Kanica
Sharma
and
Tejwant Singh
Kang‡
*
Department of Chemistry, UGC Centre for Advanced Studies (CAS-II), Guru Nanak dev University, Amritsar-143005, India. E-mail: tejwantsinghkang@gmail.com
First published on 15th November 2024
The preparation of nanomaterials employing ionic liquids (ILs) and surface active ionic liquids (SAILs) in a relatively sustainable manner for different applications is reviewed. ILs offer structure directing and templating effects via inherent bi-continuous structures formed by the segregation of polar and non-polar domains. On the other hand, SAILs offer a structure-directing effect governed by their ability to lower the surface tension, self-assembling nature and interaction with precursors via ionic head groups. Binary mixtures of ILs with other relatively greener solvents or utilization of metal-based ILs (MILs), which act as precursors of metal ions, templates and stabilizing agents propose a new way to prepare a variety of nanomaterials. The introduction of SAILs that exfoliate 2D materials under low-energy bath sonication and also aid in photoreduction and stabilization of photocatalytically active nanomaterials at the surface of 2D materials poses a distinctive perspective in sustainable preparation and utilization of nanomaterials in different photocatalytic applications. The present feature article reviews the employment of distinctive properties of ILs in precise morphological control of nanomaterials, and their after-effects on their catalytic efficiencies.
Similarly, different approaches are utilized to prepare or exfoliate 2D materials. For example, the preparation of graphene derivatives (graphene oxide and reduced graphene oxide) is generally achieved by Hummers’ process, which involves H2O2, HNO3, H2SO4, and KMnO4, with the process taking hours to complete (more than 5–6 steps).35 Other methods involve the use of conventional ionic surfactants and organic solvents for liquid phase exfoliation of 2D materials.36 Surfactants lower the interfacial tension and intercalate between the layers of the 2D materials dispersed in water, a requisite for exfoliation.16,37 However, the point charge at the ionic head group in conventional surfactants reduces their efficacy in exfoliating 2D materials whereas most of the organic solvents are toxic.
The use of chemicals is bound to chemistry, however, as humans and prominent receivers in biological ecosystems, it is our responsibility to adopt environment-friendly yet efficient strategies for experimental design based on the “12 principles of Green Chemistry”.38,39 The combination of nanotechnology and sustainability appears promising and is expected to provide the scientific community with novel discoveries and useful applications while least affecting the environment. Many approaches have been reported in recent times employing less harmful reagents in terms of their impact on the environment and providing distinctively featured nanomaterials.40–43 In this regard, the usage of plants or other biological extracts44 has been employed to prepare a variety of noble metal NPs,45 semiconductor nanocrystals,46 and quantum dots.47 The chemical entities in plant extracts act as templating and reducing agents while exerting marginal control over the size and structural features owing to their inability to control the rate of nucleation and growth of the NPs. Many times, varying compositions of the extracts result in inconsistency in the properties of NPs and hence reproducibility is a major concern.
During the past 2 decades, ionic liquids (ILs)48 have attracted great interest from researchers as a new class of benign solvents for diverse applications,49–55 alternative to VOSs. ILs were employed as suitable media for the preparation of nanoparticles (NPs)56–63 stabilized by the inherent bi-continuous structures of ILs.64–67 The tailoring of alkyl chains of ILs offered an opportunity to prepare a new class of surfactants called surface active ionic liquids (SAILs). SAILs lower the surface tension of water and other solvents and undergo self-assembly in a diverse array of solvents akin to conventional ionic surfactants.68–70 SAILs many a time show better surface activity owing to large-sized ionic head groups with delocalized charge and functionalized alkyl chains.71–73 SAILs exert electrostatic, hydrophobic, and non-covalent interactions, which play an important role in controlling the surface activities and hence, the aggregation of SAILs, which are not possessed by conventional ionic surfactants with point charges. For example, imidazolium-based SAILs have charge delocalization in the imidazolium head group which renders it less polarized and results in decreased electrostatic repulsions among the head groups and facilitates aggregation at relatively lower critical micelle concentration (cmc) as compared to conventional ionic surfactants.74 Non-covalent interactions (hydrogen bonding, π–π stacking) are considered to be responsible for compact packing of aromatic head groups in micellar aggregates.75 Hence, all these structural features of SAILs account for their enhanced surface activity and result in the formation of stable aggregates not only in an aqueous medium but also in other solvents and then provide stabilizing and templating effects for NP preparation.76 ILs and SAILs are useful in introducing anisotropy in NPs77 by differential capping of the ions and providing a surrounding medium of distinctive viscosity and surface tension that affects the rate of nucleation and growth of NPs and their resulting properties.78,79
With the passage of time, novel approaches are being proposed involving minimum use of chemicals to achieve the desired product in a facile manner. Keeping these considerations in mind along with the limitations of the conventional methods and the novelty brought by ILs in the fields of organic synthesis, catalysis, and nanotechnology, this feature article covers the synthesis and utilization of different ILs and SAILs for the preparation of a variety of nanomaterials (mixed ferrites, ferrites, noble metal and noble metal based NPs) followed by their conjugation with 2D materials (graphene and MoS2) exfoliated in situ, for enhanced photo-catalysis in a sustainable manner under visible light or sunlight (Scheme 1).
