Jinyu
Zheng
,
Xin
Xiang
,
Dongdong
Xu
* and
Yawen
Tang
*
Jiangsu Key Laboratory of New Power Batteries, Jiangsu Collaborative Innovation Center of Biomedical Functional Materials, School of Chemistry and Materials Science, Nanjing Normal University, Nanjing, 210023, P. R. China. E-mail: ddxu@njnu.edu.cn; tangyawen@njnu.edu.cn
First published on 12th August 2024
Ultrathin nanosheets possess a distinctive structure characterized by an abundance of active sites fully accessible on their surface. Concurrently, their nanoscale thickness confers an extraordinarily high specific surface area and promising electronic properties. To date, numerous strategies have been devised for synthesizing precious metal nanosheets that exhibit excellent electrocatalytic performance. In this paper, recent progress in the controlled synthesis of two-dimensional, ultrathin nanosheets by a self-assembly mechanism using functional surfactants is reviewed. The aim is to highlight the key role of functional surfactants in the assembly and synthesis of two-dimensional ultrathin nanosheets, as well as to discuss in depth how to enhance their electrochemical properties, thereby expanding their potential applications in catalysis. We provide a detailed exploration of the mechanisms employed by several long-carbon chain surfactants commonly used in the synthesis of nanosheets. These surfactants exhibit robust electrostatic and hydrophobic effects, effectively confining the crystalline growth of metals along lamellar micelles. Moreover, we present an overview of the electrocatalytic performance demonstrated by the ultrathin nanosheets synthesized through this innovative pathway. Furthermore, it offers valuable insights that may pave the way for further exploration of more functional long-chain surfactants, leading to the synthesis of ultrathin nanosheets with significantly enhanced electrocatalytic performance.
In recent years, extensive research has been conducted on the synthesis and self-assembly of noble metal nanosheets, exploring diverse approaches that each present unique advantages and challenges. These approaches aim to achieve controlled growth of nanosheets with precise morphology, thickness, and active performance. Among these strategies, the “top-down” and “bottom-up” routes have garnered significant attention.11–13 The top-down methods focus on overcoming the interlayer van der Waals interactions within nanomaterials by harnessing external energy sources.14 This involves applying external forces to dismantle bulk nanomaterials into smaller nanosheets, often achieved through techniques such as thermal treatment, organic reactions, and oxidation etching (e.g., ball milling, lithography, chemical etching, or exfoliation processes). This approach offers precise control over size and shape, making it advantageous for high-yield mass production.15–21 Conversely, the bottom-up approach involves controlling the growth process of metals to construct and assemble into target nanosheets.22–25 This method represents a frequently used, simple, and scalable strategy for synthesizing novel nanomaterials. Key methods within this framework include gas-phase deposition, wet-chemical synthesis, solution phase synthesis (such as hydrothermal and solvent heat treatment), and template-guided self-assembly synthesis. Ultrathin two-dimensional nanomaterials possess exceptional properties attributed to their high surface area and quantum confinement of two-dimensional electrons. Among the bottom-up methods, template assistance stands out as a simple and widely adopted technique for controlling the morphology and thickness of two-dimensional noble metal nanosheets. Common templates include silica, layered clay, graphene oxide, or LDH. Physical or chemical deposition (PVD/CVD) has evolved into a versatile technology capable of achieving precise control over nanosheet growth kinetics by adjusting deposition parameters.25–37 Tan et al. provided a detailed and comprehensive review of wet-chemical synthesis and applications of non-layered two-dimensional nanosheets, emphasizing the effectiveness of wet-chemical methods in producing high-yield, large-scale, and cost-effective nanosheets. They introduced strategies such as two-dimensional template synthesis and water/solvent thermal synthesis, while also offering a glimpse into the self-assembly synthesis of two-dimensional nanomaterials and the soft template method.7,38
Although numerous reports have extensively discussed the synthesis of ultrathin two-dimensional nanosheets, there remains a significant gap in detailed reviews specifically addressing the nanosheets facilitated by the self-assembly of surfactants.39–42 Functional surfactants play a pivotal role in nanosheet synthesis, not only controlling their growth and morphology but also serving as templates to shape nanostructures with precise dimensions and geometries. By carefully selecting surfactants with specific molecular structures and fine-tuning their concentration, it is possible to manipulate the molecular arrangement to preferentially bind to designated crystal planes of metals, thereby exerting a profound influence on the nanosheets' growth kinetics.43–46 The ultimate nanostructure that emerges depends on intricate interactions between ligands, metals, and the hydrophobic tail groups of the surfactants. In the realm of catalysis, ultrathin noble metal nanosheets that are synthesized through surfactant-mediated self-assembly have demonstrated exceptional catalytic activity and selectivity across a range of chemical reactions. These nanosheets boast ultrathin architectures, manifesting exceptional quantum phenomena that greatly facilitate advancements in both electrical and photonic