Next Article in Journal
Electrochemical Jet Machining of Surface Texture: Improving the Strength of Hot-Pressure-Welded AA6061-CF/PA66 Joints
Next Article in Special Issue
Minimizing Porosity in 17-4 PH Stainless Steel Compacts in a Modified Powder Metallurgical Process
Previous Article in Journal
Effects of Ionizing Radiation on the Shear Bond Strength of Composite Materials to Dentin
Previous Article in Special Issue
Dynamic FEA Analysis of the Super Lightweight External Cryogenic Fuel Tank (SLWT) Made of Aluminium Alloy 2195–Graphene Nano Composite for Launch Vehicle Aerospace Application
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Hydrolyzed Forms of Cellulose and Its Metal Composites for Hydrogen Generation: An Experimental and Theoretical Investigation

by
Omar Faye
1,*,
Inimfon A. Udoetok
2,*,
Jerzy A. Szpunar
1 and
Lee D. Wilson
2
1
Department of Mechanical Engineering, College of Engineering, University of Saskatchewan, 57 Campus Drive, Saskatoon, SK S7N 5A9, Canada
2
Department of Chemistry, University of Saskatchewan, 110 Science Place–Rm 165 Thorvaldson Building, Saskatoon, SK S7N 5C9, Canada
*
Authors to whom correspondence should be addressed.
J. Compos. Sci. 2024, 8(7), 262; https://doi.org/10.3390/jcs8070262
Submission received: 7 May 2024 / Revised: 13 June 2024 / Accepted: 1 July 2024 / Published: 6 July 2024
(This article belongs to the Special Issue Metal Composites, Volume II)

Abstract

:
The quest for a smooth transition from fossil fuels to clean and sustainable energy has warranted studies on alternative energy materials. Herein, we report on an experimental and theoretical study focused on hydrogen generation through the hydrolysis of microcrystalline cellulose (MCC) treated in different media (deionized water, sodium hydroxide) and MCC functionalized with magnesium (MCC-Mg), titanium (MCC-Ti), and niobium (MCC-Nb). The XRD results reveal the decreased crystallinity of MCC due to ball milling along with the formation of metal oxide composites between MCC and various metals (magnesium, titanium, and niobium). Theoretical studies using NVT molecular dynamic simulations with the NH chain thermostat implemented in the Dmol3 provides further support to the experimental results reported herein. The results from the experimental and theoretical studies revealed that ball milling and composite formation with metal species enhanced the kinetics of the hydrolysis of MCC and, consequently, hydrogen generation, while the addition of NaOH and urea inhibited the hydrogen yield.

Graphical Abstract

1. Introduction

Biomass is a term that has several contextual meanings, from ecology to bioenergy, etc. In terms of bioenergy, it may be described as all organic matter of plant or animal origin that exists in the biosphere, including materials obtained through natural or artificial transformations [1,2,3,4,5,6,7,8,9,10,11]. Recently, the use of biomass materials as a source of clean and sustainable energy has increased exponentially due to their unique properties such as abundance, benign nature, degradability, and relatively low cost. Biomass is a renewable organic resource derived from crop residues, forest residues, and solid animal waste. Cellulose is the major component of woody biomass and the most abundant natural polysaccharide material on earth [2]. It is a linear homopolymer with D-glucose as its basic molecular building block. Due to its low cost, biocompatibility, amenability to chemical modification, and good mechanical properties [3,4], cellulose has attracted a growing interest in biological applications [5,6], water treatment [6,7,8], the preparation of sensors [10], mechanical reinforcement agents [11,12,13], and energy applications [14,15,16].
The smooth transition from fossil fuels to renewable feedstock requires the development of biomass-based materials that can be used as clean and sustainable energy sources. Among these renewable feedstocks, hydrogen is regarded as a valuable potential energy carrier for the future due to its high energy value (141.86 kJ g−1) compared to gasoline (47.50 kJ g−1), propane (50.36 kJ g−1), and methane (55.53 kJ g−1) [17]. In addition, hydrogen is non-toxic, and its byproduct, when used as a source of clean energy, is water [18]. However, hydrogen is not readily available in nature and requires sustainable and low-cost strategies for its generation. Therefore, various hydrogen generation methods have been reported that involve the use of different materials, such as single-component and composite materials [19,20]. A review on the use of cellulosic biomass for hydrogen generation by Mohanty et al. [21] reveals the application of thermochemical processes such as pyrolysis, catalytic pyrolysis, gasification, co-gasification, and supercritical water gasification (SCWG), along with biochemical/biological conversions, including electrolytic and photolytic processes.
One of the feasible methods of hydrogen generation is the hydrolysis of alkaline earth metals [22,23] and their composites with various biomaterials, such as graphene oxide/graphite [24,25] and hydrides [26,27,28]. However, using direct hydrolysis as a method of hydrogen production remains challenging due to the low yield of hydrogen. Similarly, the direct hydrolysis of cellulosic biomass for hydrogen generation requires further research to optimize the process for sustainable and cost-effective hydrogen generation. Zagrodnik et al. [29] studied the effect of heat pretreatment of inoculums on hydrogen yield from untreated cellulose and starch. Their study reported that pretreatment at 100 °C enhanced the generation of hydrogen by cellulose up to 0.48 mol H2/molhexose [29]. Also, Caravaca et al. [30] reported that cellulose could be reformed via sunlight to generate hydrogen by utilizing metal-coated titania photocatalysts. They also noted that precious metals like Pt, Pd, and Au are less environmentally friendly than Ni [30]. Gadow et al. [31] studied the effects of temperature on the generation of hydrogen through the fermentation of cellulose and reported that temperature variation was a key factor that enhanced the bio-production of hydrogen. They further posited that thermophilic and hyper-thermophilic conditions yield more hydrogen than mesophilic conditions. Furthermore, Zou et al. [32] investigated the effects of a Ce/Fe bimetallic catalyst on hydrogen generation from cellulose steam at 500–900 °C. They showed that the catalytic impact of CeO2/FeO3 composite on hydrogen generation was superior to CeO2 or Fe2O3 alone. Studies on hydrogen generation from cellulose gasified in hot-compressed water via reduced nickel catalyst at a temperature range of 200 °C to 350 °C have also been reported [33]. The authors revealed that the catalytic activity was dependent on the size of the catalyst and not the surface area.
Composites that include metals as one of the components employed for hydrogen generation commonly include magnesium due to its low weight and ability to form mechanically resistant alloys [25]. In addition, magnesium is a metal that reacts with water to liberate hydrogen and has been reported to exhibit low toxicity. Besides magnesium, titanium has also been used due to its unique properties such as its non-toxicity, excellent corrosion resistance, good mechanical strength, and biocompatibility. These properties make titanium a promising candidate for application in various chemical, petrochemical, and biomaterial processes. Niobium is another metal that has been considered in the development of materials for hydrogen generation due to its excellent corrosion resistance and its ability to form an oxide layer when it comes in contact with air [25]. The oxide layer acts as a protective barrier that blocks additional air and moisture intrusion to prevent corrosion.
Despite the relative availability, biodegradability, biocompatibility, good stability, and mechanical stability of cellulosic biomass, the use of cellulose-based composites that contain niobium, titanium, and magnesium for hydrogen generation via the hydrolysis method is yet to be reported. Therefore, to bridge this gap, this work aims to evaluate the hydrogen generation potentials of cellulosic biomass (cellulose, microcrystalline cellulose, and ball-milled cellulose) and its composites with niobium, titanium, and magnesium. The effects of the presence of electrolytes on hydrogen generation shall also be evaluated through experimental and theoretical studies. Firstly, the composite samples were characterized with different techniques, such as powder X-ray diffraction (PXRD), scanning electron microscopy analysis (SEM), and X-ray photoelectron spectroscopy (XPS). Secondly, an experimental study was carried out to evaluate the hydrogen generation potentials of microcrystalline cellulose, ball-milled cellulose, and a composite of this cellulosic biomass with niobium, titanium, and magnesium along with the effects of the presence of electrolytes on the hydrolysis reaction. Finally, a molecular dynamic simulation to elucidate the nature of the interaction between cellulose and the metals was performed using density functional theory (DFT). The contributions of this research are highlighted as follows: (i) the use of cellulosic biomass composites with various metal ions as a “green” strategy for hydrogen generation, (ii) the role of ball milling in sustainable hydrogen generation, and (iii) the effects of electrolytes on hydrogen generation via hydrolysis of cellulosic biomass and their composite materials with metals. This study will contribute to the increased interest in the development of green and sustainable strategies based on cellulosic biomass and its composites to produce hydrogen for onboard applications.