![]() | ||
| Scheme 1 IL and SAIL-assisted preparation of a variety of NPs, their composites, and catalysis for water remediation. | ||
The future scope of ILs, SAILs and similar solvents in the area of material science is discussed thereafter. It is expected that this article will offer a new perspective to design efficient, benign and novel approaches in the field of nanoscience and nanotechnology employing ILs or SAILs for diverse applications not limited to photocatalysis but for futuristic applications such as value addition to CO2, photocatalytic water splitting, solar fuel cells and batteries, and biological applications in bio-reactive sensors and drug delivery systems.
The surface activity of SAILs was exploited for the synthesis of IL-capped AgBr nanocrystals with size and shape controlled by alkyl chain length of the IL via a hydrothermal method by Lou et al.92 The stabilizing ability of SAILs is believed to be due to their surface activity, which lowers the surface tension, viscosity, and dielectric constant of the surrounding medium resulting in slow Ostwald ripening and hence, fast nucleation and slow aggregation leading to small-sized NPs. In the case of metal NPs, ionic head groups of SAIL tend to interact with the metal ions and direct the reduction and growth of the NPs.93,94 Thus, ILs and SAILs form a useful class of new solvents for providing a template to the NPs in relatively less toxic reaction conditions with minimal use of chemical reagents to carry out the reaction. The forthcoming sections cover IL or SAIL-assisted sustainable preparation of a variety of NPs and their heterostructures with 2D (graphene and MoS2)-materials majorly reported by our research group for their photocatalytic applications.
![]() | ||
| Fig. 1 (A) Schematic representation of the synthesis of La0.9Ce0.1Fe1−xCoxO3 NPs; and (B) variation in morphologies of the NPs with changes in concentration of IL from 0 mM to 200 mM, and with change in dopant concentration as x = 0 and x = 0.5. Adopted from ref. 108 with permission. Copyright 2015 The Royal Society of Chemistry. | ||
The effect of the type of dopant was further probed by doping La–Ce ferrites with Mn2+ in place of Co2+ to prepare La0.9Ce0.1Fe1−xMnxO3110 adopting a similar approach using [C16mim][Cl] at a fixed concentration of 50 mM, which was higher than the reported cmc of the SAIL.109 In the presence of a SAIL micellar template, directional growth was favored leading to the formation of elongated rod-like structures. An increase in the concentration of the dopant, from x = 0 to x = 0.3, led to a shift from weak ferromagnetism to paramagnetism (Fig. 2), which also prevented agglomeration. The results obtained from Mössbauer studies were also found to be in accordance with the effect of the amount of Mn-doping on the magnetic and hence structural properties of the prepared nano-ferrites (Fig. 2). The appearance of a central doublet in the Mössbauer spectra with increasing amount of Mn-doping corresponds to the paramagnetic behavior of the NPs depending on the particle size. With smaller crystallite size, elongated rod-like morphology, and more surface exposure, these nano-ferrites are expected to be highly efficient in various photosensitive applications.
![]() | ||
| Fig. 2 (A)–(C) Mössbauer spectra; and (D) hysteresis loops of the prepared La0.9Ce0.1Fe1−xMnxO3 NPs employing [C16mim][Cl] (50 mM) via a hydrothermal method. Adopted from ref. 110 with permission. Copyright 2017 Elsevier B.V. | ||
Another substitute for Fe3+ ions for doping of La–Ce ferrites is Cr, due to its existence in multiple oxidation states dependent upon different methods followed for the preparation of ferrites. Super-exchange interactions between Fe2+ and Cr2+ ions are known to take place in the ordered B-site resulting in enhanced magnetic properties of the formed nanoferrites.111 Cr-doped La–Ce nano-ferrites were already reported,106 before our group exploited the surface active properties of SAIL, [C16mim][Cl], for the hydrothermal synthesis of the Cr-doped nano-ferrites and investigated their magnetic and optical properties.112 The concentration of Cr doped in La0.9Ce0.1Fe1−xCrxO3 ferrites was found to affect the crystallite size of the NPs, which first decreased with an increase in the concentration of dopant from x = 0 to x = 0.3, but then increased at x = 0.5 because of the formation of a secondary phase of La2CrO6. Here, the SAIL was found to be responsible for the formation of La2CrO6 along with the desired nano-ferrites.
At low content of Cr-doping, directional growth was observed leading to the formation of rod-like structures of nano-ferrites. At higher concentrations, an increase in magnetism increased the rate of agglomeration of the particles, which ultimately resulted in thick rods with induced branching in the nano-ferrites. Magnetic behavior was also prominent at low concentrations, but a decrease in magnetism was observed at higher dopant concentration (x = 0.5).
