Skip to content
Publicly Available Published by De Gruyter January 17, 2014

Nanomaterials engineering and applications in catalysis

  • Qiao Zhang EMAIL logo and Yadong Yin

Abstract

Heterogeneous catalysis utilizing metal particles plays an essential role in the industrial applications. Design and fabrication of highly active catalysts in an efficient and cost-effective way is thus an important topic. The emergence of nanotechnology provides an excellent opportunity for developing new catalysts. In this critical review, we present our efforts and perspective on the recent advances in engineering nanomaterials for catalysis, including synthesis, stabilization, and catalytic applications of nanoparticles. We first briefly summarize the advanced colloidal synthesis of metal nanoparticles using Ag nanoplates as the model system, and then discuss the strategies for stabilization of metal nanoparticles using both chemical and physical approaches. And finally, for practical applications, we have designed and synthesized a highly efficient, stable, and cost-effective TiO2-based photocatalyst by combining both non-metal doping and noble metal decoration.

Introduction

Catalysis, including heterogeneous, homogeneous, and enzymatic catalysis, is vitally important for the development of modern society. Improving the activity of catalysts to produce more molecules per unit area per unit time was the main focus in the 20th century, while achieving 100 % selectivity in all catalyst-based chemical processes has been the Holy Grail of catalysis science in the 21st century [1, 2]. Among the three types of catalysis, heterogeneous catalysis has some advantageous features, such as easy separation and long lifetime. This review article will focus on the topic of heterogeneous catalysis.

In a typical heterogeneous catalytic process, the catalytic reaction occurs repeatedly by a sequence of elementary steps, including molecular adsorption, surface diffusion, chemical rearrangement (bond-breaking, bond-forming, molecular rearrangement) of the adsorbed reaction intermediates, and desorption of the products [3]. The reaction can occur only when the reactants have been adsorbed onto the catalyst surface. As a result, the reaction rate of a catalytic process is highly dependent on the total surface area of the solid catalysts. And the local bonding geometries of the catalysts can play a deterministic effect on the selectivity of the specific catalytic reactions. The rise of nanotechnology over the past two decades has opened the door to a revolution in catalysis science [4]. The new technology can not only help improve the understanding of reaction mechanisms in industrial catalysis but also provide new catalysts with high activity as well as high selectivity. It has been widely accepted that a higher catalytic activity could be achieved by increasing the surface area of the specific active phase of the catalyst through the reduction of the size of the corresponding catalytic particles. The fraction of atoms on particle surfaces increases dramatically as the size of a nanoparticle decreases, providing more active sites for catalytic reactions. Additionally, the catalytic selectivity could be improved by using nanocrystals with a specific shape that have specific desirable local bonding geometries [5].

To date, there are still some challenges remaining. For example, although numerous synthetic mechanisms or hypotheses have been proposed to explain the experimental phenomena in the synthesis of various shaped nanostructures, there is no general understanding that can properly explain all systems. The lack of deep understanding of the synthetic systems also makes it difficult to reproduce many of the reported results with satisfactory yield and quality because some minor unintentional alterations to the reaction conditions may easily disturb nanostructure formation [6, 7]. It is thus highly desirable to discover and clarify the underlying mechanism. In addition, the low stability of metal nanoparticles has been a big challenge for their practical applications. Metal nanoparticles tend to agglomerate and rapidly sinter into larger particles during the catalytic process, mainly due to their large surface area and the high surface energy. This sintering process leads to the reduction of the active sites, and also to the loss of the unique properties of the nanostructured catalysts. One dramatic example of the importance of this issue can be seen with Au catalysts: Au nanoparticles dispersed on high-surface-area supports have been shown to be quite active in many low-temperature reactions, including carbon monoxide and hydrocarbon oxidations [8], but also to lose their activity over time as they sinter into larger particles.

In this review article, we summarize our recent efforts and present some thoughts on the study of metal nanoparticles for heterogeneous catalysis. First, we introduce some progress in the synthesis of anisotropic nanostructures by using Ag nanoplates, a highly anisotropic structure, as the example. Through a comprehensive study, we revealed the role of each reagent and found that H2O2 played an important role in the synthesis of Ag nanoplates. And this significantly refined understanding allows us to develop an efficient and highly reproducible process for making Ag nanoplates with controllable edge length and thickness as well as desired surface plasmon resonance bands. Second, we address the issue of enhancing the stability of metal nanoparticles through both the chemical and physical approaches. In the chemical approach, a uniform Au layer was deposited onto the Ag nanoplates to enhance its stability against various etchants. In the physical approach, the metal nanoparticle was encapsulated with a porous oxide shell. In the metal-core/oxide-shell structure, the existence of oxide shell can help prevent the physical contact of metal nanoparticles and thus eliminate the possibility of sintering. Since the outer oxide shell is mesoporous, the diffusion of both reagents and products can therefore be achieved. Then, we discuss the unique features afforded by core-shell nanostructures in the creation of desired interfaces between different components. The resulting synergetic interactions may lead to superior catalytic performance. Finally, we conclude with a short summary and our personal perspectives on the directions in which future work in this field might be focused.

Colloidal synthesis of silver nanoplates

Ag nanoplates, also named as nanoprisms or nanodisks, are 2D plasmonic nanostructures that have attracted intensive attention owing to their exceptional plasmonic properties as well as the related applications [914]. Since the seminal report by Mirkin et al. on the photo-induced synthesis of Ag nanoplates in 2001 [15], numerous colloidal strategies have been reported, including photochemical processes [1621], electrochemical synthesis [22], ligand-assisted chemical reductions [2325], templating procedures [26], and sonochemical routes [27]. Due to their isotropic face-centered-cubic (fcc) crystal structure, the formation of Wulff polyhedrons (a truncated octahedron) enclosed by a mix of {100} and {111} facets is thermodynamically more favored than the formation of extremely anisotropic Ag nanoplates [2830]. To explain the formation of such highly anisotropic nanostructures, a variety of mechanisms or hypotheses have been proposed. In the early stage of research, the “face-blocking theory” that attributes the formation of anisotropic nanostructures to capping effect is the most popular theory. It is believed that the growth rate of certain facets could be selectively slowed down by introducing a capping agent that can selectively adhere to the particular crystal facet of the growing nanocrystal [9, 10]. Later, it is realized that the crystal symmetry of the starting nuclei also plays an important effect in the formation of anisotropic structure. For materials with isotropic crystal structures, anisotropic growth is usually facilitated by breaking the isotropic symmetry with the formation of twin or other defect planes during the nucleation stage, which is, however, still a great challenge to control during synthesis. Based on crystallographic arguments, it is believed that the final morphology of the nanostructure is determined by the internal crystal structure of the original seed particle because of the limited number and variety of crystal facets available for growth [3133]. Xia et al. have recently pointed out the importance of interweaving both crystallographic and surface chemistry arguments in the elucidation of the overall mechanism of nanoplate formation [12, 34].

Formation of Ag nanoplates promoted by H2O2

Through a systematic study on a typical chemical reduction reaction, we confirmed that the formation of Ag nanoplates could be promoted by the addition of a mild oxidant, H2O2. In the direct chemical reduction route, Ag nanoplates are obtained by reducing an aqueous solution of AgNO3 with NaBH4 in the presence of trisodium citrate (TSC), polyvinyl pyrrolidone (PVP), and H2O2 [35]. In the absence of H2O2, when PVP or citrate was used as the surfactant, only quasi-spherical Ag nanoparticles could be obtained, as evidenced by the characteristic absorption peak around 400 nm (Fig. 1a) as well as the TEM characterization (Fig. 1b). When H2O2 is introduced into the system, some anisotropic structures formed, suggested by the appearance of a peak in the longer wavelength range. As shown in Fig. 1a, a sharp peak around 450 nm appeared along with a shoulder at ~400 nm when the concentration of H2O2 is 5 mM. The sharp characteristic peak around 400 nm disappeared as the concentration of H2O2 was increased to 10 mM, implying the disappearance of spherical Ag nanoparticles. The weak shoulder around 380–420 nm could be attributed to the in-plane quadrupole resonance of Ag nanoplates. When the concentration of H2O2 was increased to 20 mM, the quadrupole resonance became more pronounced and the dipole resonance red-shifted to ~600 nm. The product is triangular Ag nanoplates with almost 100 % yield, as shown in Fig. 1d. The thickness of the Ag nanoplates could be varied from ~3 to ~6 nm by varying the concentration of sodium borohydride. The reaction process has been monitored using a UV/vis spectrometer to investigate the detailed formation process of the Ag nanoplates. As shown in Fig. 1c, only the characteristic peak of spherical Ag nanoparticles (~400 nm) could be observed in the early stage. The peak intensity increased quickly and reached the maximum within 5 s after the beginning of the formation of nanoparticles. The peak intensity at 400 nm then gradually dropped, indicating the consumption of spherical particles, while another peak at ~500 nm emerged and red-shifted to longer wavelengths, suggesting the formation and growth of anisotropic structures.