device technologies. Their large specific surface area markedly enhances the reactivity of electrode materials, making them ideal candidates for energy storage and conversion processes. They efficiently catalyse a diverse array of battery reactions, enhancing energy density and electrochemical reaction efficacy. Furthermore, the fully exposed active sites of these two-dimensional nanosheets enable optimized catalytic performance through precise modulation of their thickness, defects, and compositional makeup. Consequently, they serve as exceptional catalysts for electrocatalytic oxygen reduction, hydrogen evolution, alcohol oxidation, among others, and demonstrate remarkable efficacy in the photolysis of water and degradation of organic pollutants. Moreover, the heightened sensitivity of two-dimensional nanosheets to gas and biomolecular adsorption surpasses that of other materials, thereby laying the groundwork for the development of highly sensitive nanosheet-based sensors. This versatility also translates into immense potential for applications within the biomedical realm, promising a wide spectrum of innovative solutions and advancements. The surfactant is adeptly adsorbed onto the surface of the nanomaterial, effectively diminishing surface energy and tension within the solution. This not only fosters a uniform dispersion of the sample, but also prevents stacking and agglomeration due to excessive surface energy, which could otherwise obscure active sites. Furthermore, it plays a pivotal role in stabilizing the nanosheet's shape, ensuring that the desired morphology and functionality are maintained.47–51 The surfactant-driven self-assembly process is highly sensitive to surfactant concentration. As described earlier, the surfactants are driven by hydrophobic interactions to form various micelles and further condense into stable micelle phases. However, if the concentration of surfactant is increased further, the concentration of single molecules or ions in the solution cannot be significantly increased; instead, more micelles can be formed, and eventually, these micelles aggregate into large massive micelles. These unique characteristics underscore the indispensability of surfactants in fabricating uniform and controllable nanomaterials.52–56
Herein, the objective of this article is to provide a comprehensive overview of the current research landscape regarding the self-assembly synthesis of noble metal ultrathin two-dimensional nanosheets through the utilization of functional surfactants, as schematically illustrated in Fig. 1. Initially, we delve into the growth mechanism underlying the surfactant-mediated synthesis of nanosheets, highlighting significant research outcomes in this domain. Furthermore, we present a compilation of the electrocatalytic performance demonstrated by nanosheets synthesized via this approach, encompassing, but not limited to, the hydrogen evolution reaction (HER) and alcohol oxidation reaction (AOR) (Table 1). Finally, we offer a recapitulation of the present research status and advancements, alongside an exploration of the challenges and potential avenues for future exploration in this promising field.
Types | Surfactants | Nanostructures | Ref. |
---|---|---|---|
Cationic surfactants | C22N-Py (Br−) | (100)-Pd NSs | 57 |
C22N-Py (Br−) | (110)-Pd NSs | 58 | |
C22N-Py (Br−) | PdP NDs | 59 | |
C22N-Py (Br−) | (110)-PdP NSs | 60 | |
C22N-Py (Cl−) | PdIr NSs | 61 | |
C22N-Py (Cl−) | PdPtP NDs | 62 | |
C22 TAC | PdPtCu NSs | 63 | |
CTAB | Pd NSs | 64 | |
CTAB | Pt Nanowheels | 65 | |
Anionic surfactants | SDS | Pd NSs | 66 |
Zwitterionic surfactants | C22N-COOH (Br−) | (100)-Pd NSs | 57 |
C22N-COOH (Br−) | Pt NDs | 67 | |
C22N-COOH (Br−) | PtRu NDs | 68 | |
C22N-COOH (Br−) | (100)-PdP NSs | 60 | |
C22N-COOH (Br−) | PdPt NDs | 69 | |
C22N-COOH (Br−) | Pt-α-MoS3 NDs | 70 | |
C22N-COOH (Br−) | PtP NDs | 71 | |
C22N-COOH (Cl−) | Au NDs | 72 | |
CH3AuPPh3 | Au NSs | 73 | |
Nonionic surfactants | PVP | PbTe NSs | 74 |
PVP | PdPtNi NSs | 75 | |
PVP | Au NFs | 76 |
Drawing from the aforementioned reasons, researchers are able to devise an array of innovative synthetic routes for controlling nanostructures. To date, surfactants have been categorized in numerous ways, with the most frequently utilized method in research and application being the classification into nonionic and ionic types, depending on the ion types present in an aqueous solution. Both types can effectively self-assemble with metal elements, promoting their growth in a specified direction into extremely thin nanosheets.59,86,87 Within this synthetic framework, our research group has examined several surfactants that enable metal precursors to grow along special crystal planes, ultimately yielding two-dimensional ultrathin nanosheets. These surfactants possess a long-chain alkyl hydrophobic tail (illustrated in Fig. 2c and d), facilitating their self-assembly into the desired layered micelles. Functional groups, such as pyridine, carboxyl, and quaternary ammonium, serve as hydrophilic heads, guiding the epitaxial growth of the precursors into ultrathin nanosheets in a controlled direction. Notably, these surfactants also contain halogen ions (primarily Cl− and Br−), which contribute to the growth process and facilitate the formation of a stable two-dimensional structure. Furthermore, we have expanded our research to produce non-metallic doped two-dimensional nanosheets. Phosphorus can seamlessly integrate into the nanosheets at the atomic level, modulating their internal structure without compromising the original morphology. This approach offers exciting possibilities for the development of novel nanomaterials with enhanced properties and diverse applications.