2. Experimental Procedure

2.1. Materials

Microcrystalline cellulose (medium fiber from cotton linters), niobium, urea, sodium hydroxide, and titanium powders were ordered from Sigma-Aldrich Canada Ltd. (Oakville, ON, Canada), while magnesium powder (99.8% purity, particle size 20 + 230 mesh) was obtained from Alfa Aesar, (Ward Hill, MA, USA). All materials were used without further purification.

2.2. Preparation of Cellulose Doped with Niobium, Titanium, and Magnesium

The composite materials were synthesized as follows: briefly, 2 g of the cellulosic biomass was dissolved in a NaOH/urea solution at room temperature. Following this, 0.5 g of the respective metal powders (magnesium, titanium, and niobium) was added to the cellulose solution as oxophilic agents, and the mixture was stirred for approximately 18 h. The resulting cellulosic biomass–metal composite was centrifuged, and the supernatant was decanted, followed by drying the prepared composites in an oven at 70 °C for 48 h. Then, the materials were stored in vials for further use.

2.3. Powder X-ray Diffraction

The PXRD profiles of the materials were obtained using an AXS D8 ADVANCE Bruker X-ray diffractometer (Germany) with a monochromatic source (Cu Kα1) radiation target (λ = 1.540 Å and n = 1), as previously reported [34]. The PXRD spectra were obtained in a continuous scanning mode from 5–80°, with an applied voltage and current of 40 kV and 40 mA, respectively.

2.4. Scanning Electron Microscopy

A Hitachi SU6600 field emission scanning electron microscope was utilized to evaluate the composite materials’ SEM images. The working conditions for image acquisition employed an accelerating voltage (3 kV), the working distance (8.1–8.4 mm), and a magnification of 2.5–3 k.

2.5. X-ray Photoelectron Spectroscopy

To determine the elemental composition of the materials, Kratos (Manchester, UK) AXIS Supra system was used. It is composed of a 500 mm Rowland circle monochromatic Al Kα (1486.6 eV) source, a combined hemispherical analyzer, and a spherical mirror analyzer. The survey scan spectra were collected in the 0–1200 binding energy range in 1 eV. A spot size with a hybrid slot (300 × 700 μm2) and a pass energy of 160 eV was used. Scanning of several regions was analyzed using 0.1 eV steps with a pass energy of 20 eV for the high-resolution analysis. An emission current of 15 mA and a 15 keV accelerating voltage were used for the study.

2.6. Hydrogen Production Procedure

Hydrogen generation was carried out as previously reported by Figen et al. [35], where one gram of each of the powdered composite materials was added to 200 mL of distilled water at 89 ± 3 °C with a constant stirring rate of 120 rpm in an Erlenmeyer flask. CoCl2 was used to absorb the moisture in the system. An ADM2000 flowmeter with an accuracy of 0.1 mL/min was used to measure the amount of H2 gas generated. For data acquisition, the flowmeter was connected to a computer running ADM Trend software, version 12.1 [36]. To account for any contributions from water vapor or expansion of the air in the flask during heating, the experiment was conducted in similar conditions without the presence of the composites. The amount of H2 gas generated was determined as the difference between the cumulative volume of H2 gas obtained from the two experiments, with and without the addition of the composite materials.