Binary mixtures of ILs with polar solvents offer interesting dynamics for affecting the interactions responsible for size control and structural orientation of NPs. The energetics and structure of the binary mixture of 1-ethyl-3-methylimidazoliumethylsulfate, [C2mim][C2OSO3] and water were studied by volume, density, and viscosity studies, which explain the segregation of IL in the IL-rich and water-rich regions, respectively.126–128 Inspired by the presence of bi-continuous structures in binary mixtures of IL with water, our group has introduced a greener approach to prepare photosensitive α-Fe2O3 NPs in binary mixtures of 1-ethyl-3-methylimidazoliumethylsulfate, [C2mim][C2OSO3] and ethylene glycol (EG)129via simple grinding while avoiding high-pressure conditions. The effect of IL on the structural, magnetic, and optical properties of the NPs was examined by varying the percentage of IL in the binary mixture from 0 to 25, 50, 75, and 100% (Fig. 3). H-bonding interactions between the cationic head group of the IL with –OH groups of EG, and differential aggregation behavior, depending on the content of IL in the binary mixtures, accounted for the stabilization and affected the characteristic properties of the formed hematite NPs (Fig. 3).
Easy preparation avoiding any sophisticated instrument added to the economic considerations of the proposed synthetic route. It was observed that dispersion forces dominate in the mixture, but with an increase in the content of IL, the interactions between IL and EG are compensated by increased interactions of constituent ions of IL and EG with the Fe3+ ions and α-Fe2O3 NPs. This led to greater capping by the IL and restricted growth of the particles, resulting in decreased crystallite size of the NPs with the increase in IL concentration (Fig. 4). The structural, optical and magnetic properties of the prepared α-Fe2O3 NPs were investigated (Fig. 4).
![]() | ||
| Fig. 4 (A) TEM images; and (B) Mössbauer spectra of α-Fe2O3 NPs prepared at different IL concentrations in an IL:EG binary mixture. 0, 50 and 100 in (A) and (B) represent the content of IL in the binary mixture, respectively. Adopted from ref. 129 with permission. Copyright 2015 The Royal Society of Chemistry. | ||
Due to forbidden d–d transitions and lattice relaxations, α-Fe2O3 in the bulk form does not exhibit luminescence.130 However, tuning the size of particles to the nano-size range results in a quantum confinement effect along with reduced magnetic interactions and loss of long-range order, leading to the emission of luminescence by the NPs.131 The characterization of the prepared α-Fe2O3 NPs was done by Mössbauer spectroscopy where the appearance of a central quadrupole doublet along with the characteristic sextet, corresponds to the development of superparamagnetic behavior in the NPs with increasing IL content and decreasing crystallite size (Fig. 4).
A broader central doublet in the Mössbauer spectrum of α-Fe-50 than α-Fe-100 indicates a smaller crystallite size of the α-Fe-50 NPs.
![]() | ||
| Fig. 5 (A)–(C) TEM images of α-Fe-4, α-Fe-8 and α-Fe 16, respectively (insets show their respective particle size distribution); (D) UV-vis absorption spectra of aq. RhB in the presence of α-Fe-16 under sunlight at different time intervals and corresponding images showing dye degradation; (E) rate of dye degradation and (F) recyclability of α-Fe-16 as a photocatalyst for 4 repeated catalytic cycles. Adopted from ref. 133 with permission. Copyright 2019 The Royal Society of Chemistry. | ||
Stoichiometric amounts of Fe(NO3)3·9H2O and [Cnmim][FeCl4] IL were mixed and ground in a pestle-mortar at room temperature (RT) in the presence of NaOH. The obtained paste was then washed, dried, and calcinated at 600 °C for 4 h to get the final product. α-Fe2O3 NPs were photo-catalytically active towards RhB degradation under sunlight. Although there were earlier reports concerned with IL-assisted preparation of α-Fe2O3 NPs via solvothermal or hydrothermal approaches,134,135 the impact of the structural features of the MIL was never discussed in controlling the properties of α-Fe2O3 NPs. As mentioned in our work, the FeCl4− ion is supposed to remain in close vicinity to the imidazolium head group and away from the hydrophobic alkyl chain. Thus, oxidation and nucleation of Fe3+ ions for the preparation of α-Fe2O3 were believed to take place in the polar domain of the bi-continuous assembly of the IL. This predominantly impacted the shape and size control of NPs by changing the alkyl chain length of the non-polar domain (Fig. 5). The prepared NPs (α-Fe-16) (16 represents the alkyl chain length of the MIL used) were able to decompose ∼98% dye in 105 min under sunlight exposure, while the rate of dye degradation decreased with decreasing length of the alkyl chain of the MIL used for the preparation of NPs. Sunlight was necessary as the reaction did not take place under white light, which confirms the role of sunlight in conjunction with α-Fe2O3 NPs, owing to a large band gap of the prepared NPs.