Our team successfully synthesized a range of cationic surfactants, utilizing them as flexible templates to orchestrate the self-assembly into ultrathin nanosheets. In our approach, C22N-Py (Cl−) and C22N-Py (Br−), constructed from long carbon-chain hydrophobic tails and pyridine functional moieties, served as structural guiding agents. The self-assembly process is depicted in Fig. 3a, where the surfactant dissolves in water, forming a homogeneous solution. Upon introducing metal precursors, the cationic hydrophilic heads of the surfactants and the anionic metal ions collaborate to assemble into stable layered intermediates. These intermediates grow within the confinement of micelles, ultimately yielding ultrathin two-dimensional nanomaterials under the reducing agents. Fig. 3b–f show the synthesis process using C22N-Py (Br−) as templates. It is evident that their growth mechanisms are characterized by continuous epitaxial growth, initiating from smaller, slightly thicker nanosheets and gradually evolving into the desired material. Notably, the surfactant concentration plays a pivotal role in shaping the nanosheets. Within a moderate concentration range, the surfactant effectively forms a stable adsorption layer on the nanosheet's surface, effectively lowering its surface energy while meticulously regulating the crystal growth. When the molecular structure remains constant, surface tension experiences a decline with the incremental concentration of the surfactant solution within a defined limit. Within this range, the identical surfactant molecules exhibit varying impacts on crystal growth at different concentrations, governed not solely by the surfactant's inherent properties but also intricately intertwined with the dynamics and thermodynamics of crystal growth. At relatively low concentrations, adsorption is feeble, allowing for minimal inhibition of crystal growth, resulting in rapid extension in select directions to form nanodendrites.90–92 Conversely, at relatively high concentrations, the adsorption and restricting effects intensify, directing crystal growth along a specific two-dimensional pathway, yielding a uniformly shaped, closed nanosheet as illustrated in Fig. 3b. By contrast, a slightly reduced concentration gives rise to highly branched nanosheets, as observed in Fig. 3f. These findings underscore the importance of surfactant concentration in governing the growth and morphology of nanosheets.58,93
Fig. 3 (a) Schematic illustration of the cationic surfactant-templated synthesis of 2D NSs. Reproduced with permission from ref. 93. Copyright 2020, the Royal Society of Chemistry. (b) TEM images of products captured at the different aging period of Pd NSs (inset: schematic for the formation of PdNSs). Reproduced with permission from ref. 58. Copyright 2017, the Royal Society of Chemistry. (c) HAADF-STEM, elemental mapping and HRTEM images of the PdPtP NDs (inset: the corresponding Fourier diffractogram of the PdPtP NDs). Reproduced with permission from ref. 62. Copyright 2024, the Royal Society of Chemistry. (d) High-magnification, high-resolution TEM images, HAADF-STEM images and elemental mapping of the PdPtCu NSs. Reproduced with permission from ref. 63. Copyright 2019, the Royal Society of Chemistry. (e) TEM images of PdIr NDs obtained in the presence of different amounts of Br− (the added content of Br− gradually increased). Reproduced with permission from ref. 61. Copyright 2022, the Chemistry Europe. (f) TEM images obtained from the different growth periods and elemental mapping of the ultrathin PdP NSs. Reproduced with permission from ref. 93. Copyright 2020, the Royal Society of Chemistry. |
In 2022, Guo carried out a study aimed at exploring the synthesis of PdIr nanosheets using C22N-Py (Cl−). The objective was to investigate the specific surfactant fragment that had a significant impact on the degree of nanosheet branching. As depicted in Fig. 3e, the concentration of Br− added progressively from left to right resulted in a gradual decrease in the nucleation rate due to ligand exchange between Cl− and Br−. This, in turn, influenced the reaction kinetics. Based on the observed four forms, it can be deduced that a slower reaction rate tends to favor the formation of solid nanosheets with reduced branching. Consequently, within various concentrations of the same surfactant solution, the extent of nanosheet branching is primarily determined by the halogen ion content. Extending this research in 2023, we successfully synthesized PdPtP nanodendrites exhibiting ultrathin characteristics using the same surfactant (Fig. 3c). High-resolution TEM analysis revealed that the crystals were highly crystalline and grew preferentially along specific crystal planes. Additionally, we utilized commercial C22TAC to investigate the synthesis of PdPtCu nanosheets (Fig. 3d), which exhibited an impressive specific surface area and an ideally ultrathin thickness, contributing to their overall desirable properties.61–63
Anionic surfactants are a class of surfactants with the longest development history. They have a negatively charged anionic hydrophilic head group and can be divided into four categories: carboxylate, sulfate, sulfonate, and phosphate. Similar to cationic surfactants, the hydrophobic groups consist of hydrocarbon chains with different lengths. In contrast to cationic surfactants, anionic surfactants are more suitable for guiding the formation of nanosheets in alkaline environments with higher pH. Wu et al. successfully achieved the growth of vertical nanosheets on graphite using the SDS surfactant template, resulting in materials with a high specific surface area. Similarly, Kim et al. explored the preparation of nanosheets utilizing the layered intermediate phases of SDS surfactant as flexible templates, demonstrating reversible self-assembly and disassembly properties. Some researchers have delved into the study of SDBS-modified nanosheets, comparing them to their unmodified counterparts and discovering an increase in the number of surface wrinkles. Remarkably, the addition of SDBS led to a significant reduction in the thickness of the nanosheets. In the early stages of synthesis, Yu et al. also synthesized the desired catalyst, leveraging the self-assembly assisted by sodium oleate as an anionic surfactant.94–97