3. Results and discussion

3.1. X-ray Diffraction Analysis

PXRD studies of the samples were carried out and the results are presented in Figure 1. The PXRD patterns of MCC reveal peaks at 2 θ = 22.02° and 2 θ = 18.12 ° corresponding to the (002) and (101) plains with an interlayer distance of 0.39 nm and 0.48 nm, respectively. Similarly, Figure 1 displays the minimum peak for the ball-milled MCC at 2 θ   = 18.12°, where the two peaks found in the non-ball-milled MCC seems to have merged into a broader peak at 2 θ = 22.02°, in agreement with the claim of a decline in the crystalline order of MCC due to ball milling. The results imply that MCC exhibits long-range order relative to the ball-milled MCC, in agreement with a related study [37]. Furthermore, the PXRD pattern of MCC treated with aqueous NaOH is also shown in Figure 1. The PXRD patterns further reveal that the treatment of MCC with aqueous NaOH results in sharp XRD bands at 2 θ = 33.5°, 2 θ = 47 ° , and 2 θ = 67 ° , which support the binding of NaOH to MCC. The above results suggest that the inhibitory effects of NaOH on hydrogen generation potential of MCC results from an increase in crystalline ordering of MCC through the binding of NaOH crystals to MCC.
Moreover, Figure 2a–d display the PXRD patterns of ball-milled MCC, the ball-milled MCC-Mg composite, the ball-milled MCC-Ti composite, and the ball-milled MCC-Nb composite treated with deionized water. The PXRD pattern of the ball-milled MCC-Mg composite (Figure 2b) displays bands at 2 θ = 38.10 corresponding to MgO (2 θ = 38.10°) and 2 θ = 49.88° corresponding to Mg(OH)2 in addition to signatures ascribed to MCC. Similarly, the ball-milled MCC-Nb and ball-milled MCC-Ti composites treated with deionized water (Figure 2c,d) display peaks corresponding to NbO2 (65.07°) and TiO2 (44.52°) along with the peaks ascribed to MCC. The PXRD patterns and crystallinity index obtained from the peak areas reveal greater crystalline ordering for the MCC-Mg composite treated with water and the ball-milled MCC-Ti composite [38] treated with deionized water relative to the ball-milled MCC-Nb composite [30] treated with deionized water. This trend concurs with the highest volume of hydrogen generated by the ball-milled MCC-Nb composite that was treated with deionized water, as reported earlier [25].

3.2. Scanning Electron Microscopy Analysis

The scanning electron microscopy (SEM) results for MCC, Mg, Nb, MCC-Mg, MCC-Ti, and MCC-Nb are shown in Figure 3.
The micrographs of MCC reveal different particle sizes with various geometry, as compared with the MCC that underwent 24 h of ball milling, where a smoother surface with reduced particle size is evident. These results are in agreement with other SEM studies reported in the literature [39,40]. On the other hand, the micrographs of Mg, Ti, and Nb exhibit various irregular shapes with angular and sharp edges, as shown in our previous work [25]. Moreover, the SEM micrographs reveal that the metal is randomly dispersed on the MCC surface which attests to the presence of different morphologies for MCC-Mg, MCC-Ti, and MCC-Nb, where the differences noted in the morphology of the images relate to the MCC and MCC–metal-oxide domains, respectively. The SEM results support the formation of a composite between these metals and MCC, where a similar analysis was performed on graphene oxide, as reported previously [25].

3.3. X-ray Photoelectron Spectroscopy

The high-resolution XPS spectra of the composites that underwent 24 h of ball milling (MCC-Mg, MCC-Ti, and MCC-Nb) are displayed in Figure 4 and Table 1. The results provide evidence of the existence of metals in variable oxidation states, according to our expectations for composites that contain metal species and MCC. For example, magnesium exists in the zerovalent (Mg) and divalent (Mg2+) oxidation states, where Mg2+ species were the most prevalent, in agreement with a previous study [25]. The XPS spectra for Mg supports the formation of Mg(OH)2, which concurs with a mechanism for the hydrolysis of Mg. Similarly, titanium is present in multivalent oxidation states (e.g., Ti2+, Ti3+, and Ti4+), where the Ti2+ species are the most prevalent [41]. In the case of Nb, the results reveal the existence of variable oxidation states (e.g., Nb4+ and Nb5+) [41,42], with the pentavalent state as the most abundant. These XPS results agree with the PXRD patterns, where the metal oxide species were also observed.

3.4. Computational Method

In addition to the experimental evaluation of hydrogen evolution of the MCC–metal (Mg, Ti and Nb) composites treated with deionized water, an NVT molecular dynamic simulation with the NH chain thermostat was implemented in the Dmol3 module and the material studio was performed at 500 K. The generalized gradient approximation (GGA) using Perdew, Burke, and Ernzerhof (PBE) [43] was used to approximate the effects of exchange correlation on electron–electron interactions. The semi-core pseudo potentials represent the core electrons as a single effective potential [44,45]. The final structures and the fluctuation of the temperature over the dynamical step are illustrated in Figure 5.
The simulated structures reveal the release of H2 molecule in all three cases after 800 dynamic steps. Moreover, the formation of Ti(OH)2, and Nb(OH)5 from the reaction of MCC-Ti and MCC-Nb with deionized water concurs with the presence of the metals in variable oxidation states, as supported by the PXRD and XPS results.

