Next Article in Journal
Energy-Efficient Driving for Adaptive Traffic Signal Control Environment via Explainable Reinforcement Learning
Next Article in Special Issue
Thermodynamics of Chemical Hydrogen Storage: Are Sterically Hindered and Overcrowded Molecules More Effective?
Previous Article in Journal
Segmentation of Infant Brain Using Nonnegative Matrix Factorization
Previous Article in Special Issue
Surfactant Induced Synthesis of LiAlH4 and NaAlH4 Nanoparticles for Hydrogen Storage
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

A Novel Method for Generating H2 by Activation of the μAl-Water System Using Aluminum Nanoparticles

Department of Chemistry, Saint Louis University, 3501 Laclede Ave., St. Louis, MO 63103, USA
*
Authors to whom correspondence should be addressed.
Appl. Sci. 2022, 12(11), 5378; https://doi.org/10.3390/app12115378
Submission received: 6 May 2022 / Revised: 20 May 2022 / Accepted: 23 May 2022 / Published: 26 May 2022

Abstract

:

Featured Application

Aluminum nanoparticles are integrated with micron-scale particles to enhance hydrolytic sensitivity to water and accelerate hydrogen gas production.

Abstract

A method is described for activation of the reaction of room temperature water with micron-scale aluminum particles (μAl) by the addition of poly(epoxyhexane)-capped aluminum nanoparticles (Al NPs). By themselves, Al NPs react vigorously and completely with water at ambient temperatures to produce H2. While pure μAl particles are unreactive toward water, mixtures of the μAl particles comprising 10 to 90% (by mass) of Al NPs, demonstrated appreciable hydrolytic activation. This activation is attributed to the reaction of the Al NPs present with water to produce a basic solution. Speciation modelling, pH studies, and powder X-ray diffraction analysis of the hydrolysis product confirm that the pH change is the key driver for the activation of μAl rather than residual heat from the exothermicity of Al NP hydrolysis. A mechanism is proposed by which the nonreactive aluminum oxide layer of the μAl is eroded under basic conditions. Mixtures 10% by mass of Al NPs can be used to produce the optimal quantity of H2.

Graphical Abstract

1. Introduction

Aluminum, Al, is the most abundant metal in the earth’s crust. When Al reacts with water, it produces H2 gas and aluminum hydroxide, Al(OH)3 [1]. The H2 produced can be harnessed as a clean, ecologically friendly energy source, making it a desirable fuel for many applications [2,3,4,5,6]. The development of methods for H2 production is motivated by a vision to move away from fossil fuels due to their finite availability and the environmental damage caused by their combustion [7,8,9,10]. There are many ways to generate H2 including water electrolysis [11], chemical reactions [12], steam reforming of hydrocarbons [13], and from biological sources [14]. These methods are not amenable to portable-use systems, as they are either too expensive or initiated under extreme conditions. Transporting pure H2 has challenges with its low volumetric energy density and high flammability. Thus, it is necessary to develop safe and economically feasible H2 delivery systems that can be utilized on-demand. Aluminum is deemed as a possible source of energy storage for direct H2 generation as long as its surface is passivated efficiently, effectively and safely [15].
Pure aluminum spontaneously reacts with water due to its high reduction potential (E° = −1.66 V) but can easily form a thin oxide layer when in contact with air, which leads to remarkably low reactivity [16]. This thin oxide layer develops into a 75 to 100 nm thick layer in a few months. It is also insoluble in, and impermeable to, water and prevents contact between the metal and water, hindering access to the energy-rich aluminum metal core [17]. Thus, μAl powders remain almost unhydrolyzed due to the thick oxide layer formation [18,19]. Because of this, it is important to find a method to activate the surface of μAl to allow access to the energetically rich and reactive metallic core. This problem can be solved by mixing organic capped Al nanoparticles (Al NPs) with μAl. Strategies for the activation of aluminum and aluminum alloys also have importance in additive manufacturing processes [20].
When the problems related to surface oxidation and agglomeration are overcome, Al NPs demonstrate fast reaction rates, exceptional combustion efficiency, and higher volumetric energy densities in comparison to conventional materials [21,22]. Due to their high surface-to-volume ratio, nanoparticles can have exceptionally strong interactions with other components in a nanocomposite. Thus, nanoaluminum surfaces may be stabilized by passivation and surface coated with organic polymers or polymerizable substrates to control their size and safety in energetics applications [23,24,25,26,27,28,29,30,31,32,33,34,35,36,37,38,39,40].
Our lab has previously reported that epoxyalkane-capped Al NPs demonstrated moderate air stability, as well as high active aluminum content [41,42,43]. Of these, epoxyhexane-capped Al NPs have the least protective cap owing to the shorter alkyl side-chain and react rapidly and completely with water to generate H2 at room temperature, and produce basic conditions that can be exploited for the benefit of μAl hydrolytic activation. Thus the Al NPs could work as an indirect promoter by generating Al(OH)3 to facilitate H2 production from μAl at room temperature. Wang et al. have previously reported that Al(OH)3 could catalytically promote H2 generation from the μAl-water system and was more effective compared with other promoters [44,45,46,47].
Several composite samples of μAl and Al NPs were prepared with a variety of ratios by mass and analyzed for their hydrogen generation efficiency upon hydrolysis. pH measurements and pH-dependent speciation modelling were used to determine whether to attribute the μAl activation to the exothermicity of the Al NPs’ hydrolysis or the resulting pH change. The optimal composition was determined based on the notion that the most useful, cost-effective energetic material will generate the maximal quantity of hydrogen from the minimal quantity of Al NPs in the mixture.