Fig. 1 
            (a) UV/vis spectra of Ag nanoparticles obtained by controlling the concentrations of H2O2; (b) TEM image showing the morphology of products prepared with only PVP as the surfactant and in the absence of H2O2; (c) in situ measurement of the formation of Ag nanoplates with a time interval of 7.5 s. The inset in (c) shows the UV/vis spectra of the initiation stage during the synthesis of Ag nanoplates with a time interval of 0.75 s; (d) TEM image showing the typical morphology of Ag nanoplates in the presence of H2O2. The inset in (d) shows a TEM image in which Ag nanoplates stand vertically upon their edges. Adapted with permission from ref. [7].
Fig. 1

(a) UV/vis spectra of Ag nanoparticles obtained by controlling the concentrations of H2O2; (b) TEM image showing the morphology of products prepared with only PVP as the surfactant and in the absence of H2O2; (c) in situ measurement of the formation of Ag nanoplates with a time interval of 7.5 s. The inset in (c) shows the UV/vis spectra of the initiation stage during the synthesis of Ag nanoplates with a time interval of 0.75 s; (d) TEM image showing the typical morphology of Ag nanoplates in the presence of H2O2. The inset in (d) shows a TEM image in which Ag nanoplates stand vertically upon their edges. Adapted with permission from ref. [7].

Based on the careful and systematic study, a plausible mechanism for the synthesis of Ag nanoplates has been proposed: Ag nanoparticles were first formed through the reduction of Ag+ ions by NaBH4. Thanks to the Ag-citrate coordinating interaction and the presence of the powerful etchant, H2O2, the newly formed Ag nanoparticles contain many defects, including the twinned defects that favor the planar growth into plate shapes. In the previous paper, we have suspected that H2O2 can remove the relatively unstable nanoparticles at this stage, leaving only the most stable ones [7]. Recently, our further research showed that Ag nanoparticles can trigger the decomposition of H2O2 to release active oxygen, which favors the formation of twinned defects.

This improved understanding not only provides us with a highly reproducible synthetic tool but also allows us to study the role of each reagent. For example, contrary to previous reports in which citrate has been considered to be a “magic” component critically required for the formation of Ag nanoplates, our investigation clearly points out, for the first time, that citrate is not an irreplaceable component: it is required to stabilize the formed Ag nanoplate nuclei through preferential binding to the (111) facets but can be completely substituted by many other carboxyl compounds containing two carboxylate groups. We found that PVP is not required for obtaining high-quality Ag nanoplates and the exclusion of PVP could improve the reproducibility of the synthesis. In our further research, H2O2 has been successfully used to improve the yield of traditional seed-mediated growth process, further confirming the key role of H2O2 in the synthesis of Ag nanoplates [36].

Shape conversion from metallic Ag particles to Ag nanoplates

The standard potential in the peroxide-water couple is dependent on the pH value of the solution [37, 38]. In acidic solutions

H2O2+2H++2e→2H2O (E°=1.763 V)

And in alkaline solutions

H2O2+2e→2OH(E°=0.867 V)

Since the potentials under both acidic solution (E°=1.763 V) and alkaline solution (E°=0.867 V) are higher than that of Ag+/Ag (E°=0.7996 V), H2O2 can be used as an etchant to dissolve metallic Ag and produce Ag+ ions. As a result, it can be used to convert large metallic Ag materials to Ag nanoplates, which has not been reported in the literature [24]. Figure 2 shows two examples of the conversion from pre-formed Ag nanowires and nanoparticles to Ag nanoplates, respectively. When H2O2 was added to the pre-formed colloidal nanostructures, the solution gradually became colorless with the appearance of small bubbles due to the decomposition of H2O2. When NaBH4 was then added to reduce Ag+ back to Ag0, the solution gradually became yellow, red, and blue, indicating the formation of Ag nanoplates. The successful conversion has been confirmed by both UV/vis spectra and TEM images (Fig. 2). From the point of view of synthesis, the use of metallic Ag as the source ensures a higher degree of reproducibility as there is less possibility of introducing disturbance such as the anions of the Ag salt to the reaction.

Fig. 2 
            UV/vis spectra and TEM images showing both (a–c) Ag nanowires with lengths of up to 10 μm and (d–f) Ag nanoparticles with irregular shapes can be converted to Ag nanoplates in the presence of H2O2. Adapted with permission from ref. [7].
Fig. 2

UV/vis spectra and TEM images showing both (a–c) Ag nanowires with lengths of up to 10 μm and (d–f) Ag nanoparticles with irregular shapes can be converted to Ag nanoplates in the presence of H2O2. Adapted with permission from ref. [7].

Recently, Ekgasit and co-workers have found that H2O2 can not only act as an oxidant but also a reducing agent in the conversion of Ag nanoplates [39]. Under alkaline condition

H2O2+2Ag++2HO→2Ag+2H2O+O2 (E°=0.947 V)

As a result, starch-stabilized Ag nanospheres could be converted to Ag nanoplates by H2O2 without the addition of any other reducing agent. This result, again, confirmed the critical role of H2O2 in the synthesis of Ag nanoplates.

Precise tuning of Ag nanoplates within a wide range

Since the physiochemical properties of nanostructures are highly dependent on their morphologies, it is therefore important to find a way to precisely control the shape and size of nanostructures. We recently reported a two-step procedure for the controlled growth of Ag nanoplates with extremely high aspect ratios and consequently widely tunable plasmonic bands [40]. With the advantages of precise control and high reproducibility, this two-step procedure can conveniently produce nanoplates with high aspect ratios that have not been realized previously by using one-step reactions.

Ag nanoplates with size around 20–30 nm were first synthesized based on the previous mentioned method [7, 35], in which PVP and citrate were used as the surfactants to ensure the high-yield production of plate-like seeds. It is believed that PVP and citrate can selectively bind to the (100) and (111) facets of Ag nanoplates, respectively. The as-prepared small Ag nanoplates were then washed to remove PVP to expose the side facets and served as seeds for further growth. When excess citrate was added, the growth along the vertical axis was blocked, allowing only extensive growth along the lateral axis of the plates (Fig. 3a). Generally, both self-nucleation process and seed-mediated epitaxial growth can occur in the seeded-growth process. On the basis of the Gibbs–Thomson equation, a slow reaction rate as well as a low concentration of monomer is normally favorable for the seed-mediated epitaxial growth, while a fast reaction rate is favorable for the self-nucleation process [4144]. To eliminate the self-nucleation process, the reaction rate was slowed down by using a mild reducing agent, ascorbic acid, and pre-treating Ag ions with citrate ions to form Ag-citrate complexes [45]. The as-prepared complex can be reduced by ascorbic acid and release more citrate ions, which can help maintain anisotropic growth without introducing any new ligand.

Fig. 3 
            (a) Schematic illustration of the anisotropic seeded growth of Ag nanoplates based on selective ligand adhesion; TEM image (b) and SEM images (c–g) showing the evolution of Ag nanoplates during the stepwise growth process: from original seeds (b) after removal of PVP by centrifuging and washing to larger Ag nanoplates after (c) one, (d) two, (e) four, (f) five, and (g) seven cycles of seeded growth. Adapted with permission from ref. [40].
Fig. 3

(a) Schematic illustration of the anisotropic seeded growth of Ag nanoplates based on selective ligand adhesion; TEM image (b) and SEM images (c–g) showing the evolution of Ag nanoplates during the stepwise growth process: from original seeds (b) after removal of PVP by centrifuging and washing to larger Ag nanoplates after (c) one, (d) two, (e) four, (f) five, and (g) seven cycles of seeded growth. Adapted with permission from ref. [40].

