Generation. Materials. 2022; 15(21):7769. https://doi.org/10.3390/ma15217769

Chicago/Turabian Style

Markandan, Kalaimani, and Wai Siong Chai. 2022. "Perspectives on Nanomaterials and Nanotechnology for Sustainable Bioenergy Generation" Materials 15, no. 21: 7769. https://doi.org/10.3390/ma15217769

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

Article Metrics

Back to TopTop