2. Materials and Methods

2.1. Reagents and Materials

Titanium (IV) isopropoxide (99.999% trace metals basis), N,N-dimethylethylamine alane solution (0.5 M in toluene), 1,2-epoxyhexane (97%), disodium ethylenediaminetetraacetic acid dihydrate (Na2EDTA·2H2O, 99%), zinc sulfate heptahydrate (ZnSO4·7H2O, 99.999% trace metals basis), and xylenol orange were purchased from Sigma-Aldrich (Milwaukee, WI, USA). Sodium chloride was purchased from Fisher Scientific (Waltham, MA, USA) and used to make saturated aqueous solutions. Micron-scale aluminum (μAl, 30 μm average diameter particles) was purchased from Sigma-Aldrich and used without further purification. Toluene was distilled under inert argon over Na or 4A molecular sieves to remove trace water and oxygen. The nitric acid and sodium hydroxide pellets used in the gas titration apparatus were purchased from Sigma-Aldrich.

2.2. Synthesis of Al NPs

The following is a slight modification of the previously described synthesis of epoxyhexane-capped aluminum nanoparticles [43]. All reactions were carried out on a Schlenk line under high purity nitrogen or argon. For each reaction, 6.5 mL (3.25 mmol) of 0.5 M N,N-dimethylethylamine alane solution in toluene was syringed into a Schlenk flask containing a further 10 mL of toluene. The flask was fitted with a reflux condenser and the solution was heated to 85 °C. Another solution was prepared by mixing 99 μL of titanium (IV) isopropoxide with 10 mL of toluene and then 0.8 mL of this solution was syringed into the stirred, heated alane solution, leading to a change from colorless to black. The 1,2-epoxyhexane capping agent (40 μL, 10:1 Al:epoxide molar ratio) was immediately added directly to the reaction mixture. After vigorous stirring of the reaction mixture at 85 °C for 30 min, the solvent was removed under reduced pressure and the solid residue dried in vacuo.

2.3. Characterization of Al NPs

Powder X-ray diffraction (PXRD) measurements were performed using a Rigaku Miniflex 600 diffractometer. The Rigaku had a Cu K(α) source and a scintillation counter detector. The ICDD Crystallographic Database was used to compare PXRD patterns. The instrument software used Scherrer analysis to determine nanocrystallite dimensions.
Images of nanoparticle morphology and dimensions were obtained from a JEOL 1200 EX TEM instrument, which operated at 80 kV and had 200,000× magnification.
The DSC/TGA instrument used to obtain thermal profiles was a TA Instruments SDT Q600 simultaneous DSC/TGA instrument. The instrument had a constant flow rate of 50 mL/min of N2, temperature range (25–800) °C, as well as a 10 °C/min temperature ramp. The samples were held in alumina cups.
The active and total aluminum contents of the Al NPs were determined by previously used methods of quantitative hydrolytic H2 evaluation and complexometric Zn2+−EDTA back-titration, respectively [48,49,50].

2.4. Hydrolysis Activation Experiments

Five samples composed of various mass ratios of μAl with Al NPs were prepared. Each sample weighed a total of 50 mg precisely, with Al NPs mass % values of 10, 25, 50, 75, and 90%, the remaining mass being the μAl particles. A sixth sample of pure μAl without added Al NPs was also studied. The solid fuels were placed in an inverted-Y-shaped reaction vessel connected by flexible rubber tubing to a burette filled with saturated NaCl (aq) (in order to minimize H2 dissolution), which was in turn connected via flexible plastic tubing to a glass funnel. The set-up is depicted in Figure 1. In the Y-tube side-arm was placed exactly 25.0 mL of deionized water. A timer was started upon tipping the water onto the Al solid sample and the H2 (g) volume collected measured in the burette. By appropriate leveling of the water in the funnel, the H2 pressure was recorded as the measured ambient atmospheric pressure at that time. Gas volume measurement was carried out until there appeared to be little visually discernible change in volume from one minute to the next (15 min for all samples). Samples were then left to sit in water for 3 days to allow any residual Al to hydrolyze. White solid residues from these samples were isolated by evaporating the water in an oven at 120 °C and then analyzed by PXRD.

3. Results and Discussion

3.1. Synthesis and Characterization of Al NPs

The Al NPs used in this study were synthesized by a slight modification of a previously described method [43]. The products were characterized by PXRD, which confirmed the presence of the crystalline Al cores (fcc Al, ICDD # 00-004-0787) with an average nanocrystallite diameter of 12 nm. Some polycrystallinity was observed in TEM analysis, with an average size of 30 nm. DSC/TGA also confirmed the core-shell nature of the nanoparticles, where an organic cap is removed at temperatures up to 460 °C, while aluminum oxidation onset takes place at 565 °C. All-in-all, these diagnostic characterizations (Supplementary Materials) closely matched those previously reported. Hydrolytic H2 formation and complexometric Zn2+−EDTA back-titration also confirmed an active aluminum content (Al0 as a fraction of all aluminum present) of 83 ± 5% in the samples used for the μAl hydrolysis activation study.