As shown in Fig. 3b, the original Ag seeds were triangular nanoplates with a mean size around 25 nm. The edge length of the Ag nanoplate can gradually increase to ~130 nm after one cycle of growth (Fig. 3c). The growth cycle could be repeated many times and the edge length of Ag nanoplates could be readily increased to ~4 μm (Fig. 3d–g). An interesting “size distribution focusing” effect has been observed during the process. The initial Ag seeds have a broad size distribution, which become highly uniform after extended growth. The size distribution focusing effect could be attributed to the fact that the smaller nanoparticles grow more rapidly than the larger ones as driven by the higher surface energy. In addition, larger particles have a slower growth rate even when the deposition rates of Ag atoms are the same due to the greater ratio of volume versus size for large particles, meaning more Ag is needed to produce a net edge length increase. Thanks to the effective blocking effect of citrate ions and the formation of Ag-citrate complexes, only small change in the thickness of Ag nanoplates has been observed. As a result, Ag nanoplates with extremely high aspect ratio (edge length/thickness >400) have been obtained. And the plasmon band has been readily tuned from the visible to the infrared (IR) spectrum by controlling the aspect ratio of the nanoplates. Due to the fact that blood and tissue are most transparent in the region of near-IR (NIR) range, the ability to cover the entire near-IR (NIR) spectrum makes the as-prepared Ag nanoplates ideal candidates for biomedical applications such as photothermal cancer therapy [4649].

Since the complexation can significantly suppress the self-nucleation process, this idea has been further extended to the Au nanoparticle system. A one-step seeded growth has been developed by the Yin group to obtain Au nanoparticles with widely tuned size [43]. To effectively suppress the self-nucleation process, a growth solution was first prepared by mixing HAuCl4, KI solution, PVP, and ascorbic acid solution. In this solution, HAuCl4 is the Au source and ascorbic acid is the reducing agent. KI was used as a strong coordinating ligand to Au ions to decrease the reduction potential of Au salt, eliminating the burst self-nucleation process [42, 50]. At the same time, PVP has been employed as a bi-surfactant to stabilize the atomic Au monomers and further avoid the self-nucleation process [51]. PVP can also serve as the capping agent to prevent inter-particle agglomeration. The as-prepared growth solution can stay stable even at considerably high concentration. Upon injecting small Au nanoparticles as the seeds, the growth process occurs immediately. Through this seeded growth method, monodispersed Au nanoparticles with tunable size from ~10 to ~200 nm could be easily obtained. This new approach shines some light on the effective suppression of self-nucleation regardless of the seed/Au precursor ratio, which might be able to inspire some new ideas in other systems.

Monitoring the shape evolution of Ag nanoplates using a Au marker

One interesting phenomenon we observed in the seeded-growth process is the shape change of Ag nanoplates. As shown in Fig. 3, at the earlier stage of seeded growth, the corners of Ag nanoplates are extensively rounded so that the plates appear circular in SEM images. As the size increases, sharp corners are developed, then gradually truncated, and finally producing a mixture of triangular and hexagonal nanoplates. The similar shape transition phenomena have also been observed recently [52, 53]. Xia et al. have attributed it to the possibility that a truncated structure might be more thermodynamically stable, but that the growth was kinetically controlled and thus ultimately the plates returned to their original triangular shapes. However, no solid evidence has been provided. Although the crystal structure of the 2D Ag nanoplates and Au nanoplates seems to be simple, many different model crystal structures have been proposed. For example, the 1/3{422} reflections have been observed in the electron diffraction (ED) pattern of Ag nanoplates and Au nanoplates. However, for a perfect fcc structure, the 1/3{422} reflection should be forbidden. Pileni and co-workers have attributed it to parallel stacking faults in the <111> direction [54] and further pointed out that these stacking faults should be responsible for the formation of plate structures by providing low-energy reentrant grooves which energetically favor lateral crystal growth. Later, Rocha and Zanchet conducted a careful study by using both high-resolution transmission electron microscopy (HRTEM) and X-ray diffraction (XRD). They pointed out that the internal structure of Ag nanoplates is indeed very complex, containing many twins and stacking faults [55]. They found that local hcp regions, rather than fcc structure, could form due to the existence of the planar defects in the <111> direction. The formation of hcp regions could explain the existence of forbidden 2.50 Å fringes that are observed in <111> orientated nanoplates. The formation mechanism of the hcp layer has been further elucidated by Kelly et al., and it has been used to explain the evolution of Ag nanoplates [56]. It is ascertained that Ag nanoplates are comprised of two fcc regions of different thicknesses, sandwiching an hcp layer originating from a series of internal stacking faults. In their devised crystal structure model, each nanoplate edge consists of a {111} facet and a {100} facet of different sizes. As a result, there are six edges in a truncated triangular nanoplates, half featuring a larger {111} facet and the other half with a dominant {100} facet. The formation of a triangular structure bonded by {111} dominated edges could be explained by the preferential deposition on the less thermodynamically stable {100} dominated edges.

To further elucidate the seeded-growth process and identify the crystal structure of Ag nanoplates, we designed a “marker” experiment to monitor the shape change of Ag nanoplates during the process [57]. In the experiment, Ag nanoplates were first synthesized based on the direct chemical reduction method [7], in which PVP was eliminated to avoid any disturbance. A galvanic replacement process was then applied to the Ag nanoplates by using Au cations [58]. As shown in Fig. 4b, the as-prepared seeds are nanoframes. Because the lattice mismatch between Au and Ag is negligible while the difference in atomic number is so significant, Au can be a good candidate for both seeding and serving as the marker to outline the boundaries of the Ag nanoplates.

Fig. 4 
            (a) Schematic illustration of the structural change of Ag nanoplates during seeded growth. Preferential Ag deposition on the {100} dominated edges of the seeds leads to a rounded or hexagonal intermediate with a mix of {100} and {111} dominated edges, after which further Ag deposition produces a larger triangular plate with {111} dominated edges; (b–g) step-by-step growth of Ag nanoplates using Au nanoframes as seeds. (b) Au nanoframe seeds and frame/plates after the addition of (c) 0.3 μmol, (d) 3 μmol, (e) 4.1 μmol, (f) 7.5 μmol, and (g) 10.75 μmol of Ag nitrate. Adapted with permission from ref. [57].
Fig. 4

(a) Schematic illustration of the structural change of Ag nanoplates during seeded growth. Preferential Ag deposition on the {100} dominated edges of the seeds leads to a rounded or hexagonal intermediate with a mix of {100} and {111} dominated edges, after which further Ag deposition produces a larger triangular plate with {111} dominated edges; (b–g) step-by-step growth of Ag nanoplates using Au nanoframes as seeds. (b) Au nanoframe seeds and frame/plates after the addition of (c) 0.3 μmol, (d) 3 μmol, (e) 4.1 μmol, (f) 7.5 μmol, and (g) 10.75 μmol of Ag nitrate. Adapted with permission from ref. [57].

As shown in Fig. 4c, the nanoframes can serve as seeds for the seeded growth process. A backfilling process was first observed when the seeded growth process started, followed by the normal growth into large nanoplates (Fig. 4d–g). Mirkin et al. have pointed out that the backfilling process could be attributed to the high-energy facets in the nanoframe interior [58]. The outward growth happened in the same manner as Ag nanoplates: It has a similar shape transition and ultimately returns to a triangular morphology (Fig. 4d–g). One interesting phenomenon is that there is a preference for deposition on the edges of the nanoframe/nanoplate with little to none occurring at the corners throughout the various stages of outward growth, causing the triangular to circular/hexagonal transition. The shape returns to triangular when more Ag is deposited, resulting in the formation of new corners at the deposition sites (Fig. 4f). The majority of the obtained product contains an embedded frame pointed 180° relative to the original orientation of the Ag nanoplate (Fig. 4f). No additional shape change could be observed even when more Ag atoms were deposited to make larger nanoplates (Fig. 4g). It can be concluded from the observation that the exclusive deposition of Ag atoms on the nanoplate edges, rather than the reaction kinetics, induced the shape transition.

This transition process could be well explained by using Kelly et al.’s crystal structure model, as illustrated in Fig. 4a. The edges of the initial Ag nanoplates are dominated by {100} facets, formed during the rapid reduction process. Because a weaker reducing agent, ascorbic acid, is used for the seeded growth process and the concentration of Ag ions is deliberately low, the thermodynamic consideration becomes more important, leading to selective deposition on {100} dominated sides and causing the observed shape transition to a structure with {111} dominated edges.

Stabilization of metal nanoparticles

The poor chemical and structural stability of nanomaterials has seriously limited their catalytic applications. For example, although Ag nanostructures have attracted much attention due to their exceptional structure-related plasmonic property, the practical applications of such structures have been rarely achieved. It has been reported that the structures and plasmonic properties of Ag nanoparticles are subject to change upon exposing to water [59], acids [60], halides [61], oxidative agents [62], UV irradiation [63], and heat [64]. Additionally, due to the large surface area and high surface energy, nanoparticles tend to rapidly grow larger without surface passivation, thus reducing the active surface area for catalysis. This process, often referred to catalysis as sintering, occurs with catalysts both dispersed in solution and immobilized on solid supports [65].