3.5. Hydrogen Generation from Cellulose

Due to its environmental friendliness and abundance in nature, we performed the hydrolysis of cellulose for hydrogen generation, where a diagram for this system was previously reported [25]. The cumulative volume of hydrogen yielded by MCC and its composites over time under various conditions is displayed in Figure 6.
Figure 6 shows that the hydrogen yield of MCC treated with deionized water at 89 ± 3 °C for over 4500 s is 40 mL. On the other hand, the treatment of MCC with NaOH (aq) led to a drop in hydrogen generation (14 mL) over 3000 s. The low hydrogen yield observed due to the treatment of MCC with aqueous NaOH indicates the inhibition of hydrogen production by NaOH (aq). However, the treatment of MCC with aqueous urea resulted in the generation of 35 mL of hydrogen over 3000 s. These results reveal that the effects of the urea inhibition on hydrogen generation is not as prominent when compared to NaOH (aq). The above may be ascribed to the claim that the solvation of cellulose by urea prevents agglomeration through the weakening of hydrophobic effects, due to the chaotropic properties of urea in water [40]. By contrast, solvation in aqueous NaOH involves a strong ion–dipolar interaction with cellulose due to ionization of the hydroxyl groups of the polysaccharide, which does not support hydrogen generation [31,32].
Furthermore, MCC dissolved in aqueous NaOH/urea solution was treated with titanium, magnesium, and niobium and their hydrogen generation potentials were evaluated. The results of hydrogen generation by the various composite mixtures (MCC-Mg, MCC-Ti and MCC-Nb) are displayed in Figure 6. The results reveal that the hydrogen yields are approximately 34 mL, 37 mL, and 45 mL over 3000 s for MCC-Mg, MCC-Ti, and MCC-Nb, respectively, where the volume of hydrogen gas generated by the MCC-Nb composite mixture is slightly higher than the one generated by MCC treated with H2O (40 mL). This variation in the hydrogen generation of composite materials agrees with the differences in their relative crystallinity as inferred from the PXRD results (Figure 2). Furthermore, the highest oxophilic activity of Nb, affirmed by the abundance of the metal in its +5-oxidation state from the XPS results, supports the reported trend.
Furthermore, to evaluate the effects of ball milling on the hydrogen generation potentials of MCC, the biomass was ball milled for 24 h. The hydrogen yield for samples of MCC that underwent 24 h of ball milling showed an 87.5% increase in hydrogen generation (75 mL) over 4000 s relative to 40 mL for MCC treated with deionized water. The 87.5% increase in hydrogen generation due to ball milling of MCC may be ascribed to the decrease in particle size and the crystallinity of MCC along with a change from the typical cellulose IV to cellulose II [40,46]. Ball milling results in the destruction of the hydrogen bonding that holds the hydroxyl groups in an arrangement that supports the high crystallinity of cellulose. The destruction of the hydrogen bonding makes the hydroxyl groups in cellulose available for hydrolysis. However, the addition of NaOH to the MCC-deionized water mixture led to a drop in hydrogen generation from 75 mL to approximately 40 mL over 4000 s. These results reveal that NaOH inhibits the hydrolysis of MCC, thus leading to a drop in the hydrogen yield of ball-milled MCC, as shown in Figure 7. The effects of composite formation between ball-milled MCC and metals such as magnesium (Mg), titanium (Ti), and niobium (Nb) were also studied in the presence of deionized water, which are also presented in Figure 7. The hydrogen yield of the MCC-Mg, MCC-Ti, and MCC-Nb composites are reported as 60 mL, 85 mL, and 110 mL, respectively. These findings revealed that composite formation between ball-milled MCC and the metals (Mg, Ti, and Nb) enhanced the hydrogen production potential of ball-milled MCC through a catalytic effect, where the MCC-Nb composite displayed the greatest and fastest hydrogen generation potential, in agreement with results reported earlier. Previous reports [47,48,49] have proposed that introducing oxophilic metals onto the surfaces of noble metal nanocrystals may result in the acceleration of electrochemical and hydrolytic reactions via the regulation of the adsorption energy of hydroxyl species (OHads). In particular, Kim et al. [50] reported that oxophilic Cr dopants enhanced both the H and OH binding energies of the Ni nanohelixes, resulting in the acceleration of the dissociative adsorption of water and overall kinetics of the hydrogen evolution reaction. The observed differences in the hydrogen generation potentials of the various composites relate to the oxophilic activities of the metals in the composite, where niobium with the highest oxophilic activity and the abundance of its oxides in the +5-oxidation state (cf. Table 1) supports the highest hydrogen generation volume after over 3000 s of the hydrolysis reaction.

4. Conclusions

This study reports a theoretical and experimental study of H2 production via the hydrolysis of microcrystalline cellulose (MCC) and MCC-metal composites treated with deionized water, urea and NaOH. The results demonstrate that the presence of NaOH and urea inhibit the hydrogen generation efficiency of MCC. Furthermore, ball milling of MCC for 24 h led to enhanced hydrogen generation potentials of MCC. This effect was related to a decrease in the crystalline order of MCC in agreement with PXRD spectra results, where a broader peak at 2θ = 22.02° was observed for the ball-milled MCC. Additionally, the theoretical prediction supports the formation of oxides of the metals, according to PXRD and XPS results. As well, the hydrogen yield performance adopts the following order: MCC-Nb > MCC-Ti > MCC-Mg. Finally, the results demonstrated that ball milling represents an effective way to improve the kinetics of hydrogen generation.

Author Contributions

Conceptualization, O.F. and J.A.S.; software, J.A.S. and O.F.; validation, J.A.S., O.F., and L.D.W.; formal analysis, O.F., and I.A.U.; investigation, O.F., J.A.S. and L.D.W.; data curation, O.F. and J.A.S.; writing—original draft preparation, O.F. and I.A.U.; writing—review and editing, J.A.S. and L.D.W.; supervision, J.A.S. and L.D.W.; funding acquisition, J.A.S. All authors have read and agreed to the published version of the manuscript.