3.2. Hydrolysis Activation Study

In the absence of any catalytic promoter, μAl particles hydrolyze so slowly as to be unobservable. Al NPs, on the other hand, react rapidly and completely with water to form Al(OH)3 (s) (Equation (1)). We hypothesize that the Al(OH)3 produced in this more rapid reaction then promotes the observable acceleration of the μAl reaction with water (Equation (2)). Alternatively, the exothermicity of Al NP hydrolysis in Equation (1) could be the driver for the pursuant reaction in Equation (2); the following will discern which of these is the more likely.
2 Al   NPs   ( s ) + 6 H 2 O   ( l ) 2 Al ( OH ) 3   ( s ) + 3 H 2   ( g )
2 μ Al   ( s ) + 6 H 2 O   ( l ) Catalytic   Al ( OH ) 3 2 Al ( OH ) 3   ( s ) + 3 H 2   ( g )
Experiments were thus designed where μAl and Al NPs were combined with various mass ratios, from 10 to 90% Al NPs, and then their hydrolysis reactions were studied using a gas burette assembly shown in Figure 1. Any measured volume of hydrogen produced, Vmix, at atmospheric pressure, P, was deemed to have originated only from active Al0, either from the Al NPs (primary source) or from the μAl (secondary source) and was therefore the sum of V1 and V2, respectively (Equation (3)). To calculate V2, the reaction to produce H2 (with volume V1) by the hydrolysis of the same quantity of Al NPs without the μAl present was also carried out. Volume V2 was then used to determine the number of moles, n2, of Al0 reacting from the μAl only, according to the ideal gas law (Equation (4)). Finally, the ‘activated μAl’ was quantified as the stochiometric equivalent mole amount of μAl0 that reacted to produce the n2 moles of H2 relative to the total quantity of μAl used in the experiment, nμAl (Equation (5)) and expressed as a percentage. Allowing for oxide passivation, the latter quantity was adjusted by coefficient f, the mass fraction of active Al0 (Al g/mol) in the mass of μAl used, mμAl, which was calculated to be 0.99 based on a spherical core volume fraction for 30 μm particles with a 100 nm thick oxide shell. The ‘activated μAl’ was plotted versus the mass % of Al NPs used in the mixed μAl/Al NPs samples and is shown in Figure 2.
V m i x = V 1 + V 2
n 2 = P V 2 R T = P ( V m i x V 1 ) R T
Activated   μ Al = 3 2 n 2 n μ Al × 100 % = 3 2 P ( V m i x V 1 ) × A l R T × f m μ Al × 100 %
The inset bar chart of Figure 2 shows that in the absence of Al NPs, there was no activation toward hydrolysis with less than 1% of the μAl reacting. Integration with just 10% Al NPs by mass incurred substantial activation with ca. 52% of the μAl consumed by the time H2 generation appeared to have visually ceased. As the proportion of Al NPs was raised, a steady increase in μAl activation was observed, reaching ca. 59% corresponding to 90% Al NPs by mass. As the main plot in Figure 2 shows, there is a simple linear relationship between these two quantities between 10 and 90% Al NPs by mass. The activation range thus spans only 7% and given the need to maximize H2 production with the minimum incorporation of Al NPs, our results suggest that a composite 9:1 μAl/Al NPs material (by mass) is the best possible formulation. Indeed, when the total moles of active μAl0 per unit mass of Al NPs is considered (Figure 3), it can be seen that this was highest at 0.204 mol/g for the 9:1 μAl/Al NPs sample and decreased rapidly before becoming steady at ca. 0.035 mol/g in the limit for samples with the highest proportions of Al NPs.
The hydrolysis of the μAl/Al NPs composite material is a complex process that is kinetically very difficult to model. However, tracking the hydrogen gas production from 50 mg samples as a function of reaction time (Figure 4a) clearly revealed a dual phase bimodal process. Irrespective of mass ratio, a burst of hydrogen production occurred within ca. 1 min. The rate at which hydrogen was produced during this initial phase is linearly dependent upon the mass % of Al NPs in the sample, confirming first order kinetics (Figure 4b). This is entirely consistent with our previously determined kinetic profile for the hydrolysis of poly(N-isopropylacrylamide)-capped Al nanoparticles [40]. This was followed by a steadier gas formation phase, which settled at around 0.4 mL/min during the final 3 min of measurement, irrespective of the mass % of Al NPs in the sample. This is to be expected, given the now-dominant surface-area limited hydrolysis of the much larger μAl particles. It is also important that the reaction environment created by the Al NPs hydrolysis (vide infra) affords a stable and consistent supply of hydrogen from the μAl, which can be helpful for maintaining steady fuel generation for a device such as a hydrogen fuel cell.
For the 50% Al NPs/50% μAl sample after 3 days sitting in water, analysis by PXRD of the recovered, postreaction residue (Figure 5) determined a crystalline composition that was 94% Al(OH)3 (s) and 6% elemental Al(s), presumably unreacted μAl. No AlO(OH) was detected in this measurement. The Al(OH)3 formation is typically expected in most aluminum hydrolysis reactions below 280 °C [4,51]. Additionally, Al(OH)3 is an affirmed catalytic promoter of this reaction [44,45,46,47]. In the case of our μAl/Al NPs composite samples, there was rapid consumption of the Al NPs to produce Al(OH)3 in the initial activation phase of the reaction. This is possible because the oligo(epoxyhexane) cap on the nanoparticles is easily compromised by pure water. The Al(OH)3 produced by Al NPs hydrolysis then catalytically drives the μAl hydrolysis phase by raising the reaction pH and compromising the micron-scale particle oxide shell. Higher proportions of Al NPs in the composite mixture will clearly generate more Al(OH)3 during the activation, further elevating the pH, as was observed, and further activating the μAl toward the hydrolysis and ultimately increasing the hydrogen production.
The hydrolysis of aluminum nanoparticles to bulk Al(OH)3 is a highly exothermic process (ΔH° = −1094 kJ/mol) [52]. It is conceivable, therefore, that energy released by this swift reaction could drive pursuant μAl particle hydrolysis. However, monitoring the temperature of the reaction solution using a thermocouple probe recorded a temperature increase of just 5–7 °C. Thus, it would seem the heat release rate is not reasonably sufficient to initiate degradation of the Al2O3 shell protecting the μAl particles and accelerate the μAl reaction. Most micron-scale and bulk forms of Al do not react appreciably with neutral water without some form of ball milling, chemical activation and/or elevated temperatures (≥35 °C) [1,10,53,54,55,56]. On the other hand, the Al2O3 protective shell is vulnerable to alkaline pH [1]. A sufficient increase in reaction pH might compromise the oxide layer to the extent that μAl hydrolysis is expedited. Since the Al(OH)3 produced by Al NP hydrolysis is capable of promoting such pH-driven μAl hydrolysis to produce more Al(OH)3, it could be viewed as a catalytic enhancement. Increasing the proportion of Al NPs in the composite material naturally increases the rate of initial Al(OH)3 formation in the reaction mixture (supported, of course, by the observed increase of the initial rate of formation of the other product, H2). We also measured a corresponding increase in the pH of the reaction solution at the end of the 15 min reaction monitoring period (Figure 6). Indeed, with the value of the slope close to 3, the dependence of the (basic) pH on the log(Al NP amount reacting) closely matches the 3:1 hydroxide-to-Al stoichiometry of the Al NP hydrolysis equation, which supports the rapid and complete consumption of the nanoparticles in the initial phase. The measured increase in resulting pH with the increasing Al NP proportion is offset by the reduced fraction of μAl in the sample. The result is the same steady rate of μAl hydrolysis observed in the final reaction phase of Figure 4a for all Al NP mass % samples.
The solution pH measured at the culmination of the hydrogen formation experiments ranged from 8 to 11 for the lowest and highest Al NPs mass % samples, respectively. This basic pH regime is consistent with Al3+ speciation analysis for the aqueous solutions shown in Figure 7a (and determined by equilibria equations shown in Appendix A), bearing in mind that Al(OH)3 is a sparingly soluble salt. Thus, while most of the Al(OH)3 precipitates out of solution, observed as a white solid, the thermodynamically equilibrated composition in solution ranges from ca. 50% Al(OH)3 (aq)/50% (Al(OH)4 (aq) at pH 8 to almost entirely Al(OH)4 (aq) at pH 11. These equilibria systems also confirm that the Al2O3 shell protecting the μAl particles is most likely to be compromised at a pH greater than 8, yielding primarily Al(OH)4 ions. As a result, the total saturated concentration of aqueous Al species increased by three orders of magnitude as expected in the pH range 8–11 (Figure 7b, Appendix B). This, therefore, is the most likely mode of activation of the μAl toward hydrolysis for this composite material.