Our group has actively delved into the synthesis of nanosheets using a diverse range of amphiphilic surfactants. Among these surfactants, C22N-COOH (Br−) and C22N-COOH (Cl−) can be successfully synthesized through a straightforward process involving N,N-dimethyldodecylethylamine (C24H51N), bromoacetic acid, and acetonitrile. Employing such surfactants for nanosheet synthesis offers simplicity and ease of operation, resulting in nanosheets with regular shapes and ultra-thin characteristics. These nanosheets exhibit outstanding catalytic performance. In 2019, our group achieved a significant milestone by synthesizing ultrathin, oriented Pt nanosheets utilizing C22N-COOH (Br−), as depicted in Fig. 4a. The key aspect of this synthesis lies in the clever integration of quaternary ammonium salts, carboxyl functional hydrophilic heads, and the counter ion Br−. This combination effectively drives the self-assembly process. Metal ions are confined within layered micelles and preferentially bind to the (110) crystal plane, while Br− adsorbs on the (100) crystal plane, preventing epitaxial growth in that direction. This ultimately leads to the in situ reduction of ultrathin two-dimensional nanosheets. That same year, we made another remarkable achievement by growing two-dimensional ultrathin single-crystal PtRu nanodendrites in a solution phase (Fig. 4b). Here, C22N-COOH (Br−) served as both a structural guiding template and a capping agent. The long-chain tail present in zwitterionic surfactants facilitated the self-assembly of metal ions with surfactant ions through electrostatic interactions, giving rise to stable intermediates. The quaternary ammonium salts and carboxyl groups collaborated to stabilize the growth process. Additionally, Br− selectively adsorbed onto specific fcc crystal planes, effectively blocking epitaxial growth along other crystal facets. This resulted in highly branched and dispersed two-dimensional ultrathin nanomaterials, exhibiting excellent morphology and a thickness of merely 1.8 nm, as confirmed by TEM and STEM-mapping.67,68
Fig. 4 (a) Low-magnification TEM, corresponding HRTEM images of the Pt nanostructure obtained at the reaction period (inset: the corresponding FD and low-magnification TEM image) and the proposed formation mechanism for crystalline facet-directed step-by-step in-the-plane epitaxial growth of PtNDs selectively along the 〈111〉 crystallographic orientation. Reproduced with permission from ref. 67. Copyright 2019, American Chemical Society. (b) TEM, HRTEM, HAADF-STEM and corresponding elemental mapping of PtCu NDs. Reproduced with permission from ref. 68. Copyright 2019, the Royal Society of Chemistry. (c) The corresponding intensity profiles of lattice spacing at the center site (red) and edge site (blue). Atomic schematic diagram of compressive strain. (d) TEM and corresponding HRTEM images of PtPd CNDs. Reproduced with permission from ref. 69. Copyright 2023, American Chemical Society. (e) Schematic illustration for the proposed formation mechanism of AuNDs. (f) UV-vis spectra and XRD patterns of 2-MNA AuNDs, AuNPs, and 2-MNA. (g) TEM and a schematic illustration of the formation of AuNDs. Reproduced with permission from ref. 72. Copyright 2024, Elsevier. |
In 2023, we successfully prepared ultrathin PdPt nanocrystals exhibiting long-range compressive strain, utilizing the functional surfactants C22N-COOH (Br−) and C22N-COOH (Cl−). Within the curled PdPt NDs, a remarkable difference in lattice spacing is observed between the center and the edge. Notably, the lattice at the center position (depicted in Fig. 4c) exhibits contraction, and our calculations reveal an approximate 2% compressive strain. HRTEM images of the edge reveal a smaller lattice shrinkage ratio compared to the center, indicating a gradual decrease in shrinkage rate from the interior to the exterior. The HAADF-STEM image (Fig. 4d) provides a vivid visualization of the curled ultrathin morphology of the PdPt NDs. This method of introducing strain not only enables precise control over the synthesis of nanosheets but also significantly enhances the catalytic performance of the material. This meaningful strategy offers valuable insights for the subsequent preparation of a more diverse range of nanosheets. Furthermore, we also achieved the successful synthesis of ultrathin Au nanosheets with a thickness of merely 4 nm, leveraging these surfactants along with the co-directing agent 2-mercaptonicotinic acid. The synthesis pathway is outlined in Fig. 4e. The addition of 1,2-MNA as a sulfate ligand serves multiple purposes, one being its synergistic effect with the surfactant C22N-COOH (Cl−) to smoothly generate layered micelles and consequently produce ultrathin Au nanosheets. Analysis using UV-visible absorption spectroscopy and XRD (Fig. 4f) revealed the strong localized surface plasmon resonance (LSPR) of the gray-blue Au NSs. Fig. 4g illustrates the growth process and final morphology of the Au NSs, clearly demonstrating their exceptional branching degree and ultrathin nature. Moreover, we extended this approach to synthesize AuAg, AuCoAg, and other nanosheets, overcoming the limitations of gold nanomaterials in terms of thickness and branching degree, thus offering expanded options for material science applications.69,72
Fig. 5 (a) Illustration of the mechanism for the controllable assembly of Pd NSs. (b) TEM images and XRD patterns of disassembled and assembled Pd NSs. (c) Schematic and TEM image of oxidative etching in the cases of disassembled and assembled Pd NSs. Reproduced with permission from ref. 66. Copyright 2017, the Royal Society of Chemistry. (d) TEM, SAED and HRTEM images of the Pd NSs (HRTEM images showing the presence of defects, twins, stacking faults and amorphous regions). Reproduced with permission from ref. 64. Copyright 2018, Elsevier. (e) TEM and HAADF scanning-TEM images of the platinum nanowheels. Reproduced with permission from ref. 65. Copyright 2011, the Owner Societies. (f) Morphology and structure of an ultrathin Au membrane. (g) Microstructure analysis of the HGM hydrogel. Reproduced with permission from ref. 74. Copyright 2018, Macmillan Publishers Limited. |