Numerous methods are reported for the preparation of a variety of QDs,6,47,131,136 but greener and more sustainable approaches are always looked upon for enhanced utilization of such materials in various applications. Our group was the first one to report137 the preparation of ZnS QDs using Zn containing IL via a one-pot facile approach (Fig. 6).
![]() | ||
| Fig. 6 (A) Schematic representation of the synthesis of ZnS QDs stabilized by [Cnmim][ZnCl3]; (B) TEM image, (C) particle size distribution, (D) HR-TEM image showing lattice planes of ZnS QD-C8; (E) absorption spectra of an aqueous solution of RhB in the presence of QD-C8 exposed to UV irradiation for different time intervals; (F) photocatalytic performances of the different QDs and (G) comparison of the reaction rates together with the rate constants for the different QDs for degradation of RhB in aqueous solutions. Adopted from ref. 137 with permission. Copyright 2017 The Royal Society of Chemistry and the Centre National de la Recherche Scientifique. | ||
The synthetic procedure involved mixing and grinding equimolar amounts of ZnS and 1-alkyl-3-methylimidazolium trichlorozincate [Cnmim][ZnCl3] (n = 4, 8, 16) IL in a pestle-mortar to yield ZnS QDs (QD-C4, QD-C8, QD-C16, where 4, 8, 16 denote the alkyl chain length of the IL used). MIL plays three important roles in the preparation of the QDs, (i) providing a medium for the reaction to take place, (ii) acting as a precursor of Zn2+ ions for the reaction, and (iii) acting as a templating agent to control the size and shape of the QDs. The length of the alkyl chain of the IL affects the dielectric constant, viscosity, and surface tension of the medium. This in turn affects the rate of nucleation and rate of growth of the NPs and hence controls the rate of agglomeration and particle size thereafter. It was observed that the smallest crystallite size forms in the case when [C16mim][ZnCl3] MIL was used for the preparation of ZnS QDs. The segregation of polar and non-polar domains of the MIL was proposed to affect the size and shape of the QDs.
Optical properties of the ZnS QDs were found to be dependent on the particle size, as observed from the photoluminescence and photocatalytic studies. The prepared QDs were able to decompose rhodamine B dye in an aqueous solution under UV-light irradiation. The activity of the QDs followed an increasing trend with an increase in chain length of the employed MIL up to n = 8, owing to a decrease in the crystallite size. However, despite being the smallest in size, the probability of residual MIL and the highest band gap energy renders QD-C16 less efficient in photocatalysis as compared to QD-C8 (Fig. 6). QD-C8 showed maximal activity towards photodegradation of the dye with 100% degradation in 75 minutes.
![]() | ||
| Fig. 7 (A) Conventional, (B) metal–halide mediated and (C) SAIL mediated strategies for the preparation of Ag@AgX NPs; (D) and (E) show the photographs of formation and non-formation of Ag@AgBr JNPs under different conditions, respectively; (F) TEM; and (G) HR-TEM images of Ag@AgBr JNPs; and (H) shows the variation of the hydrodynamic diameter (Dh) along with the ζ-potential of Ag@AgBr JNPs as a function of time when stored in the dark at 25 °C. Adopted from ref. 146 with permission. Copyright 2019 The Royal Society of Chemistry. | ||
The economic and environmental viability of the procedure lies in the minimal use of chemicals as the SAIL used acts as the precursor for providing the Br− ion required for the formation of AgBr, a templating agent for the capping of the NPs, and the reducing agent for the reduction of Ag+ ions to Ag0. The shape and size control by the SAIL were exquisite as distinguished multi-phased and highly stable Janus-shaped NPs (JNPs) were obtained with this methodology (Fig. 7). The SAIL concentration used for the JNP synthesis was below the reported cmc,147 which means that SAIL provided the templating effect to the growing Ag@AgBr NPs in the monomeric form. The interactions of the photosensitive nicotinium cation with the Ag+ ions on the AgBr surface resulted in in situ reduction and symmetry-breaking induced generation of Janus-shaped Ag/AgBr NPs under sunlight. The NPs were found to be photo-catalytically active for the efficient degradation of model water-pollutant such as RhB dye and reduction of 4-nitrophenol under sunlight with a relatively higher rate constant (Fig. 8). A series of Ag@AgBr NPs were synthesized employing varied concentrations of the SAIL and AgNO3. However, the highest photocatalytic activity was shown by Ag/AgBr-1 JNPs (prepared by using [AgNO3] = 10 mM and [SAIL] = 2 mM), which showed distinguished structural anisotropy (Fig. 7). The Ag/AgBr-1 JNPs were able to degrade ∼100% of the RhB dye in 25 minutes of sunlight irradiation with a rate constant of 0.15 min−1 (Fig. 8). Here, we proposed a new sustainable approach for synthesizing photo-catalytically active nanomaterials with enhanced structural anisotropy using relatively greener alternatives. The novelty of the process lies in the non-requirement of separate reducing and capping agents, and complex media for precise shape and size control on the NPs.