4. Conclusions

The sensitivity and reactivity of aluminum nanoparticles (Al NPs) were exploited to activate the hydrolytic reactivity of micron-scale aluminum particles (μAl) through the catalytic promotion by Al(OH)3 and the associated increase in solution pH rather than by thermal activation. This method provides for the fabrication of a simple binary admixture of μAl with Al NPs that could be packaged as a portable solid and used to produce hydrogen gas on demand, such as that which might be needed for a fuel cell. Given that Al NPs are typically much more expensive and hazardous than μAl, this work demonstrates that it is beneficial to formulate a small (≤10%) proportion of the composite material as Al NPs relative the μAl to achieve the optimal μAl activation. Proportions of Al NPs >10% would likely not provide a substantially cost-effective performance benefit improvement in terms of overall H2 production.

Supplementary Materials

The following supporting information can be downloaded at: https://www.mdpi.com/article/10.3390/app12115378/s1, Figure S1: PXRD pattern of the Al NPs; Figure S2: TEM images of the Al NPs; Figure S3: DSC/TGA traces of the Al NPs.

Author Contributions

Conceptualization, W.Z., M.S.K., E.J., S.W.B., P.A.J.; methodology, M.S.K., W.Z., E.J., S.W.B., P.A.J.; formal analysis, M.S.K., E.J., W.Z.; investigation, M.S.K., E.J., W.Z.; resources, S.W.B., P.A.J.; data curation, M.S.K., E.J., P.A.J., S.W.B.; writing—original draft preparation, M.S.K., W.Z., E.J.; writing—review and editing, P.A.J., S.W.B.; supervision, S.W.B., P.A.J.; project administration, S.W.B., P.A.J. All authors have read and agreed to the published version of the manuscript.

Funding

This research received no external funding.

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

Not applicable.

Acknowledgments

The authors wish to thank Michael Briscoe for technical support and custom glassware fabrication.

Conflicts of Interest

The authors declare no conflict of interest.

Appendix A

The following speciation reactions and equations were used to construct Figure 7a. Aluminum oxide in water is shown in reaction (A1):
Al2O3.3H2O (s) ⇌ 2Al(OH)3 (s)
Al3+ reacts with water to produce the following ions in aqueous solution:
Al3+ (aq) + H2O (l) ⇌ Al(OH)2+ (aq) + H+ (aq); KF1 = β1 = 10−4.97
Al3+ (aq) + 2H2O (l) ⇌ Al(OH)2+ (aq) + 2H+ (aq); β2 = 10−9.3
Al3+ (aq) + 3H2O (l) ⇌ Al(OH)3 (aq) + 3H+ (aq); β3 = 10−15
Al3+ (aq) + 4H2O (l) ⇌ Al(OH)4 (aq) + 4H+ (aq); β4 = 10−23
The corresponding equilibrium constants for these reactions include:
K F 1 = β 1 = 10 4.97 = [ Al ( OH ) 2 + ] [ H + ] [ Al 3 + ]
β 2 = 10 9.3 = [ Al ( OH ) 2 + ] [ H + ] 2 [ Al 3 + ]
β 3 = 10 15 = [ Al ( OH ) 3 ] [ H + ] 3 [ Al 3 + ]
β 4 = 10 23 = [ Al ( OH ) 4 ] [ H + ] 4 [ Al 3 + ]
The fractional composition equations for all of the different forms of aluminum in solution (i.e., Al3+, AlOH2+, Al(OH)2+, Al(OH)3 (aq), and Al(OH)4) are given by their corresponding α equations, For example, the speciation equations for the Al3+ ion in aqueous solution are given in Equations (A4):
α 0 = [ Al 3 + ] [ Al 3 + ] + [ Al ( OH ) 2 + ] + [ Al ( OH ) 2 + ] + [ Al ( OH ) 3 ] + [ Al ( OH ) 4 ]
α 0 = 1 1 + [ Al ( OH ) 2 + ] [ Al 3 + ] + [ Al ( OH ) 2 + ] [ Al 3 + ] + [ Al ( OH ) 3 ] [ Al 3 + ] + [ Al ( OH ) 4 ] [ Al 3 + ]
α 0 = 1 1 + K 1 [ H + ] + β 2 [ H + ] 2 + β 3 [ H + ] 3 + β 4 [ H + ] 4 = 1 D
The corresponding speciation equations for the remaining forms of dissolved aluminum in solution are given by:
α 1 = K F 1 [ H + ] × D   for   Al ( OH ) 2 + ( a q )
α 2 = β 2 [ H + ] 2 × D   for   Al ( OH ) 2 + ( a q )  
α 3 = β 3 [ H + ] 3 × D   for   Al ( OH ) 3 ( a q )
α 4 = β 4 [ H + ] 4 × D   for   Al ( OH ) 4 ( a q )