Furthermore, Gao et al. employed a straightforward one-pot method to synthesize PdPtNi nanosheets in a mixed solution containing PVP and CTAC. The resulting product is characterized by its ultra-thin and porous nature, boasting numerous active sites on its surface, enhanced conductivity, and superior catalytic kinetics.75 Niu et al. fabricated a gold film that stands out as the thinnest single-crystal gold sheet reported at the micrometer scale (Fig. 5f). Their approach involved the use of a surfactant system composed of non-ionic HGM and co-surfactant SDS. As illustrated in Fig. 5g, the assembled nanosheets exhibit remarkable thinness, reaching a mere 3.6 nanometers.106 PVP, serving as a potential crystal surface inhibitor, exhibits the ability to selectively adsorb onto specific planes, thereby promoting the anisotropic growth of nanosheets, regulating nucleation kinetics, and ultimately exerting control over the formation of nanosheets. Our team has employed oleamine and glucose as structural directing agents to successfully synthesize two-dimensional ultrathin PdPtMoCrCoNi nanosheets (NSs) with a uniform structural morphology.107 To further delve into the significance of surfactants in the synthesis process of nanosheets, we carried out a comparative study, exploring both surfactant-assisted and surfactant-free synthesis methods. TEM images revealed that in the presence of surfactants, C22-Py (Br−) spontaneously assembles into a layered structure, serving as a confinement agent. This confined environment facilitates the reduction of metal precursors within a limited space, leading to the formation of ultrathin, approximately square nanosheets. Conversely, without the addition of surfactants, the nanosheets exhibit a hexagonal morphology, resulting from the adsorption and reduction of carbon monoxide (CO).71
Drawing upon CPP data and preceding research endeavors, it becomes evident that alkyl chains of elongated lengths contribute to diminished surface tension and heightened hydrophobic effects, which in turn yield elevated charge counts and a denser packing within micelles. This enhanced configuration fortifies the stability of nanosheets and impedes the development of alternative crystal facets through steric hindrance mechanisms.108,109 As a consequence, this orchestrated self-assembly process promotes an increase in the aspect ratio. The surfactant molecules adhere to the hydrophobic surface along their tails, with the adsorption strength augmenting proportionally to the chain length. Conversely, shorter hydrophobic chains prove inadequate in constructing effective micelles that can proficiently hinder growth, underscoring the significance of chain length in achieving optimal growth control.110,111 For instance, the Ryoo group successfully synthesized zeolite nanosheets using designed surfactant, C22-6-6 (2Br−). They investigated the impact of surfactant tail length on the synthesis process. When the long chain was shortened, the number of hydrophobic groups at the tail decreased, inhibiting the transformation of the single-layer structure into a multi-layer membrane structure.112 Luo et al. also examined the effects of surfactants with varying hydrophobic alkyl chain lengths (including DTAB, TTAB, STAB, CTAB, and others). The formation of nanosheets is influenced by a combination of spatial hindrance, electrostatic interactions, and other factors that arise from the differing lengths of the alkyl chains.103
Our group has delved into the unique characteristics of surfactants. In 2018, we successfully synthesized various Cn-PyB surfactants, each differing in carbon chain length, to fabricate two-dimensional ultrathin nanosheets. Under electrostatic forces, these surfactants, carrying specific charges, naturally align with negatively charged precursor salt ions, undergoing self-assembly into primitive inorganic–organic nanosheets. Subsequently, hydrophobic interactions driven by surfactants lead to the formation of refined nanosheets. Upon careful observation, it becomes evident that even under near-identical synthesis conditions, nanosheets assembled by pyridine-based surfactants with diverse carbon chain lengths exhibit distinct morphologies. As depicted in Fig. 6a, surfactants with shorter carbon chains tend to form large nanoparticles, failing to achieve the desired nanosheet structure. Conversely, only when surfactants with a C22 carbon chain length are employed, can ultrathin two-dimensional nanosheets be reliably obtained. This phenomenon suggests that for these surfactants, the longer the carbon chain, the more pronounced is the hydrophobic effect. This strengthened hydrophobicity is advantageous for limiting the self-assembly growth of metals and surfactants to follow the predetermined crystal plane and direction, thereby enhancing the formation of high-quality nanosheets.57
Fig. 6 (a) TEM images of Pd NSs synthesized using surfactants with different alkyl lengths. Binding behaviors of (b) Me-Py*, (c) QA*, and (d) HCOO-* onto different Pd crystal planes (100), (110), and (111). Reproduced with permission from ref. 57. Copyright 2018, the Royal Society of Chemistry. (e) TEM images of PD NSs prepared by using surfactants C22-Py C and C22-Py I. Reproduced with permission from ref. 58. Copyright 2017, the Royal Society of Chemistry. |
Therefore, regardless of the type of hydrophilic head or the length of the hydrophobic tail, the morphology of nanosheets is regulated by affecting adsorption behavior and crystal growth dynamics. A hydrophilic head with strong polarity and a larger volume, along with a long hydrophobic chain, can reduce surface energy, realize the growth of specific crystal faces, and ultimately form nanosheets with regular shapes. This precise control of nanosheet growth can be achieved by optimizing the surfactant structure.