Funding

The authors express their appreciation to the National Engineering Research Council of Canada and the Canada Research Chairs program for providing support for this research.

Data Availability Statement

Data are contained within the article.

Acknowledgments

We acknowledge that we had access to high-performance supercomputers, Compute Canada, and Plato at the University of Saskatchewan. This work was supported by the National Engineering Research Council of Canada and the Canada Research Chairs program. The authors acknowledge that this work was carried out in the Homeland of the Métis and Treaty 6 Territory, and we pay our respect to the First Nations and Métis ancestors of this place and reaffirm our relationship with one another. We are very grateful to Andrew Olkowski.

Conflicts of Interest

The authors declare that there is no conflict of interest.

References

  1. Perea-Moreno, M.-A.; Samerón-Manzano, E.; Perea-Moreno, A.-J. Biomass as Renewable Energy: Worldwide Research Trends. Sustainability 2019, 11, 863. [Google Scholar] [CrossRef]
  2. Ji, S.; Jang, J.; Cho, E.; Kim, S.H.; Kang, E.S.; Kim, J.; Kim, H.K.; Kong, H.; Kim, S.K.; Kim, J.Y.; et al. High Dielectric Performances of Flexible and Transparent Cellulose Hybrid Films Controlled by Multidimensional Metal Nanostructures. Adv. Mater. 2017, 29, 1700538. [Google Scholar] [CrossRef]
  3. Wang, S.; Lu, A.; Zhang, L. Recent Advances in Regenerated Cellulose Materials. Prog. Polym. Sci. 2016, 53, 169–206. [Google Scholar] [CrossRef]
  4. Fridrihsone, V.; Zoldners, J.; Skute, M.; Grinfelds, U.; Filipova, I.; Sivacovs, I.; Spade, M.; Laka, M. Dissolution of Various Cellulosic Materials and Effect of Regenerated Cellulose on Mechanical Properties of Paper. Key Eng. Mater. 2019, 800, 138–144. [Google Scholar] [CrossRef]
  5. Fu, F.; Gu, J.; Cao, J.; Shen, R.; Liu, H.; Zhang, Y.; Liu, X.; Zhou, J. Reduction of Silver Ions Using an Alkaline Cellulose Dope: Straightforward Access to Ag/ZnO Decorated Cellulose Nanocomposite Film with Enhanced Antibacterial Activities. ACS Sustain. Chem. Eng. 2018, 6, 738–748. [Google Scholar] [CrossRef]
  6. Chang, S.T.; Chen, L.C.; Lin, S.B.; Chen, H.H. Nano-Biomaterials Application: Morphology and Physical Properties of Bacterial Cellulose/Gelatin Composites via Crosslinking. Food Hydrocoll. 2012, 27, 137–144. [Google Scholar] [CrossRef]
  7. Mohammed, N.; Grishkewich, N.; Tam, K.C. Cellulose Nanomaterials: Promising Sustainable Nanomaterials for Application in Water/Wastewater Treatment Processes. Environ. Sci. Nano 2018, 5, 623–658. [Google Scholar] [CrossRef]
  8. Rahman, N.S.A.; Yhaya, M.F.; Azahari, B.; Ismail, W.R. Utilisation of Natural Cellulose Fibres in Wastewater Treatment. Cellulose 2018, 25, 4887–4903. [Google Scholar] [CrossRef]
  9. Peng, B.; Yao, Z.; Wang, X.; Crombeen, M.; Sweeney, D.G.; Tam, K.C. Cellulose-Based Materials in Wastewater Treatment of Petroleum Industry. Green Energy Environ. 2020, 5, 37–49. [Google Scholar] [CrossRef]
  10. Jang, M.; Kim, S.K.; Lee, J.; Ji, S.; Song, W.; Myung, S.; Lim, J.; Lee, S.S.; Jung, H.K.; Lee, J.; et al. Amplifying Gas Sensor Performance by Embedding a Cellulose-Based Buffer Layer in Organic Transistors. J. Mater. Chem. C Mater. 2019, 7, 14504–14510. [Google Scholar] [CrossRef]
  11. Peres, B.U.; Vidotti, H.A.; de Carvalho, L.D.; Manso, A.P.; Ko, F.; Carvalho, R.M. Nanocrystalline Cellulose as a Reinforcing Agent for Electrospun Polyacrylonitrile (PAN) Nanofibers. J. Oral. Biosci. 2019, 61, 37–42. [Google Scholar] [CrossRef] [PubMed]
  12. Xu, X.; Liu, F.; Jiang, L.; Zhu, J.Y.; Haagenson, D.; Wiesenborn, D.P. Cellulose Nanocrystals vs. Cellulose Nanofibrils: A Comparative Study on Their Microstructures and Effects as Polymer Reinforcing Agents. ACS Appl. Mater. Interfaces 2013, 5, 2999–3009. [Google Scholar] [CrossRef]
  13. Xu, J.T.; Chen, X.Q.; Shen, W.H.; Li, Z. Spherical vs Rod-like Cellulose Nanocrystals from Enzymolysis: A Comparative Study as Reinforcing Agents on Polyvinyl Alcohol. Carbohydr. Polym. 2021, 256, 117493. [Google Scholar] [CrossRef] [PubMed]
  14. Wang, X.; Yao, C.; Wang, F.; Li, Z. Cellulose-Based Nanomaterials for Energy Applications. Small 2017, 13, 1702240. [Google Scholar] [CrossRef] [PubMed]
  15. Dutta, S.; Kim, J.; Ide, Y.; Ho Kim, J.; Hossain, M.S.A.; Bando, Y.; Yamauchi, Y.; Wu, K.C.W. 3D Network of Cellulose-Based Energy Storage Devices and Related Emerging Applications. Mater. Horiz. 2017, 4, 522–545. [Google Scholar] [CrossRef]