Appendix B

The following speciation reactions and equations were used to construct Figure 7b. The total concentration of all Al species may be calculated by Equation (A6) based on the distribution of Al in its various forms in aqueous solution:
[Al]total = [Al3+] + [Al(OH)2+] + [Al(OH)2+] + [Al(OH)3] + [Al(OH)4]
The individual terms for (A6) are obtained from the corresponding equilibrium terms for each individual form:
Al(OH)3 (s) + 3H+ (aq) ⇌ Al3+ (aq) + 3H2O (l); K = 108.5
[ Al 3 + ] = K × [ H + ] 3
Al(OH)3 (s) + 3H+ (aq) ⇌ Al3+ (aq) + 3H2O (l); K
Al3+ (aq) + H2O (l) ⇌ Al(OH)2+ (aq) + H+ (aq); KF1
[ Al ( OH ) 2 + ] = K × K F 1 × [ H + ] 2
Al(OH)3 (s) + 3H+ (aq) ⇌ Al3+ (aq) + 3H2O (l); K
Al3+ (aq) + 2H2O (l) ⇌ Al(OH)2+ (aq) + 2H+ (aq); β2
[ Al ( OH ) 2 + ] = K × β 2 × [ H + ]
Al(OH)3 (s) + 3H+ (aq) ⇌ Al3+ (aq) + 3H2O (l); K
Al3+ (aq) +3H2O (l) ⇌ Al(OH)3 (aq) + 3H+ (aq); β3
[ Al ( OH ) 3 ] = K × β 3
Al(OH)3 (s) + 3H+ (aq) ⇌ Al3+ (aq) + 3H2O (l); K
Al3+ (aq) + 4H2O (l) ⇌ Al(OH)4 (aq) + 4H+ (aq); β4
[ Al ( OH ) 4 ] = K × β 4 [ H + ]