Fig. 7 (a) Schematic illustration for the formation of PtP NDs via a post-phosphating process. (b) HRTEM image (inset: the corresponding FFT pattern), the schematic diagram of three types of P atom doping, HAADF-STEM image and elemental mapping images of PtP NDs. Reproduced with permission from ref. 71. Copyright 2022, Wiley-VCH GmbH. (c) Schematic illustration for the growth mechanism, (d) HAADF-STEM (inset: the structure of MoS3) and corresponding elemental mapping of Pt-α-MoS3 NDs. Reproduced with permission from ref. 70. Copyright 2023, Wiley-VCH GmbH. (e) HAADF-STEM images and elemental mapping images (inset: the corresponding Fourier diffractogram), and (f) XRD pattern, PdP (100) and schematic diagram of PdP (100). Reproduced with permission from ref. 60. Copyright 2022, American Chemical Society. |
Nanostructures | Reaction | Electrolyte | Performance |
---|---|---|---|
(100)-Pd NSs58 | GOR | 1 M GI + 1 M KOH | 76.7 mA cm−2 |
(100)-Pd NSs | HER | 0.5 M H2SO4 | 67 mV at 10 mA cm−2 |
(110)-Pd NSs57 | 158 mV at 10 mA cm−2 | ||
(111)-Pd NSs | 227 mV at 10 mA cm−2 | ||
MOR | 1 M MeOH + 1 M KOH | 2.06 A mgPd−1 | |
PdIr NSs61 | EOR | 1 M EtOH + 1 M KOH | 3.70 A mgPd−1 |
GOR | 1 M GI + 1 M KOH | 1.36 A mgPd−1 | |
PdPtCu NSs63 | MOR | 1 M MeOH + 1 M KOH | 2.67 A mgNM−1 |
PtRu NDs68 | MOR | 1 M MeOH + 1 M KOH | 3.05 A mgPt−1 |
PdPt NDs69 | HER | 0.5 M H2SO4 | 10.8 mV at 10 mA cm−2 |
PdP NDs93 | EOR | 1 M EtOH + 1 M KOH | 3.20 A mgPd−1 |
(100)-PdP NSs71 | EOR | 1 M EtOH + 1 M KOH | 4.07 A mgPd−1 |
PtP NDs70 | HER | 0.5 M H2SO4 | 13.3 mV at 10 mA cm−2 |
MOR | 1 M MeOH + 1 M KOH | 4.2 A mgPt−1 | |
Pt-α-MoS3 NDs60 | HER | 0.5 M H2SO4 | −11.5 mV at 10 mA cm−2 |
1 M KOH | −16.3 mV at 10 mA cm−2 | ||
PdPtP NDs62 | EOR | 1 M EtOH + 1 M KOH | 14.3 A mgPd+Pt−1 |
Fig. 8 (a) and (b) CV curves of PdNSs with different sizes and commercial Pd black for glycerol electrooxidation. (c) Chronoamperometry curve. Reproduced with permission from ref. 58. Copyright 2017, the Royal Society of Chemistry. Electrocatalytic HER performances of different PdNSs in 0.5 M H2SO4. (d) LSV curves of PdNSs, PdB and cPt. (e) LSV curves of PdNSs before and after repeating CV scans of 15000 cycles. (f) Time-dependent curves of PdNSs and commercial PdB under a constant overpotential of 42 mV. Reproduced with permission from ref. 57. Copyright 2018, the Royal Society of Chemistry. (g)–(i) CV, LSV curves and time-dependent curves of HER activity and stability of PtNDs in 0.5 M H2SO4 electrolyte. Reproduced with permission from ref. 67. Copyright 2019, American Chemical Society. |
Fig. 9 (a) CV curves and i–t in 1.0 M KOH and 1.0 M methanol, (b) CO stripping voltammograms obtained in 1.0 M KOH of the trimetallic PdPtCu nanosheets, bimetallic PdPt and PdCu nanosheets, monometallic Pd nanosheets, PdPtCu nanoparticles, and commercial Pt and Pd nanoparticles. Reproduced with permission from ref. 63. Copyright 2019, the Royal Society of Chemistry. (c) CV curves and summarized mass activities of Pt, Pt95Ru5, Pt90Ru10, Pt80Ru20, Pt75Ru25 NDs, and commercial PtC in 1.0 M KOH and 1.0 M methanol. Reproduced with permission from ref. 68. Copyright 2019, the Royal Society of Chemistry. (d) Normalized mass activity CV curves and normalized mass and specific activities of PdIr NDs, PdIr NDs-1, PdIr NDs-2 and PdIr NSs obtained in 1.0 M KOH and 1.0 M EtOH (e) for MeOH and (f) for GI. Reproduced with permission from ref. 61. Copyright 2022, Wiley-VCH GmbH. (g) DFT study of the HER on PtPd CNDs and PtPd NDs. (h) LSV curves at a scan rate of 5 mV s−1, and summarized j0 and η10 values of HER activities of PtPd CNDs, PtPd NDs, and commercial Pt/C in 0.5 M H2SO4. Reproduced with permission from ref. 69. Copyright 2023, American Chemical Society. |
Numerous studies have delved into the strain effects, leading us to successfully fabricate bent PdPt NDs endowed with long-range compressive strain. A density functional theory (DFT) analysis, depicted in Fig. 9g, reveals that this strain effectively diminishes the adsorption strength of catalytic intermediates, thereby accelerating the kinetics of catalytic reactions. Additionally, it fine-tunes the local coordination environment of active sites, theoretically leading to a substantial enhancement in catalytic performance. Remarkably, under standard testing conditions, the linear sweep voltammetry (LSV) curve of the PdPt NDs, measured in a 0.5 M sulfuric acid solution, exhibited a commendable overpotential of just 10.8 mV. This represents a significant improvement in both electrocatalytic activity and stability compared to unstrained and flat PdPt NDs. The surfactant-induced strategy for synthesizing ultrathin nanosheets with extensive strain holds immense potential for diverse applications, paving the way for the production of a wider array of