![]() | ||
| Fig. 8 (A) Time-dependent UV-vis absorbance spectra of an aqueous solution of RhB in the presence of Ag/AgBr-1 as a catalyst; (B) catalytic performances of different NPs and (C) recycling efficiency of the Ag/AgBr-1 catalyst for 5 repeated catalytic cycles; and (D) represents the plausible mechanism similar to that reported earlier and the band gap values of Ag/AgBr-1 JNPs mentioned in the scheme has been calculated using UV-visible measurements. Adopted from ref. 146 with permission. Copyright 2019 The Royal Society of Chemistry. | ||
Taking into consideration the applicability of NCs of noble metals with semiconductors, multicomponent NCs can be tried for their distinctive structural and morphological properties. Au–Ag@AgBr was first synthesized by the Au seed-mediated method using a high-intensity Hg lamp as the energy source with PVP acting as the capping and reducing agent in the presence of EG as the solvent.148 Another group reported the synthesis of Au/Ag@AgBr NCs using CTAB as a capping agent and Au seed particles to induce nucleation and growth of the nanoparticles with insignificantly induced anisotropy.149 Both reports explained the formation of AgBr on the pre-formed Au seeds in their respective environments, followed by in situ reduction of Ag+ to Ag0. We were curious to understand the dynamics of stabilization of noble metal-based NCs assisted by surface-active agents, which lead to anisotropic growth of the particles. Our group reported the preparation of AuAg alloy@AgBr JNPs with precise shape and size control offered by CTAB as the capping agent in aqueous medium and reduction of the metal ions induced by a natural source, ascorbic acid, under sunlight exposure.150 We achieved control over the morphology of the NPs by variation in the pH of the reaction medium. The viability of this report lies in the pH-dependent dissipative self-assembly of CTAB-stabilized AgBr colloids in a non-conventional top-down approach, leading to shape and size controlled by the interactions prevailing among the surfactant, metal ions and reducing agent in the reaction medium (Fig. 9). The synthesis involved the preparation of AgBr colloids stabilized by CTAB, followed by simultaneous reduction of Ag+ and Au3+ in the presence of sunlight. Precisely multi-phased Janus-shaped nanoparticles were observed from the TEM with size-dependent optical properties.
![]() | ||
| Fig. 9 (A) Diagrammatic representation of the mechanism followed during the preparation of JNP-1 and JNP-2; (B) and (C) TEM and HR-TEM images of JNP-1; and (D) and (E) TEM and HR-TEM images of JNP-2. Adopted from ref. 150 with permission. Copyright 2024 The Royal Society of Chemistry. | ||
JNP-1 and JNP-2 were prepared under the same experimental conditions with a difference in pH of the medium (JNP-1 at pH ∼ 4.2 and JNP-2 at pH ∼ 9.4). The prepared Janus-shaped AuAg alloy@AgBr NPs were found to be photo-catalytically efficient in sunlight-driven degradation of RhB dye as a model pollutant (Fig. 10), and for bacterial growth inhibition with very low minimum inhibitory concentration (∼35 μg mol−1).
![]() | ||
| Fig. 10 (A) Time-dependent UV-vis absorbance spectra showing the degradation of RhB in the presence of JNP-2; (B) catalytic activities of JNP-1 and JNP-2; (C) recyclability of JNPs; and (D) schematic representation of the formation of charge carriers involved in the photocatalysis. Adopted from ref. 150 with permission. Copyright 2024 The Royal Society of Chemistry. | ||
![]() | ||
| Fig. 11 (A) TEM image of an exfoliated graphene sheet; (B) TEM image of the prepared α-Fe2O3@G NCs; (C) photocatalytic antibiotic sulfamethoxazole (SMX) degradation efficiency of graphene, α-Fe2O3, and α-Fe2O3@G and their corresponding apparent quantum efficiencies (AQEs) under visible light; and (D) schematic representation showing the possible mechanism for electron transfer for the degradation of SMX antibiotic. Inset of (A) shows the SAED pattern of exfoliated graphene sheets. Adopted from ref. 156 with permission. Copyright 2020 The Royal Society of Chemistry and the Centre National de la Recherche Scientifique. | ||
Another type of semiconductor-graphene NC reported by our group was ZnS@G (Fig. 12). The low-energy bath sonication exfoliation of graphene in aqueous medium assisted by surface-active Zn-containing SAIL, [C16mim][ZnCl3] followed by in situ decoration of ZnS QDs results in ZnS@G NCs for photocatalytic applications.157 The exfoliated graphene flakes were 0.5–1 nm in height and up to 5-layered thick with negligible lattice defects. In situ decoration of graphene flakes with ZnS QDs was done in the aqueous colloidal dispersion of graphene having 50 mM [C16mim][ZnCl3]. It was mixed with equimolar amounts of Na2S and subjected to magnetic stirring for 30 minutes at room temperature to obtain the ZnS@G NCs. The band gap calculated from the Tauc plot showed a decrease from 3.45 eV for ZnS QDs (prepared under the same reaction conditions in the absence of graphene) to 3.09 eV for ZnS@G NCs. Lowering of the band gap of ZnS@G NCs was attributed to the graphene, which prevented the electron–hole recombination and resulted in an increased electron density in the CB of ZnS, responsible for the enhancement of the photocatalytic activity of the NCs in visible light. Photoluminescence quenching observed in the presence of graphene confirmed the role of graphene in preventing electron–hole pair recombination. The prepared ZnS@G NCs were able to degrade RhB dye completely and antibiotic ciprofloxacin (CIP) under sunlight exposure, showing better catalytic activity than bare ZnS QDs reported earlier137 which were photo-catalytically active in UV light (Fig. 12).