To enhance the stability of nanoparticles, many methods have been developed, which can be roughly divided into two main categories: the chemical and physical approaches [66]. The chemical approaches rely on substrate effects or on the formation of alloy or hybrid materials [67]. For example, the strong metal-support interaction (SMSI) has been realized for a long time. It has been found that remarkable thermal stability of metal nanoparticles could be achieved when deposited onto certain supports. Inorganic materials with large surface area, such as charcoal, salt, and oxides including SiO2, Al2O3, and TiO2, are the most widely used support materials in catalysis [6872]. Recently, multiple-oxide supports have been utilized to obtain stable metallic nanocatalysts, by exploring synergistic effects among the different components [73]. For instance, by coating a thin alumina layer on anatase TiO2 and subsequently depositing Au on the Al2O3/TiO2 dual oxide, Dai and co-workers have successfully produced a highly stable catalyst which retains high activity for CO oxidation even after calcination at 773 K [74, 75]. Additionally, forming alloy or hybrid materials can also significantly enhance the stability of metal nanoparticles. One notable example is the enhanced thermal stability of PtSn alloys in the propane dehydrogenation reaction, which operates at high temperature (~550–650°C) [76, 77]. In another case, to improve the stability and durability of Pt nanocatalysts in proton-exchange-membrane fuel celles (PEMFCs) [7882], some transition metals such as Ni and Co have been alloyed with Pt in order to suppress the growth of Pt nanoparticles [8388].

The physical approach is relatively more straightforward: a physical barrier is used to protect nanoparticles from contacting each other and thus prevent coalescence. Colloidal nanoparticles dispersed in solution are usually stabilized by electrostatic forces or steric stabilizers such as surfactants, ligands, or dendrimers [8999]. But those inter-particle forces or ligand binding can be easily perturbed during catalytic processes, thus leading to coagulation and the loss of the high activity associated with the state of colloidal dispersion [100]. Due to their poor thermal stability, although organic polymers have been used to stabilize nanoparticles in solution, they cannot be used when a catalytic reaction is performed at elevated temperatures or in highly oxidizing environments [101]. Recently, the core/shell structure, in which metal nanoparticles are encapsulated within a more stable shell, has attracted much attention due to the enhanced stability [102109].

Stabilizing Ag nanoplates through the Ag@Au core/shell nanostructure

As we mentioned above, the poor chemical and structural stability of Ag nanostructure has been the major concern in their practical applications. The Yin group recently developed a Au coating strategy that can not only enhance the stability of Ag nanoplates but also keep their excellent plasmonic property [50]. It is well known that galvanic replacement would happen when Ag nanoparticles are exposed to Au salt due to the different oxidative potentials. As a result, the key to uniformly depositing a thin layer of Au on the Ag nanoplates is to minimize the galvanic replacement process.

Similar to the above-mentioned one-step seeded growth of Au nanoparticles, an aqueous solution containing HAuCl4, KI, and PVP was first prepared as a growth solution, which was then slowly added into a mixture containing PVP, diethylamine, ascorbic acid, and Ag nanoplates. In the growth solution, thanks to the complexation between Au ions and I ions, the reduction potential of Au salt has been significantly decreased to +0.56 V (vs. standard hydrogen electrode, SHE), which is much lower than those for Au ions (E°(Au3+)=+1.52, E°(AuCl4)=+0.93 V, vs. SHE). The addition of diethylamine can ensure homogeneous Au deposition on the surface of Ag nanoplates, due to the unselectively binding of diethylamine on the surface of Ag nanoplates. Figure 5a shows a typical TEM image of the obtained Ag@Au core/shell nanoplates. Although some nanoplates contain small voids on the surface that are caused by galvanic replacement, the original triangular/hexagonal shapes are well maintained, suggesting that the galvanic replacement has been significantly suppressed. Notably, the plasmonic property has been well maintained, as shown in Fig. 5b. The intensity of the plasmon band of the final Ag@Au core/shell nanoplates is very close to that of the original Ag nanoplates. A slight blue-shift of the plasmonic peak has been observed, which can be attributed to the decrease of the aspect ratio. The core/shell structure has been further confirmed by the high-resolution energy-dispersive X-ray spectroscopy (EDS). Figure 5c shows a typical EDS line scan across the surface of a single nanoplate. Both Ag and Au have been detected in the bulk area of the core/shell nanoplates. About 0.5 nm from the edge of the nanoplates, the atomic percentage of Ag drops to zero and that of Au rises to 100 %, which is consistent with the fact that the edge Au atoms are deposited through reduction of Au cations instead of partial replacement of Ag.

Fig. 5 
            (a) TEM image of the Ag@Au core/shell nanoplates; (b) UV/vis/NIR spectra of the original Ag nanoplates and Ag@Au core/shell nanostructures prepared by adding various amounts of Au growth solution; (c) atomic percentage profile of a Ag@Au nanoplate calculated from the relative counts in the EDS line scan indicated by the arrow in the STEM image below; (d–i) stability of the Ag@Au nanoplates in (d) phosphate buffer solution (10 mm, pH 7.4, PVP 0.5 %), (e) NaCl solution (20 mm, PVP 0.5 %), (f) PBS (10 mm, NaCl 150 mm, pH 7.4, PVP 0.5 %), and (g) H2O2 solution (2.1 %), monitored by UV/vis/NIR spectrophotometry. A TEM image of the Ag@Au nanoplates after treatment in H2O2 for 2 h is shown in the inset to (g). Stability of (h) pristine Ag nanoplates and (i) MHA-stabilized Ag nanoplates in H2O2 (2.1 %). Adapted with permission from ref. [50].
Fig. 5

(a) TEM image of the Ag@Au core/shell nanoplates; (b) UV/vis/NIR spectra of the original Ag nanoplates and Ag@Au core/shell nanostructures prepared by adding various amounts of Au growth solution; (c) atomic percentage profile of a Ag@Au nanoplate calculated from the relative counts in the EDS line scan indicated by the arrow in the STEM image below; (d–i) stability of the Ag@Au nanoplates in (d) phosphate buffer solution (10 mm, pH 7.4, PVP 0.5 %), (e) NaCl solution (20 mm, PVP 0.5 %), (f) PBS (10 mm, NaCl 150 mm, pH 7.4, PVP 0.5 %), and (g) H2O2 solution (2.1 %), monitored by UV/vis/NIR spectrophotometry. A TEM image of the Ag@Au nanoplates after treatment in H2O2 for 2 h is shown in the inset to (g). Stability of (h) pristine Ag nanoplates and (i) MHA-stabilized Ag nanoplates in H2O2 (2.1 %). Adapted with permission from ref. [50].

Thanks to the protection of Au layer, the Ag@Au core/shell structure shows excellent stability against various etchants. As shown in Fig. 5d–f, the position and the intensity of the in-plane dipole plasmon resonance bands remained essentially unchanged for 4 days or longer when the nanoplates were dispersed in (D) a phosphate buffer solution, (E) a NaCl solution, and (F) a phosphate-buffered saline (PBS) solution, respectively. Although a slight decrease in peak intensity was observed in the concentrated NaCl solution treated case, it could be primarily attributed to the aggregation of the nanoplates in a solution of high ionic strength. The as-obtained core/shell structure shows outstanding stability against H2O2, a strong oxidant that can easily destroy Ag nanoplates protected by conventional methods (Fig. 3g). The inset TEM image in Fig. 3g shows the morphology of Ag@Au nanoplates stored in H2O2 for 2 h, which is quite similar to the original Ag nanoplates, further confirming their excellent stability against strong etching agents. The protecting effect can be further proved by the control experiments, in which Ag nanoplates without any protection were quickly etched by H2O2 (Fig. 3h), as evidenced by a significant shift in the peak position and a dramatic decrease in the intensity over a relatively short period. The protecting effect from thiols, such as 16-mercaptohexadecanoic acid (MHA), is also very limited. As shown in Fig. 3i, the Ag–MHA nanoplates gradually degraded in the presence of H2O2.