References

  1. Wang, H.Z.; Leung, D.Y.C.; Leung, M.K.H.; Ni, M. A review on hydrogen production using aluminum and aluminum alloys. Renew. Sustain. Energy Rev. 2009, 13, 845–853. [Google Scholar] [CrossRef]
  2. Xu, S.; Zhao, X.; Liu, J. Liquid metal activated aluminum-water reaction for direct hydrogen generation at room temperature. Renew. Sustain. Energy Rev. 2018, 92, 17–37. [Google Scholar] [CrossRef]
  3. Alinejad, B.; Mahmoodi, K. A novel method for generating hydrogen by hydrolysis of highly activated aluminum nanoparticles in pure water. Int. J. Hydrog. Energy 2009, 34, 7934–7938. [Google Scholar] [CrossRef]
  4. Ma, G.L.; Zhuang, D.W.; Dai, H.; Wang, P. Controlled hydrogen generation by reaction of aluminum with water. Prog. Chem. 2012, 24, 650–658. [Google Scholar]
  5. Uehara, K.; Takeshita, H.; Kotaka, H. Hydrogen gas generation in the wet cutting of aluminum and its alloys. J. Mater. Process. Technol. 2002, 127, 174–177. [Google Scholar] [CrossRef]
  6. Elitzur, S.; Rosenband, V.; Gany, A. Study of hydrogen production and storage based on aluminum–water reaction. Int. J. Hydrog. Energy 2014, 39, 6328–6334. [Google Scholar] [CrossRef]
  7. Sartbaeva, A.; Kuznetsov, V.L.; Wells, S.A.; Edwards, P.P. Hydrogen nexus in a sustainable energy future. Energy Environ. Sci. 2008, 1, 79–85. [Google Scholar] [CrossRef]
  8. Wang, W.; Zhao, X.M.; Chen, D.M.; Yang, K. Insight into the reactivity of Al–Ga–In–Sn alloy with water. Int. J. Hydrog. Energy 2012, 37, 2187–2194. [Google Scholar] [CrossRef]
  9. Parmuzina, A.V.; Kravchenko, O.V. Activation of aluminium metal to evolve hydrogen from water. Int. J. Hydrog. Energy 2008, 33, 3073–3076. [Google Scholar] [CrossRef]
  10. Rosenband, V.; Gany, A. Application of activated aluminum powder for generation of hydrogen from water. Int. J. Hydrog. Energy 2010, 35, 10898–10904. [Google Scholar] [CrossRef]
  11. Rashid, M.; Al Mesfer, M.; Naseem, H.; Danish, M. Hydrogen production by water electrolysis: A review of alkaline water electrolysis, PEM water electrolysis and high temperature water electrolysis. Int. J. Eng. Adv. Technol. 2015, 4, 80–93. [Google Scholar]
  12. Kojima, Y.; Suzuki, K.-i.; Fukumoto, K.; Sasaki, M.; Yamamoto, T.; Kawai, Y.; Hayashi, H. Hydrogen generation using sodium borohydride solution and metal catalyst coated on metal oxide. Int. J. Hydrog. Energy 2002, 27, 1029–1034. [Google Scholar] [CrossRef]
  13. Brown, L.F. A comparative study of fuels for on-board hydrogen production for fuel-cell-powered automobiles. Int. J. Hydrog. Energy 2001, 26, 381–397. [Google Scholar] [CrossRef]
  14. Sivaramakrishnan, R.; Shanmugam, S.; Sekar, M.; Mathimani, T.; Incharoensakdi, A.; Kim, S.-H.; Parthiban, A.; Edwin Geo, V.; Brindhadevi, K.; Pugazhendhi, A. Insights on biological hydrogen production routes and potential microorganisms for high hydrogen yield. Fuel 2021, 291, 120136. [Google Scholar] [CrossRef]
  15. Newell, A.; Thampi, K.R. Novel amorphous aluminum hydroxide catalysts for aluminum–water reactions to produce H2 on demand. Int. J. Hydrog. Energy 2017, 42, 23446–23454. [Google Scholar] [CrossRef]
  16. Torvi, A.; Naik, S.; Kariduraganavar, M. Development of supercapacitor systems based on binary and ternary nanocomposites using chitosan, graphene and polyaniline. Chem. Data Collect. 2018, 17–18, 459–471. [Google Scholar] [CrossRef]
  17. Thakur, V.K.; Thakur, M.K. Recent advances in graft copolymerization and applications of chitosan: A review. ACS Sustain. Chem. Eng. 2014, 2, 2637–2652. [Google Scholar] [CrossRef]
  18. Madhumitha, G.; Fowsiya, J.; Mohana Roopan, S.M.; Thakur, V.K. Recent advances in starch–clay nanocomposites. Int. J. Polym. Anal. Charact. 2018, 23, 331–345. [Google Scholar] [CrossRef]
  19. Casar, G.; Li, X.; Malič, B.; Zhang, Q.M.; Bobnar, V. Impact of structural changes on dielectric and thermal properties of vinylidene fluoride–trifluoroethylene-based terpolymer/copolymer blends. Phys. B Phys. Cond. Matt. 2015, 461, 5–9. [Google Scholar] [CrossRef] [Green Version]
  20. Veron, F.; Lanoue, F.; Baco-Carles, V.; Kiryukhina, K.; Vendier, O.; Tailhades, P. Selective laser powder bed fusion for manufacturing of 3D metal-ceramic multi-materials assemblies. Addit. Manuf. 2022, 50, 102550. [Google Scholar] [CrossRef]
  21. Dreizin, E.L. Metal-based reactive nanomaterials. Prog. Energy Combust. Sci. 2009, 35, 141–167. [Google Scholar] [CrossRef]
  22. Yetter, R.A.; Risha, G.A.; Son, S.F. Metal particle combustion and nanotechnology. Proc. Combust. Inst. 2009, 32, 1819–1838. [Google Scholar] [CrossRef]
  23. Jelliss, P.A.; Thomas, B.J.; Patel, A. Polymerization passivation strategies for the stabilization of energetic aluminum nanomaterials. Int. J. Chem. 2014, 3, 122–131. [Google Scholar]
  24. Ates, B.; Koytepe, S.; Ulu, A.; Gurses, C.; Thakur, V.K. Chemistry, structures, and advanced applications of nanocomposites from biorenewable resources. Chem. Rev. 2020, 120, 9304–9362. [Google Scholar] [CrossRef]
  25. Zeng, W.; Nyapete, C.; Benziger, A.; Jelliss, P.; Buckner, S. Encapsulation of reactive nanoparticles of aluminum, magnesium, zinc, titanium, or boron within polymers for energetic applications. Curr. Appl. Polym. Sci. 2019, 3, 3–13. [Google Scholar] [CrossRef]