![]() | ||
| Fig. 12 (A) Schematic representation of exfoliation of graphene and subsequent preparation of ZnS QD decorated graphene; (B) TEM and (C) HR-TEM images of an exfoliated graphene sheet; (D) TEM image of the prepared ZnS@G NCs; (E) catalytic performances of ZnS@G, ZnS and graphene and their rate constants; and (F) photocatalytic CIP removal % of graphene, ZnS and ZnS@G catalysts and their corresponding apparent quantum efficiencies (AQE) under visible light. Inset of (A) shows the SAED pattern of exfoliated graphene sheets. Adopted from ref. 157 with permission. Copyright 2020 The Royal Society of Chemistry. | ||
Although the activity of Ag/AgBr NPs was significantly enhanced compared to that of both Ag and AgBr NPs, their conjunction with 2D materials such as graphene is a field of interest. Prompted by the heterojunction formation between graphene and NPs supported by a large surface area of graphene, we combined the photocatalytic properties of Ag/AgBr NPs with a large surface area of graphene in a two-step one-pot sustainable strategy. Arora et al.158 employed a nicotinium-based [C12ENic][Br] SAIL, reported earlier147 by our group, for exfoliation of graphene in an aqueous medium, followed by in situ decoration of Ag/AgBr JNPs on the exfoliated graphene sheets (Fig. 13) under sunlight. Not only this, the conjugated Ag/AgBr@graphene NCs showed enhanced photocatalytic efficiency under sunlight. Maximum exfoliation of graphene was achieved at [SAIL] = 10 mM, near to its cmc,147 which points towards the dominance of the monomeric form of SAIL in graphene exfoliation. Graphene sheets of 2.4 nm thickness were obtained via low-energy bath sonication in good yield with negligible lattice defects and appreciable colloidal stability for up to 15 days as observed by a highly positive zeta potential value (+35.5 mV). The same SAIL-stabilized aqueous dispersion of graphene was used for in situ generation of Ag/AgBr JNPs, after diluting it to get a final SAIL concentration of 2 mM as it was reported to produce Ag/AgBr JNPs with effectively induced symmetry-breaking under sunlight.146 As discussed earlier, [C12ENic][Br] SAIL reduces Ag+ to Ag0 in less than 1 minute of sunlight exposure but in the presence of graphene, around 5 minutes were taken for the growth of Ag/AgBr@G NCs at the desired concentration. This was assigned to the reduced exposure of Ag+ ions towards sunlight because of the presence of graphene sheets which slowed down the rate of reduction and growth of the NPs. The prepared NCs were able to decompose 99% of ciprofloxacin within 120 minutes of light exposure as determined by HPLC measurements. Moreover, the rate of photocatalysis was enhanced by 2–3 fold for the degradation of RhB dye and 4-NP by Ag/AgBr JNPs in the presence of graphene. Enhanced photocatalytic activity and stability of the prepared Ag/AgBr@G NCs are attributed to the presence of graphene as an electron-accepting agent in conjunction with Ag/AgBr, which assists in electron–hole separation and charge transfer, along with providing extended surface area for adsorption of reactant molecules.