Stabilizing metal nanoparticles through encapsulation within a porous shell

In the physical approach, encapsulating metal nanoparticle with a porous shell to prevent the direct contact from each other is one of the major strategies [110114]. Figure 6b shows the typical TEM image of a model catalysts, in which Pt nanoparticles with size around 2–3 nm were supported on the surface of ~100 nm silica beads. After calcined at 1075 K, only several large crystalline particles with size ~10–20 nm can be observed (Fig. 6c), confirming the occurrence of a serious sintering process. When the SiO2/Pt composite was coated with a thin layer of silica, great improvement in thermal stability has been achieved. As shown in Fig. 6d–f, no obvious particle size increase can be observed after being treated at 1075 K, implying successful suppressing of sintering.

Fig. 6 
            Illustration of the use of mesoporous silica layers for protection against sintering of dispersed metal nanoparticles [125]. As shown schematically in panel (a), the original small Pt nanoparticles supported on SiO2 beads [panel (b)] coalesce into a few larger structures [panel (c)] after calcination at 1075 K. The corresponding TEM images from a catalyst coated with a mesoporous SiO2 shell, shown in panels (e) and (f), prove the enhanced stability afforded by such treatment [panel (d)]. Adapted with permission from ref. [125].
Fig. 6

Illustration of the use of mesoporous silica layers for protection against sintering of dispersed metal nanoparticles [125]. As shown schematically in panel (a), the original small Pt nanoparticles supported on SiO2 beads [panel (b)] coalesce into a few larger structures [panel (c)] after calcination at 1075 K. The corresponding TEM images from a catalyst coated with a mesoporous SiO2 shell, shown in panels (e) and (f), prove the enhanced stability afforded by such treatment [panel (d)]. Adapted with permission from ref. [125].

Although the encapsulation of metal nanoparticles within oxide shells could significantly suppress the sintering, dense coating will block the accessibility of the active sites and make them useless. It is thus desirable to have a porous shell coating rather than a dense one. We have developed a new concept, called “surface-protected etching”, to precisely convert dense oxide shell to porous shell and explored their potential applications in catalysis [66, 115124]. In a typical surface-protected etching process, solid oxide shells were first protected by a layer of polymeric ligands, followed by a preferential etching of material from the interior of the oxide shells. Due to the protection of polymeric ligands, the original size of the oxide shell could be well maintained, while the selective etching at the interior produces porous or even hollow structures. To transfer dense silica shells to porous ones, we used PVP as the polymeric ligand and NaOH as the etchant. By varying the etching time, the porosity of silica shell could be readily tuned, resulting in well-controlled catalytic activity. The hydrogenation of cis-2-butene using Pt nanoparticles as the catalyst has been used as the model reaction, as shown in Fig. 7. When SiO2/Pt/SiO2 composites without etching were used as the catalyst, a slow reaction rate, indicated by the slope of the traces at time zero, has been observed, which can be attributed to the microporous nature of sol-gel obtained silica shell. The reaction rate can be significantly higher when the samples were treated with a surface-protected etching process (by a factor of almost an order of magnitude).

Fig. 7 
            Illustration of the use of the surface-protected etching procedure to regain the activity of encapsulated catalysts. The kinetics of hydrogenation of cis-2-butene, in the form of the accumulation of the butane product in the batch reactor as a function of reaction time, is reported for three catalysts, namely, for 10-nm-mesoporous-silica/1-wt%-Pt/silica-bead catalysts as prepared (bottom TEM, blue) and after 40 (center TEM, green) and 60 (top TEM, red) min of etching [66]. Pt re-exposure and full catalytic activity are regained upon the 60 min etching treatment. Adapted with permission from ref. [66].
Fig. 7

Illustration of the use of the surface-protected etching procedure to regain the activity of encapsulated catalysts. The kinetics of hydrogenation of cis-2-butene, in the form of the accumulation of the butane product in the batch reactor as a function of reaction time, is reported for three catalysts, namely, for 10-nm-mesoporous-silica/1-wt%-Pt/silica-bead catalysts as prepared (bottom TEM, blue) and after 40 (center TEM, green) and 60 (top TEM, red) min of etching [66]. Pt re-exposure and full catalytic activity are regained upon the 60 min etching treatment. Adapted with permission from ref. [66].

Design and synthesis of functional composites for catalysis

A grand challenge facing the research community in the area of solar energy harvesting is to develop materials capable of transforming solar photons efficiently into chemical or electrical work [127, 128]. As one of the most widely studied photocatalysts, titanium dioxide (TiO2) has attracted much attention. It has been realized that higher photocatalytic activity can be achieved by improving many aspects of TiO2, including crystal phase, specific surface area, surface properties, and crystallinity [129131]. Recently, Sanz and co-workers have theoretically predicted that Au pre-adsorption on TiO2 surfaces can significantly stabilize implanted N, increase the reachable amount of N loading in the oxide, and enhance Au-surface adhesion energy due to an electron transfer from the Au 6s orbitals to the partially occupied N 2p orbitals [132]. However, the thermal stability of Au nanoparticles during the nitrogen doping process has been a concern in making the predicted structure.

We have designed and synthesized a sandwich structure that comprises a SiO2 core, a layer of Au nanoparticles, and a TiO2 shell (Fig. 8a,b), to solve the problem [126]. Non-metal doping, including N and C, is achieved by introducing 3-aminopropyl-triethoxysilane, which originally acts as the ligands to enhance the adhesion of Au nanoparticles to support surface but later upon decomposition serves as N and C source for doping. Au nanoparticles are sandwiched between the SiO2 core and TiO2 shell. Compared to the traditional Au/TiO2 composites, in which Au nanoparticles are usually decorated on the surface of TiO2 such that they are unstable during high-temperature calcination and application, the as-obtained sandwich structures have several intrinsic advantages. First, Au nanoparticles are embedded inside the TiO2 matrix, which protects them from moving and forming aggregations, making them more stable, as evidenced by the remained particle size of Au nanoparticles (Fig. 8c). Second, the encapsulation increases the contact area between Au nanoparticles and TiO2 matrix and therefore allows more efficient electron transfer than the previous anchored cases. The unique sandwich-structure also offers the small and tunable crystal grains of titania nanoparticles (~8–15 nm) in the outer shell and the synergetic effect of the interfacial non-metal doping and plasmonic metal decoration, which are all believed to be beneficial to the photocatalytic efficiency. Additionally, the large Au/TiO2 interface makes it possible to achieve high catalytic performance with only a small loading amount of Au (~0.1 %), thus dramatically cutting down the cost and making the catalysts practically useful in practical applications. The expected sandwich structure has been clearly confirmed by the energy-dispersive X-ray elemental mapping characterization, as shown in Fig. 8c. The as-prepared composites show outstanding catalytic activity under direct sunlight. As shown in Fig. 9, the sandwich structure can completely decompose RhB molecules within 40 min, while a commercial P25 photocatalyst can decompose only ~38 % and a commercial anatase powder can remove only ~27 %.

Fig. 8 
          (a) Schematic illustration of the fabrication process of sandwich-structured SiO2@Au@TiO2 photocatalysts; (b) a typical TEM image of the composite photocatalyst; (c) elemental mapping of a single particle, with the distribution of individual elements shown in the bottom row. Adapted with permission from ref. [126].
Fig. 8

(a) Schematic illustration of the fabrication process of sandwich-structured SiO2@Au@TiO2 photocatalysts; (b) a typical TEM image of the composite photocatalyst; (c) elemental mapping of a single particle, with the distribution of individual elements shown in the bottom row. Adapted with permission from ref. [126].

Fig. 9 
          Photodegradation of RhB without catalyst and with commercial anatase TiO2, commercial P25, and SiO2@Au@TiO2 (Au loading amount is 0.1 wt %) as the photocatalysts under direct sunlight illumination. All conversions are referred to the same total weight of TiO2 catalyst. Adapted with permission from ref. [126].
Fig. 9

Photodegradation of RhB without catalyst and with commercial anatase TiO2, commercial P25, and SiO2@Au@TiO2 (Au loading amount is 0.1 wt %) as the photocatalysts under direct sunlight illumination. All conversions are referred to the same total weight of TiO2 catalyst. Adapted with permission from ref. [126].

Conclusion and outlook

Since the establishment of the first industrialized catalytic process in 1913, catalysis has been of critical importance to the modern society. The development of nanotechnology over the past two decades has provided some great opportunities to the catalysis research community. This review article highlights some of our efforts in nanomaterials engineering and their application in catalysis. By using Ag nanoplates as the model system, we figured out the key component for making such an anisotropic structure. On the basis of improved understanding, we are able to develop a reliable method to make uniform Ag nanoplates with controllable size and shape. Both chemical and physical approaches have been developed to stabilize metal nanoparticles. And a highly visible-light-active photocatalyst has been designed and synthesized by utilizing our improved knowledge in both synthesis and stabilization of nanomaterials.