  16. Chen, Y.; Zhang, L.; Mei, C.; Li, Y.; Duan, G.; Agarwal, S.; Greiner, A.; Ma, C.; Jiang, S. Wood-Inspired Anisotropic Cellulose Nanofibril Composite Sponges for Multifunctional Applications. ACS Appl. Mater. Interfaces 2020, 12, 35513–35522. [Google Scholar] [CrossRef] [PubMed]
  17. Moreira, F.S.; Machado, R.G.; Romão, B.B.; Batista, F.R.X.; Ferreira, J.S.; Cardoso, V.L. Improvement of Hydrogen Production by Biological Route Using Repeated Batch Cycles. Process Biochem. 2017, 58, 60–68. [Google Scholar] [CrossRef]
  18. Faye, O.; Szpunar, J.A. An Efficient Way to Suppress the Competition between Adsorption of H2 and Desorption of n H2-Nb Complex from Graphene Sheet: A Promising Approach to H2 Storage. J. Phys. Chem. C 2018, 122, 28506–28517. [Google Scholar] [CrossRef]
  19. Kayfeci, M.; Keçebaş, A.; Bayat, M. Hydrogen Production. In Solar Hydrogen Production: Processes, Systems and Technologies; Elsevier Science: London, UK; San Diego, CA, USA, 2019; ISBN 9780128148549. [Google Scholar]
  20. Coutanceau, C.; Baranton, S.; Audichon, T. Hydrogen Electrochemical Production. In Hydrogen Electrochemical Production; Pollet, B.G., Ed.; Academic Press: Cambridge, UK, 2018; pp. 1–6. ISBN 978-0-12-811250-2. [Google Scholar]
  21. Mohanty, P.; Pant, K.K.; Mittal, R. Hydrogen Generation from Biomass Materials: Challenges and Opportunities. WIREs Energy Environ. 2015, 4, 139–155. [Google Scholar] [CrossRef]
  22. Huang, M.; Ouyang, L.; Wang, H.; Liu, J.; Zhu, M. Hydrogen Generation by Hydrolysis of MgH2 and Enhanced Kinetics Performance of Ammonium Chloride Introducing. Int. J. Hydrogen Energy 2015, 40, 6145–6150. [Google Scholar] [CrossRef]
  23. Tan, Z.H.; Ouyang, L.Z.; Huang, J.M.; Liu, J.W.; Wang, H.; Shao, H.Y.; Zhu, M. Hydrogen Generation via Hydrolysis of Mg2Si. J. Alloys Compd. 2019, 770, 108–115. [Google Scholar] [CrossRef]
  24. Ma, M.; Yang, L.; Ouyang, L.; Shao, H.; Zhu, M. Promoting Hydrogen Generation from the Hydrolysis of Mg-Graphite Composites by Plasma-Assisted Milling. Energy 2019, 167, 1205–1211. [Google Scholar] [CrossRef]
  25. Faye, O.; Udoetok, I.A.; Szpunar, J.A.; Wilson, L.D. Experimental and Theoretical Studies of Hydrogen Generation by Binary Metal (Oxide)-Graphene Oxide Composite Materials. Int. J. Hydrogen Energy 2021, 46, 19802–19813. [Google Scholar] [CrossRef]
  26. Wang, C.; Wang, Q.; Fu, F.; Astruc, D. Hydrogen Generation upon Nanocatalyzed Hydrolysis of Hydrogen-Rich Boron Derivatives: Recent Developments. Acc. Chem. Res. 2020, 53, 2483–2493. [Google Scholar] [CrossRef] [PubMed]
  27. Chen, K.; Ouyang, L.; Wang, H.; Liu, J.; Shao, H.; Zhu, M. A High-Performance Hydrogen Generation System: Hydrolysis of LiBH4-Based Materials Catalyzed by Transition Metal Chlorides. Renew. Energy 2020, 156, 655–664. [Google Scholar] [CrossRef]
  28. Zhang, X.; Zhang, Q.; Xu, B.; Liu, X.; Zhang, K.; Fan, G.; Jiang, W. Efficient Hydrogen Generation from the NaBH4 Hydrolysis by Cobalt-Based Catalysts: Positive Roles of Sulfur-Containing Salts. ACS Appl. Mater. Interfaces 2020, 12, 9376–9386. [Google Scholar] [CrossRef] [PubMed]
  29. Zagrodnik, R.; Seifert, K. Direct Fermentative Hydrogen Production from Cellulose and Starch with Mesophilic Bacterial Consortia. Pol. J. Microbiol. 2020, 69, 109–120. [Google Scholar] [CrossRef] [PubMed]
  30. Caravaca, A.; Jones, W.; Hardacre, C.; Bowker, M. H2 Production by the Photocatalytic Reforming of Cellulose and Raw Biomass Using Ni, Pd, Pt and Au on Titania. Proc. R. Soc. A Math. Phys. Eng. Sci. 2016, 472, 20160054. [Google Scholar]
  31. Gadow, S.I.; Li, Y.Y.; Liu, Y. Effect of Temperature on Continuous Hydrogen Production of Cellulose. Int. J. Hydrog. Energy 2012, 37, 15465–15472. [Google Scholar] [CrossRef]
  32. Zou, J.; Oladipo, J.; Fu, S.; Al-Rahbi, A.; Yang, H.; Wu, C.; Cai, N.; Williams, P.; Chen, H. Hydrogen Production from Cellulose Catalytic Gasification on CeO2/Fe2O3 Catalyst. Energy Convers. Manag. 2018, 171, 241–248. [Google Scholar] [CrossRef]
  33. Minowa, T.; Ogi, T. Hydrogen Production from Cellulose Using a Reduced Nickel Catalyst. Catal. Today 1998, 45, 411–416. [Google Scholar] [CrossRef]
  34. Udoetok, I.A.; Wilson, L.D.; Headley, J.V. Self-Assembled and Cross-Linked Animal and Plant-Based Polysaccharides: Chitosan-Cellulose Composites and Their Anion Uptake Properties. ACS Appl. Mater. Interfaces 2016, 8, 33197–33209. [Google Scholar] [CrossRef] [PubMed]