![]() | ||
| Fig. 13 (A) Schematic representation for in situ preparation of Ag/AgBr@G NCs and its photocatalytic activity; (B) TEM and (C) HR-TEM images of an exfoliated graphene sheet; (D) TEM image of Ag/AgBr@G NCs; (E) photocatalytic efficacy for the photodegradation of CIP and their corresponding apparent quantum efficiency (AQE%) using Ag/AgBr@G NC and their counterparts; and (F) plausible mechanism for the photocatalytic activity of Ag/AgBr@G NCs. Inset of (A) shows the SAED pattern of exfoliated graphene sheets. Adopted from ref. 158 with permission. Copyright 2021 Elsevier B.V. | ||
Progress in graphene exfoliation has led to the exploration of other 2D materials with interesting properties. Transition metal dichalcogenides (TMDs), with a general formula of MX2 (where, M is a transition metal and X = S, Se, Te), cover a wide range of materials for electronic and catalytic applications.14 The exfoliation of bulk TMDs into 2D nanosheets would account for enhancement in their structural and electronic properties owing to the quantum confinement effect.131 There have been reports of the conjugation of TMD nanosheets with other metal NPs for various applications.24,159,160 Our group reported exfoliation of MoS2 sheets assisted by a biological surfactant, choline deoxycholate [Cho][Doc], followed by in situ decoration with Ag NPs in a facile and sustainable manner.161 Maximum exfoliation of MoS2 was achieved by using 3 mM of [Cho][Doc] in an aqueous medium and then Ag@MoS2 NCs were prepared at different concentrations of [Cho][Doc] and characterized for their morphological and optical properties. We were able to exfoliate 2–8 nm thick sheets of MoS2 with good colloidal stability for at least 16 days. The in situ generation of Ag NPs took place under sunlight via [Cho]+ induced reduction of Ag+ to Ag0 and the NPs capped by [Cho][Doc] (Fig. 14).
![]() | ||
| Fig. 14 (A) Molecular structure of [Cho][Doc] and exfoliation of MoS2 in aqueous [Cho][Doc]; (B) TEM image showing Ag NPs adhered to the MoS2; (C) time-dependent UV absorbance of RhB in the presence of Ag@MoS2(10) under sunlight; and (D) comparison of the kinetics of Ag@MoS2(10) NCs, Ag@MoS2(25) NCs and Ag@MoS2(50) NCs. Adopted from ref. 161 with permission. Copyright 2024 Elsevier B.V. | ||
The size of the NPs was found to decrease with increasing concentration of [Cho][Doc] used in the preparation (10, 50 and 100 mM). The interaction of Ag NPs with MoS2 seems to be reduced at high concentrations of the surfactant. At higher concentrations of [Cho][Doc], the number of free micelles increases and Ag+ ions orient towards the negatively charged [Cho][Doc] micelles and not with the monomers intercalated among the MoS2 sheets. Hence, more of the free Ag NPs are observed when 50 mM [Cho][Doc] is used, in contrast to the appreciable amount of Ag@MoS2 NCs observed in the case of 10 mM [Cho][Doc] used. Thereafter, the photocatalytic efficiency of Ag@MoS2(10) towards the degradation of RhB under sunlight was observed to be the best among all other NCs prepared. The photocatalytic efficiency of the NCs is attributed to the involvement of reactive oxygen species in oxidizing the organic dye (RhB) in the aqueous medium under sunlight. The NCs were found to be catalytically stable for up to 5 cycles with unaltered activity.
The journey of finding new ways of synthesizing more efficient and sustainable ILs does not end. We have reported the synthesis of a functionalized SAIL based on nicotinium cation,147 and its application in the sunlight-driven preparation of Ag/AgBr JNPs with induced symmetry-breaking induced by the templating effect of the SAIL in aqueous medium.146 The economic and environmental viability of this work lies in the minimal use of harsh chemical reagents and energy sources to carry out the synthesis of anisotropic Ag/AgBr JNPs with enhanced photocatalytic activity. The photocatalytic activity of these NPs has been found to be better than bare Ag and AgBr NPs, respectively, owing to Fermi-level equilibration of electronic bands of Ag and AgBr as well as striking anisotropy in the shape control in the JNPs. The same SAIL has also been reported to be efficient in exfoliating graphene (G) nanosheets in aqueous medium via bath sonication and in situ decoration of Ag/AgBr JNPs onto exfoliated G to form Ag/AgBr@G NCs.158 The band gap tuning by conjunction with graphene leads to enhancement in photocatalytic efficiency compared to the previously prepared bare Ag/AgBr JNPs. The templating effect of surface active agents in controlling the shape and size of the nanoparticles has been explained by our group in a recent report on the synthesis of CTAB-stabilized AuAg alloy@AgBr JNPs reduced by ascorbic acid in sunlight.150 Dissipative self-aggregation behavior of CTAB micelles owing to the metal-surfactant interactions at different pH seems to induce symmetry-breaking and anisotropy in the Janus-shaped NPs, also affecting their photocatalytic and antibacterial properties. The impact on the optical properties of noble metal NPs in multicomponent NCs such as these open doors to new combinations of metal–semiconductor NCs for their band gap tuning and photosensitivity in visible light. We have summarized the preparation of Ag@MoS2 NCs stabilized and reduced by a bio-based surfactant [Cho][Doc] in an aqueous medium under sunlight irradiation.161 The surfactant was used for liquid phase exfoliation of MoS2 nanosheets followed by in situ reduction of Ag NPs on the 2D nanosheets to form the photocatalytically active Ag@MoS2 NCs. The area of research concerned with the sustainable development of various nanomaterials assisted by ILs is not limited but continues to grow with novel methodological approaches introduced with the discovery of new ILs. The collaboration of the fields of SAILs and nanotechnology seems to be highly anticipated and a successful one for those involved in the development of new methodological approaches for sustainable photocatalytic systems to cater to the needs of alternate energy sources, such as to harvest energy sources such as sunlight. The conversion of solar energy into thermal energy to provide the activation energy for up-hill reactions has been successfully employed in organic synthesis.162,163 Plasmon heating and the generation of charge carriers on the absorption of solar energy account for the enhancement in the rate of reactions, adding to the sustainability and feasibility of the processes.164 There are distinctive nanomaterials, 2D materials, metal nanoparticles and their composites available, and their photosensitive properties have been employed in organic synthetic reactions by harvesting light energy into chemical energy.4,165–167 The advent of photocatalytic nanomaterials for organic reactions marks the transformation of conventional complicated approaches to relatively greener ones for the preparation of organic molecules of commercial importance. We aim to design more photosensitive and competent nanomaterials with precise size and morphological control using greener reagents and reaction conditions and to utilize them in significant photocatalytic processes (Scheme 2).