Although impressive progress has been achieved, many challenges still remain. For example, in the synthesis of nanomaterials, more mechanistic studies should be conducted to reveal the underlying mechanisms and establish some general understanding that works for many different systems. With the advanced nanotechnology, people should be able to design nanocatalysts with computer-based software and then produce them quickly and reliably. To achieve industrial applications, we should be able to obtain nanocatalysts with desired morphology as well as desired properties, such as high activity and high selectivity, in large-volume production and in a low-cost manner. Anisotropic nanomaterials, e.g., Ag nanoplates, have shown great potential in catalytic applications due to the structure-related properties. However, they have been suffering from both shape and size change during the catalytic process, resulting in the loss of catalytic activity. More effort should be devoted to developing new strategies to enhance their stabilities. Additionally, future research in this field should pay more attention to the interfacial interactions and the possible synergistic effect among different components.


Corresponding author: Qiao Zhang, Department of Chemistry, University of California Berkeley, Berkeley, CA 94720, USA, e-mail: ; and Materials Science Division, Lawrence Berkeley National Laboratory, Berkeley, CA 94720, USA


A collection of invited, peer-reviewed articles by the winners of the 2013 IUPAC Prize for Young Chemists.


We thank all the colleagues and collaborators whose names are mentioned in the references. Q.Z. would like to thank Profs. A. Paul Alivisatos and Gabor A. Somorjai for their support in his postdoctoral research at the University of California Berkeley. Q.Z. thanks IUPAC for the prestigious IUPAC Prize for Young Chemists as well as the invitation to write this critical review.

References

[1] G. A. Somorjai, K. McCrea. Appl. Catal., A 222, 3 (2001).10.1016/S0926-860X(01)00825-0Search in Google Scholar

[2] G. A. Somorjai, Y. G. Borodko. Catal. Lett. 76, 1 (2001).10.1023/A:1016711323302Search in Google Scholar

[3] G. A. Somorjai, Y. M. Li. Introduction to Surface Chemistry and Catalysis, 2nd ed., John Wiley, Hoboken, NJ (2010).Search in Google Scholar

[4] H. C. Zeng. Acc. Chem. Res. 46, 226 (2013).10.1021/ar3001662Search in Google Scholar PubMed

[5] C. Burda, X. B. Chen, R. Narayanan, M. A. El-Sayed. Chem. Rev. 105, 1025 (2005).10.1021/cr030063aSearch in Google Scholar PubMed

[6] Q. Zhang, Y. D. Yin. Chem. Commun. 49, 215 (2013).10.1039/C2CC34733DSearch in Google Scholar

[7] Q. Zhang, N. Li, J. Goebl, Z. D. Lu, Y. D. Yin. J. Am. Chem. Soc. 133, 18931 (2011).10.1021/ja2080345Search in Google Scholar PubMed

[8] G. J. Hutchings, M. Haruta. Appl. Catal., A 291, 2 (2005).10.1016/j.apcata.2005.05.044Search in Google Scholar

[9] I. Pastoriza-Santos, L. M. Liz-Marzan. J. Mater. Chem. 18, 1724 (2008).10.1039/b716538bSearch in Google Scholar

[10] J. E. Millstone, S. J. Hurst, G. S. Metraux, J. I. Cutler, C. A. Mirkin. Small 5, 646 (2009).10.1002/smll.200801480Search in Google Scholar PubMed

[11] K. M. Mayer, J. H. Hafner. Chem. Rev. 111, 3828 (2011).10.1021/cr100313vSearch in Google Scholar PubMed

[12] M. R. Jones, K. D. Osberg, R. J. Macfarlane, M. R. Langille, C. A. Mirkin. Chem. Rev. 111, 3736 (2011).10.1021/cr1004452Search in Google Scholar PubMed

[13] M. Rycenga, C. M. Cobley, J. Zeng, W. Y. Li, C. H. Moran, Q. Zhang, D. Qin, Y. N. Xia. Chem. Rev. 111, 3669 (2011).10.1021/cr100275dSearch in Google Scholar PubMed PubMed Central

[14] Y. G. Sun, W. G. Yang, Y. Ren, L. Wang, C. H. Lei. Small 7, 606 (2011).10.1002/smll.201002201Search in Google Scholar PubMed

[15] R. C. Jin, Y. W. Cao, C. A. Mirkin, K. L. Kelly, G. C. Schatz, J. G. Zheng. Science 294, 1901 (2001).10.1126/science.1066541Search in Google Scholar PubMed

[16] R. C. Jin, Y. C. Cao, E. C. Hao, G. S. Metraux, G. C. Schatz, C. A. Mirkin. Nature 425, 487 (2003).10.1038/nature02020Search in Google Scholar PubMed

[17] M. Maillard, P. R. Huang, L. Brus. Nano Lett. 3, 1611 (2003).10.1021/nl034666dSearch in Google Scholar

[18] C. Xue, C. A. Mirkin. Angew. Chem., Int. Ed. 46, 2036 (2007).10.1002/anie.200604637Search in Google Scholar PubMed

[19] Y. G. Sun, Y. N. Xia. Adv. Mater. 15, 695 (2003).10.1002/adma.200304652Search in Google Scholar

[20] C. Xue, G. S. Metraux, J. E. Millstone, C. A. Mirkin. J. Am. Chem. Soc. 130, 8337 (2008).10.1021/ja8005258Search in Google Scholar PubMed PubMed Central

[21] J. Zhang, M. R. Langille, C. A. Mirkin. J. Am. Chem. Soc. 132, 12502 (2010).10.1021/ja106008bSearch in Google Scholar PubMed

[22] Y. G. Sun, G. P. Wiederrecht. Small 3, 1964 (2007).10.1002/smll.200700484Search in Google Scholar PubMed

[23] S. H. Chen, D. L. Carroll. Nano Lett. 2, 1003 (2002).10.1021/nl025674hSearch in Google Scholar

[24] Y. G. Sun, B. Mayers, Y. N. Xia. Nano Lett. 3, 675 (2003).10.1021/nl034140tSearch in Google Scholar

[25] Y. Xiong, A. R. Siekkinen, J. Wang, Y. Yin, M. J. Kim, Y. Xia. J. Mater. Chem. 17, 2600 (2007).10.1039/b705253gSearch in Google Scholar

[26] E. C. Hao, K. L. Kelly, J. T. Hupp, G. C. Schatz. J. Am. Chem. Soc. 124, 15182 (2002).10.1021/ja028336rSearch in Google Scholar PubMed

[27] L. P. Jiang, S. Xu, J. M. Zhu, J. R. Zhang, J. J. Zhu, H. Y. Chen. Inorg. Chem. 43, 5877 (2004).10.1021/ic049529dSearch in Google Scholar PubMed

[28] G. Wulff. Z. Kristallogr. 34, 449 (1901).10.1524/zkri.1901.34.1.449Search in Google Scholar

[29] J. L. Elechiguerra, J. Reyes-Gasga, M. J. Yacaman. J. Mater. Chem. 16, 3906 (2006).10.1039/b607128gSearch in Google Scholar

[30] E. Leontidis, K. Kleitou, T. Kyprianidou-Leodidou, V. Bekiari, P. Lianos. Langmuir 18, 3659 (2002).10.1021/la011368sSearch in Google Scholar

[31] R. W. Berriman, R. H. Herz. Nature 180, 293 (1957).10.1038/180293a0Search in Google Scholar

[32] D. R. Hamilton, R. G. Seidensticker. J. Appl. Phys. 31, 1165 (1960).10.1063/1.1735796Search in Google Scholar

[33] Y. Hosoya, S. Urabe. J. Imaging Sci. Technol. 42, 487 (1998).Search in Google Scholar

[34] Y. Xia, Y. J. Xiong, B. Lim, S. E. Skrabalak. Angew. Chem., Int. Ed. 48, 60 (2009).10.1002/anie.200802248Search in Google Scholar PubMed PubMed Central

[35] G. S. Metraux, C. A. Mirkin. Adv Mater 17, 412 (2005).10.1002/adma.200401086Search in Google Scholar

[36] N. Li, Q. Zhang, S. Quinlivan, J. Goebl, Y. Gan, Y. D. Yin. ChemPhysChem 13, 2526 (2012).10.1002/cphc.201101018Search in Google Scholar PubMed

[37] J. P. Hoare. Standard Potentials in Aqueous Solution, Marcel Dekker, New York (1985).Search in Google Scholar