high-performance electrocatalysts. In 2022, PdIr NSs with different structures and compositions were prepared using self-synthesized surfactants to investigate their synergistic advantages for electrocatalytic alcohol oxidation. We conducted extensive measurements across various alcohol types, including methanol, ethanol, and glycerol, to assess the universality of this approach. The results, presented in Fig. 9d, e, and f, are promising. Ultrathin PdIr NSs with a thickness of 2.1 nm exhibited the highest mass activity, and interestingly, an increase in branching degree led to a corresponding enhancement in alcohol oxidation reaction (AOR) performance. This underscores the intricate relationship between elemental composition, material morphology, and electrocatalytic performance. Therefore, the meticulous design of two-dimensional nanomaterials, harnessing the synergistic control of self-assembled surfactants and halides, offers a powerful tool for elevating electrocatalytic activity. This approach not only expands our understanding of nanoscale catalysis but also points to new inspiration in materials science and catalysis research.61,69
Fig. 10 (a) Schematic illustrating the electrocatalytic EOR of ultrathin (100)-PdP NSs. Reproduced with permission from ref. 71. Copyright 2022, American Chemical Society. (b) CVs curves for the PdP NSs, Pd NSs, and Pd/C collected in 1.0 M KOH and 1.0 M EtOH (inset: the corresponding mass activities). Reproduced with permission from ref. 93. Copyright 2020, the Royal Society of Chemistry. (c) CV curves and summarized mass/specific activities of PdPtP, PdPt, PdP, Pd NDs, and Pd/C collected in 1.0 M KOH and 1.0 M ethanol. Reproduced with permission from ref. 62. Copyright 2024, the Royal Society of Chemistry. (d) and (e) CV curves and summarized mass activities of PtP NDs with different P content collected in 1.0 m KOH and 1.0 m MeOH. (f) and (g) LSV curves and summarized η10 values at a scan rate of 5 mV s−1 of PtP NDs with different P contents. (h) Schematic illustration for atomic-level P-doped effect for boosting electrocatalytic performance of both the HER and MOR. Reproduced with permission from ref. 70. Copyright 2022, Wiley-VCH GmbH. (i) Summarized mass activity and (j) i–t chronoamperometry curves of (100)-PdP NSs with different amounts of doping P collected in 1.0 M KOH and 1.0 M EtOH. Reproduced with permission from ref. 71. Copyright 2022, American Chemical Society. (k) and (l) Steady stability test of the electrocatalysts of commercial Pt/C, Pt NDs, and Pt-α-MoS3 NDs for the HER (inset: chronoamperometry of Pt-α-MoS3 NDs at a fixed potential of η10), and LSV curves that collected before and after the stability test of Pt-α-MoS3 NDs. Reproduced with permission from ref. 60. Copyright 2023, Wiley-VCH GmbH. |
Encouraged by the above findings, we conducted ethanol catalytic oxidation tests on (100)-PdP NSs, which further validated the exceptional catalytic capabilities of this material.71,93 The mass activity and stability is optimized when the P content reaches 3.75%. Volcano plots were constructed based on the catalytic performance exhibited by different concentrations of phosphorus doping (Fig. 10i and j). This study was further extrapolated to synthesize other multicomponent alloy nanosheets doped with phosphorus, aiming to enhance their electrocatalytic performance.71 When compared with other catalysts (Fig. 10c), PdPtP NDs doped with phosphorus exhibit the lowest initial potential, as well as the highest mass and specific activity, during ethanol oxidation reactions. This underscores the pivotal role of phosphorus doping in alloy regulation.62 Due to the significant difference in electronegativity between Pt and P, a substantial amount of electron transfer occurs during the formation of nanosheets. This transfer not only facilitates the hydrogen evolution reaction (HER) but also significantly enhances the methanol oxidation reaction (MOR) performance. We delve into the underlying mechanism (Fig. 10h). Compared to Pt NDs without phosphorus doping, our PtP nanosheets exhibit superior reaction kinetics in both the HER and MOR. The introduction of phosphorus leads to electron transfer activity from Pt to P, resulting in optimized electronic states. During the HER catalytic process, the phosphorus sites are more adept at adsorbing H+ in acidic environments. Conversely, in the MOR process, positively charged Pt atoms readily combine with methanol molecules to form adsorbed carbon monoxide species (COads). Concurrently, both Pt and oxygenophilic P atoms facilitate the formation of adsorbed hydroxyl species (OHads), which subsequently react rapidly with adjacent COads. Therefore, the key to enhancing catalyst performance through phosphorus doping lies in the augmentation of reaction kinetics.