The Ag/AgX nanoparticles are well-known photocatalysts owing to the light-sensitive nature of AgX as already discussed. The excellent photocatalytic abilities of AgCl and AgBr have inspired the scientific community to design and investigate the properties of mixed halides. Ternary silver halides (AgClxBr1−x) are known to have optical properties and photosensitivity depending on the composition of the material.168 Numerous reports are available in the literature concerned with the synthesis of AgClxBr1−x NPs and investigating their optical properties for photocatalytic applications.169–171In situ generation of Ag@AgClxBr1−x NCs has been reported recently for improved photocatalytic activities.172–174 However, the methods available to date for the synthesis of AgClxBr1−x based NCs include the use of surfactants and non-polar solvents to ensure proper diffusion and anion exchange in the AgX lattice for finely tuned halide composition in the lattice for the required properties. Our group aims to design greener methodologies based on IL-assisted capping and reaction media for improved photocatalytic efficiencies of such NCs along with the pre-requisite requirement of size and shape control by using SAILs.
Moreover, 2D nanomaterials are gaining sufficient attention in the field of photocatalysis. The long-term stability of photo-catalytically active materials when combined with 2D materials can be enhanced by immobilizing such materials into a hydrogel using a SAIL.175 Such systems pave the way for synthesizing catalytic nanomaterials and storing them in the form of catalytic gels for long-term durability and stability. Despite of popularity of ILs in various fields of research, some limitations are encountered in the synthesis and applications of ILs, and more green alternative solvents are coming into the frame. These are called deep eutectic solvents (DESs). DES is a mixture of two components bound by H-bonding interactions (one component is an H-bond donor and the other is an H-bond acceptor) with properties different and a lower melting point from the individual components.176 Ever since the onset of DESs, there has been remarkable interest in utilizing them in various applications.177 DESs have proven to be better alternatives for solvents in reaction media owing to their facile synthesis and environment-friendly nature. Our group has used a metal-based DES in the dissolution of polymeric materials and investigation of the regenerated material for its structural and chemical properties.178,179 We expect that the exquisite physical properties of DESs can be employed to test their ability to stabilize NPs. Our group has already started working on the mixture of DES and IL as an efficient solvent mixture for the synthesis of N-doped carbon dots (CDs) for light harvesting applications.180 A DES composed of choline-chloride and ethylene glycol was used to support the self-assembly of surface active choline oleate [Cho][Ole], which stabilized the CDs produced by gelatin as the carbon source.
There is still a lot to investigate at the smallest scale of dimensions tuned by reaction conditions for useful applications in the fields of chemical sciences that are both unknown and interesting at the same time. The need for more photocatalytic materials is important also because we need more recyclable and vast sources of energy as alternatives to conventional ones. Therefore, nanomaterials that are efficient in harvesting solar energy, visible light at best, can prove to be helpful in maintaining sustainable surroundings in catalytic reactions. This feature article paves the way for the synthesis of novel and environment-friendly molecules to provide a templating effect for efficient shape and size control of nanomaterials. The synthetic route employed for the NP formation also leaves an impact on their properties which drives the mass-scale application of nanotechnology in various fields of importance.
Footnotes |
| † This feature article is dedicated to Prof. Nobuo Kimizuka, Professor, Department of Applied Sciences, Graduate School of Engineering, Kyushu University, Japan on his 65th Birthday, who was among the pioneers in preparing IL templated hollow TiO2 and discrete gold nano-architectures. |
| ‡ Current Address: Department of Applied Chemistry, Graduate School of Engineering, Kyushu University, 744 Moto-oka, Nishi-ku, Fukuoka 819-0395, Japan. |
| This journal is © The Royal Society of Chemistry 2024 |