  35. Kantürk Figen, A.; Coşkuner, B.; Pişkin, S. Hydrogen Generation from Waste Mg Based Material in Various Saline Solutions (NiCl2, CoCl2, CuCl2, FeCl3, MnCl2). Int. J. Hydrogen Energy 2015, 40, 7483–7489. [Google Scholar] [CrossRef]
  36. Shetty, T.; Szpunar, J.A.; Faye, O.; Eduok, U. A Comparative Study of Hydrogen Generation by Reaction of Ball Milled Mixture of Magnesium Powder with Two Water-Soluble Salts (NaCl and KCl) in Hot Water. Int. J. Hydrogen Energy 2020, 45, 25890–25899. [Google Scholar] [CrossRef]
  37. Amini, M.; Rahimipour, M.R.; Tayebifard, S.A.; Palizdar, Y. Effect of Milling Time on XRD Phases and Microstructure of a Novel Al 67 Cu 20 Fe 10 B 3 Quasicrystalline Alloy Effect of Milling Time on XRD Phases and Microstructure of a Novel Al67Cu20Fe10B3 Quasicrystalline Alloy. Mater. Res. Express 2020, 7, 065011. [Google Scholar] [CrossRef]
  38. Shen, X.; Liu, F.; Guan, J.; Dong, F.; Zhang, Y.; Guo, Z.; Yuan, Y.; Wang, B.; Luo, L.; Su, Y.; et al. Effect of hydrogen on thermal deformation behavior and microstructure evolution of MoNbHfZrTi refractory high-entropy alloy. Intermetallics 2024, 166, 108193. [Google Scholar] [CrossRef]
  39. Das, K.; Ray, D.; Bandyopadhyay, N.R.; Sengupta, S. Study of the Properties of Microcrystalline Cellulose Particles from Different Renewable Resources by XRD, FTIR, Nanoindentation, TGA and SEM. J. Polym. Environ. 2010, 18, 355–363. [Google Scholar] [CrossRef]
  40. Zheng, Y.; Fu, Z.; Li, D.; Wu, M. Effects of Ball Milling Processes on the Microstructure and Rheological Properties of Microcrystalline Cellulose as a Sustainable Polymer Additive. Materials 2018, 11, 1057. [Google Scholar] [CrossRef] [PubMed]
  41. Guan, W.; Sun, G.; Yin, L.; Zhang, Z.; Tian, S. Ti4O7/g-C3N4 Visible Light Photocatalytic Performance on Hypophosphite Oxidation: Effect of Annealing Temperature. Front. Chem. 2018, 6, 37. [Google Scholar] [CrossRef]
  42. Hryniewicz, T.; Rokosz, K.; Sandim, H.R.Z. SEM/EDX and XPS Studies of Niobium after Electropolishing. Appl. Surf. Sci. 2012, 263, 357–361. [Google Scholar] [CrossRef]
  43. Chukwuike, V.; Kesavan, R.; Barik, R.C. Surface and Electrochemical Corrosion Analysis of Niobium Oxide Film Formed in Various Wet Media. Appl. Surf. Sci. Adv. 2021, 4, 100079. [Google Scholar] [CrossRef]
  44. Perdew, J.P.; Burke, K.; Ernzerhof, M. Generalized Gradient Approximation Made Simple. Phys. Rev. Lett. 1996, 77, 3865. [Google Scholar] [CrossRef] [PubMed]
  45. Perdew, J.P.; Chevary, J.A.; Vosko, S.H.; Koblar, A.J.; Pederson, M.R.; Singh, D.J.; Fiolhais, C. Atoms, Molecules, Solids and Surfaces: Applications of the GGA for Exchange Correlation. Phys. Rev. B 1992, 46, 6671–6687. [Google Scholar] [CrossRef] [PubMed]
  46. Su, S.T.; Xiong, J.; Ye, J. Effect of Ball Milling on Structure of Microcrystalline Cellulose. Appl. Mech. Mater. 2013, 394, 201–204. [Google Scholar] [CrossRef]
  47. Luc, W.; Jiang, Z.; Chen, J.G.; Jiao, F. Role of Surface Oxophilicity in Copper-Catalyzed Water Dissociation. ACS Catal. 2018, 8, 9327–9333. [Google Scholar] [CrossRef]
  48. Jin, H.; Liu, X.; Jiao, Y.; Vasileff, A.; Zheng, Y.; Qiao, S.-Z. Constructing Tunable Dual Active Sites on Two-Dimensional C3N4@ MoN Hybrid for Electrocatalytic Hydrogen Evolution. Nano Energy 2018, 53, 690–697. [Google Scholar] [CrossRef]
  49. Wu, F.; Zhang, L.; Lai, J.; Luque, R.; Niu, W.; Xu, G. Modulating the Oxophilic Properties of Inorganic Nanomaterials for Electrocatalysis of Small Carbonaceous Molecules. Nano Today 2019, 29, 100802. [Google Scholar] [CrossRef]
  50. Kim, J.; Jung, H.; Jung, S.-M.; Hwang, J.; Kim, D.Y.; Lee, N.; Kim, K.-S.; Kwon, H.; Kim, Y.-T.; Han, J.W.; et al. Tailoring Binding Abilities by Incorporating Oxophilic Transition Metals on 3D Nanostructured Ni Arrays for Accelerated Alkaline Hydrogen Evolution Reaction. J. Am. Chem. Soc. 2021, 143, 1399–1408. [Google Scholar] [CrossRef]
Figure 1. XRD patterns of pure microcrystalline cellulose (MCC), ball-milled microcrystalline cellulose, and ball-milled microcrystalline cellulose plus sodium hydroxide (NaOH).
Figure 1. XRD patterns of pure microcrystalline cellulose (MCC), ball-milled microcrystalline cellulose, and ball-milled microcrystalline cellulose plus sodium hydroxide (NaOH).
Jcs 08 00262 g001
Figure 2. XRD spectral patterns of ball-milled microcrystalline cellulose (a), and ball-milled microcrystalline cellulose doped with magnesium (b), titanium (c), and niobium (d). (The indices in colored fonts relate to the metals while those in black font relate to cellulose).