  26. Patel, A.; Becic, J.; Buckner, S.W.; Jelliss, P.A. Reactive aluminum metal nanoparticles within a photodegradable poly(methyl methacrylate) matrix. Chem. Phys. Lett. 2014, 591, 268–272. [Google Scholar] [CrossRef]
  27. Zeng, W.H.; Buckner, S.W.; Jelliss, P.A. Poly(methyl methacrylate) as an environmentally responsive capping material for aluminum nanoparticles. ACS Omega 2017, 2, 2034–2040. [Google Scholar] [CrossRef] [Green Version]
  28. Abdelkader, E.M.; Jelliss, P.A.; Buckner, S.W. Main group nanoparticle synthesis using electrical explosion of wires. Nano-Struct. Nano-Objects 2016, 7, 23–31. [Google Scholar] [CrossRef]
  29. Atmane, Y.A.; Sicard, L.; Lamouri, A.; Pinson, J.; Sicard, M.; Masson, C.; Nowak, S.; Decorse, P.; Piquemal, J.Y.; Galtayries, A.; et al. Functionalization of aluminum nanoparticles using a combination of aryl diazonium salt chemistry and iniferter method. J. Phys. Chem. C 2013, 117, 26000–26006. [Google Scholar] [CrossRef]
  30. Esmaeili, B.; Chaouki, J.; Dubois, C. Nanoparticle encapsulation by a polymer via in situ polymerization in supercritical conditions. Polym. Eng. Sci. 2012, 52, 637–642. [Google Scholar] [CrossRef]
  31. Ghanta, S.R.; Muralidharan, K. Solution phase chemical synthesis of nano aluminium particles stabilized in poly(vinylpyrrolidone) and poly(methylmethacrylate) matrices. Nanoscale 2010, 2, 976–980. [Google Scholar] [CrossRef] [PubMed]
  32. Ghanta, S.R.; Muralidharan, K. Chemical synthesis of aluminum nanoparticles. J. Nanopart. Res. 2013, 15, 1715. [Google Scholar] [CrossRef]
  33. Gottapu, S.; Padhi, S.K.; Krishna, M.G.; Muralidharan, K. Poly(vinylpyrrolidone) stabilized aluminum nanoparticles obtained by the reaction of SiCl4 with LiAlH4. New J. Chem. 2015, 39, 5203–5207. [Google Scholar] [CrossRef]
  34. Huang, X.Y.; Jiang, P.K.; Kim, C.N.; Ke, Q.Q.; Wang, G.L. Preparation, microstructure and properties of polyethylene aluminum nanocomposite dielectrics. Compos. Sci. Technol. 2008, 68, 2134–2140. [Google Scholar] [CrossRef]
  35. Huang, X.Y.; Kim, C.N.; Ma, Z.S.; Jiang, P.K.; Yin, Y.; Li, Z. Correlation between rheological, electrical, and microstructure characteristics in polyethylene/aluminum nanocomposites. J. Polym. Sci. Polym. Phys. 2008, 46, 2143–2154. [Google Scholar] [CrossRef]
  36. Huang, X.Y.; Ma, Z.S.; Wang, Y.Q.; Jiang, P.K.; Yin, Y.; Li, Z. Polyethylene/aluminum nanocomposites: Improvement of dielectric strength by nanoparticle surface modification. J. Appl. Polym. Sci. 2009, 113, 3577–3584. [Google Scholar] [CrossRef]
  37. Madhankumar, A.; Nagarajan, S.; Rajendran, N.; Nishimura, T. EIS evaluation of protective performance and surface characterization of epoxy coating with aluminum nanoparticles after wet and dry corrosion test. J. Solid State Electrochem. 2012, 16, 2085–2093. [Google Scholar] [CrossRef]
  38. Nishimura, T.; Raman, V. Corrosion prevention of aluminum nanoparticles by a polyurethane coating. Materials 2014, 7, 4710–4722. [Google Scholar] [CrossRef] [Green Version]
  39. Nishimura, T.; Raman, V. Epoxy polymer coating to prevent the corrosion of aluminum nanoparticles. Polym. Advan. Technol. 2016, 27, 712–717. [Google Scholar] [CrossRef]
  40. Zeng, W.; Jelliss, P.A.; Buckner, S.W. Synthesis and hydrogen production kinetics of temperature-responsive aluminum-poly(N-isopropylacrylamide) core-shell nanoparticles. Mater. Chem. Phys. 2018, 220, 233–239. [Google Scholar] [CrossRef]
  41. Chung, S.W.; Guliants, E.A.; Bunker, C.E.; Hammerstroem, D.W.; Deng, Y.; Burgers, M.A.; Jelliss, P.A.; Buckner, S.W. Capping and passivation of aluminum nanoparticles using alkyl-substituted epoxides. Langmuir 2009, 25, 8883–8887. [Google Scholar] [CrossRef] [PubMed]
  42. Hammerstroem, D.W.; Burgers, M.A.; Chung, S.W.; Guliants, E.A.; Bunker, C.E.; Wentz, K.M.; Hayes, S.E.; Buckner, S.W.; Jelliss, P.A. Aluminum nanoparticles capped by polymerization of alkyl-substituted epoxides: Ratio-dependent stability and particle size. Inorg. Chem. 2011, 50, 5054–5059. [Google Scholar] [CrossRef] [PubMed]
  43. Jelliss, P.A.; Buckner, S.W.; Chung, S.W.; Patel, A.; Guliants, E.A.; Bunker, C.E. The use of 1,2-epoxyhexane as a passivating agent for core-shell aluminum nanoparticles with very high active aluminum content. Solid State Sci. 2013, 23, 8–12. [Google Scholar] [CrossRef]
  44. Wang, H.-W.; Chin, M.-S. Rapid hydrogen generation from aluminum-water system using synthesized aluminum hydroxide catalyst. Int. J. Chem. Eng. Appl. 2015, 6, 146–151. [Google Scholar] [CrossRef] [Green Version]
  45. Chen, Y.-K.; Teng, H.-T.; Lee, T.-Y.; Wang, H.-W. Rapid hydrogen generation from aluminum–water system by adjusting water ratio to various aluminum/aluminum hydroxide. Int. J. Energy Environ. Eng. 2014, 5, 87. [Google Scholar] [CrossRef] [Green Version]
  46. Prabu, S.; Hsu, S.C.; Lin, J.S.; Wang, H.W. Rapid hydrogen generation from the reaction of aluminum powders and water using synthesized aluminum hydroxide catalysts. Top. Catal. 2018, 61, 1633–1640. [Google Scholar] [CrossRef]
  47. Teng, H.-T.; Lee, T.-Y.; Chen, Y.-K.; Wang, H.-W.; Cao, G. Effect of Al(OH)3 on the hydrogen generation of aluminum–water system. J. Power Sources 2012, 219, 16–21. [Google Scholar] [CrossRef]