[38] C. M. Ho, S. K. Yau, C. N. Lok, M. H. So, C. M. Che. Chem. Asian J. 5, 285 (2010).10.1002/asia.200900387Search in Google Scholar PubMed

[39] T. Parnklang, C. Lertvachirapaiboon, P. Pienpinijtham, K. Wongravee, C. Thammacharoen, S. Ekgasit. RSC Adv. 3, 12886 (2013).10.1039/c3ra41486hSearch in Google Scholar

[40] Q. Zhang, Y. X. Hu, S. R. Guo, J. Goebl, Y. D. Yin. Nano Lett. 10, 5037 (2010).10.1021/nl1032233Search in Google Scholar PubMed

[41] J. W. Mullin. Crystallization, 1st ed., Butterworth, London (1972).Search in Google Scholar

[42] C. B. Gao, Q. Zhang, Z. D. Lu, Y. D. Yin. J. Am. Chem. Soc. 133, 19706 (2011).10.1021/ja209647dSearch in Google Scholar PubMed

[43] C. B. Gao, J. Vuong, Q. Zhang, Y. D. Liu, Y. D. Yin. Nanoscale 4, 2875 (2012).10.1039/c2nr30300kSearch in Google Scholar PubMed

[44] C. B. Gao, J. Goebl, Y. D. Yin. J. Mater. Chem. C 1, 3898 (2013).10.1039/c3tc30365aSearch in Google Scholar

[45] S. Djokic. Bioinorg. Chem. Appl. 436 (2008).Search in Google Scholar

[46] L. Bickford, J. Sun, K. Fu, N. Lewinski, V. Nammalvar, J. Chang, R. Drezek. Nanotechnology 19, 315102 (2008).10.1088/0957-4484/19/31/315102Search in Google Scholar PubMed

[47] C. Loo, A. Lowery, N. J. Halas, J. West, R. Drezek. Nano Lett. 5, 709 (2005).10.1021/nl050127sSearch in Google Scholar PubMed

[48] H. Goesmann, C. Feldmann. Angew. Chem., Int. Ed. 49, 1362 (2010).10.1002/anie.200903053Search in Google Scholar PubMed

[49] L. R. Hirsch, A. M. Gobin, A. R. Lowery, F. Tam, R. A. Drezek, N. J. Halas, J. L. West. Ann. Biomed. Eng. 34, 15 (2006).10.1007/s10439-005-9001-8Search in Google Scholar PubMed

[50] C. B. Gao, Z. D. Lu, Y. Liu, Q. Zhang, M. F. Chi, Q. Cheng, Y. D. Yin. Angew. Chem., Int. Ed. 51, 5629 (2012).10.1002/anie.201108971Search in Google Scholar PubMed

[51] Y. G. Sun, Y. N. Xia. Science 298, 2176 (2002).10.1126/science.1077229Search in Google Scholar PubMed

[52] J. Zeng, X. H. Xia, M. Rycenga, P. Henneghan, Q. G. Li, Y. N. Xia. Angew. Chem., Int. Ed. 50, 244 (2011).10.1002/anie.201005549Search in Google Scholar PubMed

[53] B. Tang, S. P. Xu, J. An, B. Zhao, W. Q. Xu. J. Phys. Chem. C 113, 7025 (2009).10.1021/jp810711aSearch in Google Scholar

[54] V. Germain, J. Li, D. Ingert, Z. L. Wang, M. P. Pileni. J. Phys. Chem. B 107, 8717 (2003).10.1021/jp0303826Search in Google Scholar

[55] T. C. R. Rocha, D. Zanchet. J. Phys. Chem. C 111, 6989 (2007).10.1021/jp0702696Search in Google Scholar

[56] D. Aherne, D. M. Ledwith, M. Gara, J. M. Kelly. Adv. Funct. Mater. 18, 2005 (2008).10.1002/adfm.200800233Search in Google Scholar

[57] J. Goebl, Q. Zhang, L. He, Y. D. Yin. Angew. Chem., Int. Ed. 51, 552 (2012).10.1002/anie.201107240Search in Google Scholar PubMed

[58] G. S. Metraux, Y. C. Cao, R. C. Jin, C. A. Mirkin. Nano Lett. 3, 519 (2003).10.1021/nl034097+Search in Google Scholar

[59] X. C. Jiang, Q. H. Zeng, A. B. Yu. Langmuir 23, 2218 (2007).10.1021/la062797zSearch in Google Scholar PubMed

[60] Y. Chen, C. G. Wang, Z. F. Ma, Z. M. Su. Nanotechnology 18, 325602 (2007).10.1088/0957-4484/18/32/325602Search in Google Scholar

[61] J. An, B. Tang, X. L. Zheng, J. Zhou, F. X. Dong, S. P. Xu, Y. Wang, B. Zhao, W. Q. Xu. J. Phys. Chem. C 112, 15176 (2008).10.1021/jp802694pSearch in Google Scholar

[62] Y. S. Xia, J. J. Ye, K. H. Tan, J. J. Wang, G. Yang. Anal. Chem. 85, 6241 (2013).10.1021/ac303591nSearch in Google Scholar PubMed

[63] Q. Zhang, J. P. Ge, T. Pham, J. Goebl, Y. X. Hu, Z. Lu, Y. D. Yin. Angew. Chem., Int. Ed. 48, 3516 (2009).10.1002/anie.200900545Search in Google Scholar PubMed

[64] B. Tang, J. An, X. L. Zheng, S. P. Xu, D. M. Li, J. Zhou, B. Zhao, W. Q. Xu. J. Phys. Chem. C 112, 18361 (2008).10.1021/jp806486fSearch in Google Scholar

[65] J. Gabaldon, M. Bore, A. Datye. Top. Catal. 44, 253 (2007).10.1007/s11244-007-0298-4Search in Google Scholar

[66] Q. Zhang, I. Lee, J. P. Ge, F. Zaera, Y. D. Yin. Adv. Funct. Mater. 20, 2201 (2010).10.1002/adfm.201000428Search in Google Scholar

[67] C. L. Fang, K. Qian, J. H. Zhu, S. B. Wang, X. X. Lv, S. H. Yu. Nanotechnology 19, 125601 (2008).10.1088/0957-4484/19/12/125601Search in Google Scholar PubMed

[68] C. Pham-Huu, N. Keller, L. J. Charbonniere, R. Ziessle, M. J. Ledoux. Chem. Commun. 1871 (2000).10.1039/b005306fSearch in Google Scholar

[69] J. P. Tessonnier, L. Pesant, G. Ehret, M. J. Ledoux, C. Pham-Huu. Appl. Catal., A 288, 203 (2005).10.1016/j.apcata.2005.04.034Search in Google Scholar

[70] W. F. Yan, S. Brown, Z. W. Pan, S. M. Mahurin, S. H. Overbury, S. Dai. Angew. Chem., Int. Ed. 45, 3614 (2006).10.1002/anie.200503808Search in Google Scholar PubMed

[71] Z. Ma, H. F. Yin, S. H. Overbury, S. Dai. Catal. Lett. 126, 20 (2008).10.1007/s10562-008-9627-xSearch in Google Scholar

[72] Z. Ma, F. Zaera. Encyclopedia of Inorganic Chemistry, R. B. King (Ed.), 2nd ed., John Wiley, New York (2005).Search in Google Scholar

[73] C. M. Wang, K. N. Fan, Z. P. Liu. J. Phys. Chem. C 111, 13539 (2007).10.1021/jp074530uSearch in Google Scholar

[74] W. Yan, S. M. Mahurin, Z. Pan, S. H. Overbury, S. Dai. J. Am. Chem. Soc. 127, 10480 (2005).10.1021/ja053191kSearch in Google Scholar PubMed

[75] W. Yan, S. Mahurin, S. Overbury, S. Dai. Top. Catal. 39, 199 (2006).10.1007/s11244-006-0058-xSearch in Google Scholar

[76] O. A. Barias, A. Holmen, E. A. Blekkan. J. Catal. 158, 1 (1996).10.1006/jcat.1996.0001Search in Google Scholar

[77] A. Iglesias-Juez, A. M. Beale, K. Maaijen, T. C. Weng, P. Glatzel, B. M. Weckhuysen. J. Catal. 276, 268 (2010).10.1016/j.jcat.2010.09.018Search in Google Scholar

[78] P. Bindra, S. J. Clouser, E. Yeager. J. Electrochem. Soc. 126, 1631 (1979).10.1149/1.2129345Search in Google Scholar