Figure 2. XRD spectral patterns of ball-milled microcrystalline cellulose (a), and ball-milled microcrystalline cellulose doped with magnesium (b), titanium (c), and niobium (d). (The indices in colored fonts relate to the metals while those in black font relate to cellulose).
Jcs 08 00262 g002
Figure 3. Representation of the scanning electron microscopy of MCC, Mg, Nb, Ti, 24 h of ball milling of MCC, and the corresponding functionalized MCC-Mg, MCC-Nb, and MCC-Ti.
Figure 3. Representation of the scanning electron microscopy of MCC, Mg, Nb, Ti, 24 h of ball milling of MCC, and the corresponding functionalized MCC-Mg, MCC-Nb, and MCC-Ti.
Jcs 08 00262 g003
Figure 4. Representation of the XPS spectra of the metals and the various MCC composites (MCC-Mg, MCC-Nb, and MCC-Ti), where the black line is the envelop and green line is the background.
Figure 4. Representation of the XPS spectra of the metals and the various MCC composites (MCC-Mg, MCC-Nb, and MCC-Ti), where the black line is the envelop and green line is the background.
Jcs 08 00262 g004
Figure 5. The final structures of the dynamics of (a) MCC-Mg, (b) MCC-Ti, and (c) MCC-Nb treated with water. The colour coding of the spheres (atoms) is given as follows: Green (magnesium), dark green niobium), white (hydrogen), red (oxygen), grey (titanium), and dark grey (carbon).
Figure 5. The final structures of the dynamics of (a) MCC-Mg, (b) MCC-Ti, and (c) MCC-Nb treated with water. The colour coding of the spheres (atoms) is given as follows: Green (magnesium), dark green niobium), white (hydrogen), red (oxygen), grey (titanium), and dark grey (carbon).
Jcs 08 00262 g005
Figure 6. Hydrogen yield in (mL) of microcrystalline cellulose and its composites over time (seconds). The black (MCC + Urea + NaOH + Ti + H2O), orange (MCC + Urea + NaOH + Nb + H2O), green (MCC+ NaOH + H2O), red (MCC + Urea + H2O), dark red (MCC + Urea + NaOH + Mg + H2O), and blue (MCC + H2O) lines denote the hydrogen production for the various systems studied.
Figure 6. Hydrogen yield in (mL) of microcrystalline cellulose and its composites over time (seconds). The black (MCC + Urea + NaOH + Ti + H2O), orange (MCC + Urea + NaOH + Nb + H2O), green (MCC+ NaOH + H2O), red (MCC + Urea + H2O), dark red (MCC + Urea + NaOH + Mg + H2O), and blue (MCC + H2O) lines denote the hydrogen production for the various systems studied.
Jcs 08 00262 g006
Figure 7. Hydrogen generation of ball-milled MCC treated with pure water and salt water, and microcrystallaine cellulose functionalized with magnesium, titanium and niobium. The coloured lines are defined as follows: orange (ball-milled MCC + H2O), green (ball-milled MCC + NaOH + H2O), dark red (ball-milled MCC + Mg + H2O), black (ball-milled MCC + Ti + H2O) and blue (ball-milled MCC + Nb + H2O) for various conditions of hydrogen production.
Figure 7. Hydrogen generation of ball-milled MCC treated with pure water and salt water, and microcrystallaine cellulose functionalized with magnesium, titanium and niobium. The coloured lines are defined as follows: orange (ball-milled MCC + H2O), green (ball-milled MCC + NaOH + H2O), dark red (ball-milled MCC + Mg + H2O), black (ball-milled MCC + Ti + H2O) and blue (ball-milled MCC + Nb + H2O) for various conditions of hydrogen production.
Jcs 08 00262 g007
Table 1. Binding energies and atom content (wt.%) of composite materials.
Table 1. Binding energies and atom content (wt.%) of composite materials.
Elements/
Materials
Magnesium (Mg 2p)Titanium (Ti 2p)Niobium (Nb 3d)
PositionArea%PositionArea%PositionArea%
MCC-Mg49.2711769.7NANA
50.1309830.3
MCC-TiNA46032.954.6NA
45724.340.3
4563.065.08
MCC-NbNANA20636.926.0
20979.155.6
21126.318.5
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

Faye, O.; Udoetok, I.A.; Szpunar, J.A.; Wilson, L.D. Hydrolyzed Forms of Cellulose and Its Metal Composites for Hydrogen Generation: An Experimental and Theoretical Investigation. J. Compos. Sci. 2024, 8, 262. https://doi.org/10.3390/jcs8070262

AMA Style

Faye O, Udoetok IA, Szpunar JA, Wilson LD. Hydrolyzed Forms of Cellulose and Its Metal Composites for Hydrogen Generation: An Experimental and Theoretical Investigation. Journal of Composites Science. 2024; 8(7):262. https://doi.org/10.3390/jcs8070262

Chicago/Turabian Style

Faye, Omar, Inimfon A. Udoetok, Jerzy A. Szpunar, and Lee D. Wilson. 2024. "Hydrolyzed Forms of Cellulose and Its Metal Composites for Hydrogen Generation: An Experimental and Theoretical Investigation" Journal of Composites Science 8, no. 7: 262. https://doi.org/10.3390/jcs8070262

Article Metrics

Back to TopTop