  48. Thomas, B.J.; Bunker, C.E.; Guliants, E.A.; Hayes, S.E.; Kheyfets, A.; Wentz, K.M.; Buckner, S.W.; Jelliss, P.A. Synthesis of aluminum nanoparticles capped with copolymerizable epoxides. J. Nanopart. Res. 2013, 15, 1729. [Google Scholar] [CrossRef]
  49. Glotov, O.G.; Zyryanov, V.Y. The effect of pressure on characteristics of condensed combustion products of aluminized solid propellant. Arch. Combust. 1991, 11, 251–262. [Google Scholar]
  50. Yang, S.-P.; Tsai, R.-Y. Complexometric titration of aluminum and magnesium ions in commercial antacids. An experiment for general and analytical chemistry laboratories. J. Chem. Ed. 2006, 83, 906–909. [Google Scholar] [CrossRef]
  51. Digne, M.; Sautet, P.; Raybaud, P.; Toulhoat, H.; Artacho, E. Structure and stability of aluminum hydroxides:  A theoretical study. J. Phys. Chem. B 2002, 106, 5155–5162. [Google Scholar] [CrossRef]
  52. Chung, S.W.; Guliants, E.A.; Bunker, C.E.; Jelliss, P.A.; Buckner, S.W. Size-dependent nanoparticle reaction enthalpy: Oxidation of aluminum nanoparticles. J. Phys. Chem. Solids 2011, 72, 719–724. [Google Scholar] [CrossRef]
  53. Dupiano, P.; Stamatis, D.; Dreizin, E.L. Hydrogen production by reacting water with mechanically milled composite aluminum-metal oxide powders. Int. J. Hydrog. Energy 2011, 36, 4781–4791. [Google Scholar] [CrossRef]
  54. Xiao, F.; Guo, Y.; Li, J.; Yang, R. Hydrogen generation from hydrolysis of activated aluminum composites in tap water. Energy 2018, 157, 608–614. [Google Scholar] [CrossRef]
  55. Jia, Y.; Shen, J.; Meng, H.; Dong, Y.; Chai, Y.; Wang, N. Hydrogen generation using a ball-milled Al/Ni/NaCl mixture. J. Alloys Comp. 2014, 588, 259–264. [Google Scholar] [CrossRef]
  56. Ilyukhina, A.V.; Ilyukhin, A.S.; Shkolnikov, E.I. Hydrogen generation from water by means of activated aluminum. Int. J. Hydrog. Energy 2012, 37, 16382–16387. [Google Scholar] [CrossRef]
Figure 1. Hydrolysis activation and H2 collection apparatus.
Figure 1. Hydrolysis activation and H2 collection apparatus.
Applsci 12 05378 g001
Figure 2. Activated μAl (as defined in Equation (5)) versus mass % of Al NPs.
Figure 2. Activated μAl (as defined in Equation (5)) versus mass % of Al NPs.
Applsci 12 05378 g002
Figure 3. Total amount of active Al0 per unit mass of Al NPs as a function of mass % of Al NPs in the sample.
Figure 3. Total amount of active Al0 per unit mass of Al NPs as a function of mass % of Al NPs in the sample.
Applsci 12 05378 g003
Figure 4. (a) Tracking of total H2 gas volume formed by the hydrolysis of 50 mg-μAl/Al NPs composite samples. (b) Dependence of initial H2 formation rate on mass % of Al NPs.
Figure 4. (a) Tracking of total H2 gas volume formed by the hydrolysis of 50 mg-μAl/Al NPs composite samples. (b) Dependence of initial H2 formation rate on mass % of Al NPs.
Applsci 12 05378 g004
Figure 5. PXRD pattern for the white solid material recovered after 3 days hydrolysis of the 50% Al NPs/50% μAl composite sample. Peaks labeled with * are due to elemental Al (s) (fcc Al, ICDD #00-004-0787) and all remaining peaks are attributable to Al(OH)3 (s) (bayerite, ICDD #01-074-1119). The three strongest peaks for each identified phase are labeled with their Miller indices.
Figure 5. PXRD pattern for the white solid material recovered after 3 days hydrolysis of the 50% Al NPs/50% μAl composite sample. Peaks labeled with * are due to elemental Al (s) (fcc Al, ICDD #00-004-0787) and all remaining peaks are attributable to Al(OH)3 (s) (bayerite, ICDD #01-074-1119). The three strongest peaks for each identified phase are labeled with their Miller indices.
Applsci 12 05378 g005
Figure 6. Dependence of reaction solution pH on Al NP mass %.
Figure 6. Dependence of reaction solution pH on Al NP mass %.
Applsci 12 05378 g006
Figure 7. (a) Compositional fractions of Al3+ species in aqueous solution as a function of pH (see Appendix A). (b) Total concentration of all dissolved Al species in solution as a function of pH (see Appendix B).
Figure 7. (a) Compositional fractions of Al3+ species in aqueous solution as a function of pH (see Appendix A). (b) Total concentration of all dissolved Al species in solution as a function of pH (see Appendix B).
Applsci 12 05378 g007
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Kader, M.S.; Zeng, W.; Johnston, E.; Buckner, S.W.; Jelliss, P.A. A Novel Method for Generating H2 by Activation of the μAl-Water System Using Aluminum Nanoparticles. Appl. Sci. 2022, 12, 5378. https://doi.org/10.3390/app12115378

AMA Style

Kader MS, Zeng W, Johnston E, Buckner SW, Jelliss PA. A Novel Method for Generating H2 by Activation of the μAl-Water System Using Aluminum Nanoparticles. Applied Sciences. 2022; 12(11):5378. https://doi.org/10.3390/app12115378

Chicago/Turabian Style

Kader, Mohammad S., Wenhui Zeng, Elisabeth Johnston, Steven W. Buckner, and Paul A. Jelliss. 2022. "A Novel Method for Generating H2 by Activation of the μAl-Water System Using Aluminum Nanoparticles" Applied Sciences 12, no. 11: 5378. https://doi.org/10.3390/app12115378

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