Based on the aforementioned insights, we conducted a dual functional analysis of PtP nanosheets. Fig. 10d and e demonstrate their catalytic oxidation activity towards methanol, revealing a regular trend in mass activity depending on the amount of phosphorus doping. Accordingly, a volcano plot was generated to visualize the impact of phosphorus atom content on catalytic performance. Similarly, the catalytic analysis of the HER (Fig. 10f and g) also yields volcano plots exhibiting distinct patterns. These tests collectively indicate that appropriate phosphorus doping can significantly enhance the catalytic performance of nanosheets; however, excessive doping leads to a decline in catalytic activity due to reduced catalyst conductivity and an increase in positively charged Pt sites, which hinder the removal of adsorbed hydrogen (Hads) and carbon monoxide (COads) species during the HER and MOR. Doping phosphorus into nanosheets represents an effective approach to enhance the electrocatalytic performance of catalysts. Analogously, anchoring MoS3 is another research objective of ours. The high dispersion of MoS3 significantly boosts the activity of Pt catalysts under both acidic and alkaline conditions. Notably, Pt-α-MoS3 nanosheets maintain excellent HER catalytic activity even after 24 hours of operation, with impressive η10 potentials (Fig. 10h–l). The stability of Pt-α-MoS3 nanosheets surpasses that of commercial Pt/C and nanosheets without α-MoS3 doping. Our work underscores the importance of atomic engineering of active sites. Although the electrocatalytic performance of precious metals and precious metal-based alloys doped with phosphorus has been significantly improved, numerous challenges remain. As a nonmetal, phosphorus exhibits limited conductivity, hence excessive doping can detrimentally affect the electrocatalytic performance. Moreover, P atoms are prone to oxidation during the electrocatalytic process, highlighting the need for discovering catalysts with more stable structures.60,70
Despite notable advancements in the synthesis of two-dimensional nanosheets, numerous challenges persist in developing straightforward methods for fabricating nanosheets with tunable morphologies and significantly enhanced catalytic performance. A notable hurdle lies in the purification step, where the complete removal of surfactants from the samples remains elusive. Even trace amounts of surfactant residue can obstruct active sites on the nanosheets, thereby impeding accurate characterization and testing of their various properties. Additionally, while laboratory-scale synthesis and evaluation of nanosheets have reached satisfactory levels, scalable production of high-quality 2D nanosheets and the precise engineering of their morphology, size, and composition remain crucial bottlenecks hindering large-scale commercialization and industrial utilization.125–127
As we gaze ahead into the future, the potential of nanosheet electrocatalysts and their practical applications appear vast and limitless, offering boundless opportunities:
(1) Firstly, researchers must consistently delve into surfactant molecules (including novel structure designs and functionalities) to enhance the precision of self-assembly. They must strive to develop surfactants that promote greater stability, repeatability, reliability, and ease of removal from metallic nanosheet catalysts. Additionally, the quest for efficient electrocatalysts capable of driving fundamental energy conversion processes remains paramount. The main aim is to synthesize metallic nanosheets with a cleaner surface by reducing surfactant adsorption on electrocatalysts.
(2) Secondly, among the numerous promising avenues, nanosheet catalysts offer the tantalizing possibility of being tailored and modulated to meet specific functional and application requirements. This customization extends to various aspects, including but not limited to element composition and microelectronic motion. Their adaptability in terms of chemical properties enables them to accommodate diverse energy conversion processes. Their remarkable structural attributes and performance have the potential to revolutionize numerous energy conversion and storage technologies. Consequently, this may pave the way for unprecedented and highly efficient clean energy conversion methods. Central to this endeavor are efforts to refine the electronic motion within nanosheets and fine-tune their active sites, ultimately enhancing the activity and stability of these catalysts.
(3) Thirdly, the commercialization, scalability, and seamless integration of two-dimensional nanosheets into functional hardware represent crucial areas of research. Given their exceptional performance and tunable properties, the future of these nanosheets is poised to intersect with numerous diverse fields. In particular, their application in electrocatalysis exhibits remarkable sensitivity, selectivity, and catalytic activity, making them suitable for the detection and degradation of pollutants, toxic gases, and harmful chemicals. Notably, fuel cells stand as the cornerstone of sustainable energy technology, and as such, the precise control of self-assembly processes and the accelerated translation of research outcomes into practical applications play a pivotal role in addressing urgent environmental challenges, mitigating climate change, and promoting energy recycling efforts.
(4) Fourthly, there is an ongoing necessity for focused research efforts to delve deeper into exploring synthesis routes that offer increased convenience, reproducibility, and high yields for high-performance catalysts. This pursuit must be complemented by a constant quest to minimize costs and expand production capabilities. Leveraging advancements in nanotechnology, researchers are now able to fine-tune the molecular-level characteristics of fuel cell components, thereby optimizing their performance and addressing longstanding challenges. The development of scalable production technologies and manufacturing processes will serve as the linchpin for the widespread commercial and industrial adoption of two-dimensional nanosheets.
In essence, ongoing research in these areas is anticipated to further deepen our comprehension of the synthesis and self-assembly mechanisms for two-dimensional nanosheets. This, in turn, will fully unlock their catalytic potential, laying the foundation for the creation of novel nanomaterials with superior performance and functionality that are scalable for mass production.
This journal is © The Royal Society of Chemistry 2024 |