[79] P. J. Ferreira, G. J. la O′, Y. Shao-Horn, D. Morgan, R. Makharia, S. Kocha, H. A. Gasteiger. J. Electrochem. Soc. 152, A2256 (2005).10.1149/1.2050347Search in Google Scholar

[80] Z. Chen, M. Waje, W. Li, Y. Yan. Angew. Chem., Int. Ed. 46, 4060 (2007).10.1002/anie.200700894Search in Google Scholar PubMed

[81] E. Antolini, J. R. C. Salgado, E. R. Gonzalez. J. Power Sources 160, 957 (2006).10.1016/j.jpowsour.2006.03.006Search in Google Scholar

[82] J. B. Joo, P. Kim, W. Kim, Y. Kim, J. Yi. J. Appl. Electrochem. 39, 135 (2009).10.1007/s10800-008-9645-9Search in Google Scholar

[83] Y. Takasu, R. Matsuyama, S. Konishi, W. Sugimoto, Y. Murakami. Electrochem. Solid State Lett. 8, B34 (2005).10.1149/1.1979448Search in Google Scholar

[84] H. R. Colon-Mercado, B. N. Popov. J. Power Sources 155, 253 (2006).10.1016/j.jpowsour.2005.05.011Search in Google Scholar

[85] H. R. Colon-Mercado, H. Kim, B. N. Popov. Electrochem. Commun. 6, 795 (2004).10.1016/j.elecom.2004.05.028Search in Google Scholar

[86] P. Yu, M. Pemberton, P. Plasse. J. Power Sources 144, 11 (2005).10.1016/j.jpowsour.2004.11.067Search in Google Scholar

[87] S. Koh, J. Leisch, M. F. Toney, P. Strasser. J. Phys. Chem. C 111, 3744 (2007).10.1021/jp067269aSearch in Google Scholar

[88] Z. C. Liu, C. F. Yu, I. A. Rusakova, D. X. Huang, P. Strasser. Top. Catal. 49, 241 (2008).10.1007/s11244-008-9083-2Search in Google Scholar

[89] K. A. Flanagan, J. A. Sullivan, H. Mueller-Bunz. Langmuir 23, 12508 (2007).10.1021/la7015897Search in Google Scholar PubMed

[90] R. Narayanan, M. A. El-Sayed. J. Am. Chem. Soc. 125, 8340 (2003).10.1021/ja035044xSearch in Google Scholar PubMed

[91] S. Kidambi, J. Dai, J. Li, M. L. Bruening. J. Am. Chem. Soc. 126, 2658 (2004).10.1021/ja038804cSearch in Google Scholar PubMed

[92] B. P. S. Chauhan, J. S. Rathore, T. Bandoo. J. Am. Chem. Soc. 126, 8493 (2004).10.1021/ja049604jSearch in Google Scholar PubMed

[93] A. B. Lowe, B. S. Sumerlin, M. S. Donovan, C. L. McCormick. J. Am. Chem. Soc. 124, 11562 (2002).10.1021/ja020556hSearch in Google Scholar PubMed

[94] Y. Liu, C. Khemtong, J. Hu. Chem. Commun. 398 (2004).10.1039/B313210MSearch in Google Scholar PubMed

[95] S. V. Ley, C. Mitchell, D. Pears, C. Ramarao, J. Q. Yu, W. Zhou. Org. Lett. 5, 4665 (2003).10.1021/ol0358509Search in Google Scholar PubMed

[96] R. B. Grubbs. Polym. Rev. 47, 197 (2007).10.1080/15583720701271245Search in Google Scholar

[97] M. Zhao, L. Sun, R. M. Crooks. J. Am. Chem. Soc. 120, 4877 (1998).10.1021/ja980438nSearch in Google Scholar

[98] M. Zhao, R. M. Crooks. Angew. Chem., Int. Ed. 38, 364 (1999).10.1002/(SICI)1521-3773(19990201)38:3<364::AID-ANIE364>3.0.CO;2-LSearch in Google Scholar

[99] R. M. Crooks, M. Zhao, L. Sun, V. Chechik, L. K. Yeung. Acc. Chem. Res. 34, 181 (2001).10.1021/ar000110aSearch in Google Scholar

[100] Y. Li, J. Petroski, M. A. El-Sayed. J. Phys. Chem. B 104, 10956 (2000).10.1021/jp002569sSearch in Google Scholar

[101] M. Kralik, A. Biffis. J. Mol. Catal., A 177, 113 (2001).10.1016/S1381-1169(01)00313-2Search in Google Scholar

[102] M. Ohmori, E. Matijevic. J. Colloid Interface Sci. 150, 594 (1992).10.1016/0021-9797(92)90229-FSearch in Google Scholar

[103] A. P. Philipse, M. P. B. van Bruggen, C. Pathmamanoharan. Langmuir 10, 92 (1994).10.1021/la00013a014Search in Google Scholar

[104] L. M. Liz-Marzan, M. Giersig, P. Mulvaney. Langmuir 12, 4329 (1996).10.1021/la9601871Search in Google Scholar

[105] Y. Lu, Y. Yin, Z. Y. Li, Y. Xia. Nano Lett. 2, 785 (2002).10.1021/nl025598iSearch in Google Scholar

[106] P. M. Arnal, M. Comotti, F. Schüth. Angew. Chem. 118, 8404 (2006).10.1002/ange.200603507Search in Google Scholar

[107] M. Giersig, T. Ung, L. M. Liz-Marzan, P. Mulvaney. Adv. Mater. 9, 570 (1997).10.1002/adma.19970090712Search in Google Scholar

[108] T. Ung, L. M. Liz-Marzan, P. Mulvaney. Langmuir 14, 3740 (1998).10.1021/la980047mSearch in Google Scholar

[109] J. Liu, H. Q. Yang, F. Kleitz, Z. G. Chen, T. Y. Yang, E. Strounina, G. Q. Lu, S. Z. Qiao. Adv. Funct. Mater. 22, 591 (2012).10.1002/adfm.201101900Search in Google Scholar

[110] C. T. Kresge, M. E. Leonowicz, W. J. Roth, J. C. Vartuli, J. S. Beck. Nature 359, 710 (1992).10.1038/359710a0Search in Google Scholar

[111] M. E. Davis. Nature 417, 813 (2002).10.1038/nature00785Search in Google Scholar PubMed

[112] D. Zhao, J. Feng, Q. Huo, N. Melosh, G. H. Fredrickson, B. F. Chmelka, G. D. Stucky. Science 279, 548 (1998).10.1126/science.279.5350.548Search in Google Scholar PubMed

[113] D. Zhao, Q. Huo, J. Feng, B. F. Chmelka, G. D. Stucky. J. Am. Chem. Soc. 120, 6024 (1998).10.1021/ja974025iSearch in Google Scholar

[114] S. H. Joo, J. Y. Park, C. K. Tsung, Y. Yamada, P. D. Yang, G. A. Somorjai. Nat. Mater. 8, 126 (2009).10.1038/nmat2329Search in Google Scholar PubMed

[115] Q. Zhang, T. R. Zhang, J. P. Ge, Y. D. Yin. Nano Lett 8, 2867 (2008).10.1021/nl8016187Search in Google Scholar PubMed

[116] Q. Zhang, W. S. Wang, J. Goebl, Y. D. Yin. Nano Today 4, 494 (2009).10.1016/j.nantod.2009.10.008Search in Google Scholar

[117] Q. Zhang, J. P. Ge, J. Goebl, Y. X. Hu, Y. G. Sun, Y. D. Yin. Adv. Mater. 22, 1905 (2010).10.1002/adma.200903748Search in Google Scholar PubMed

[118] Q. Zhang, J. P. Ge, J. Goebl, Y. X. Hu, Z. D. Lu, Y. D. Yin. Nano Res. 2, 583 (2009).10.1007/s12274-009-9060-5Search in Google Scholar

[119] I. Lee, M. A. Albiter, Q. Zhang, J. P. Ge, Y. D. Yin, F. Zaera. Phys. Chem. Chem. Phys. 13, 2449 (2011).10.1039/C0CP01688HSearch in Google Scholar PubMed

[120] J. B. Joo, Q. Zhang, I. Lee, M. Dahl, F. Zaera, Y. D. Yin. Adv. Funct. Mater. 22, 166 (2012).10.1002/adfm.201101927Search in Google Scholar

[121] J. B. Joo, Q. Zhang, M. Dahl, F. Zaera, Y. D. Yin. J. Mater. Res. 28, 362 (2013).10.1557/jmr.2012.280Search in Google Scholar

[122] J. B. Joo, Q. Zhang, M. Dahl, I. Lee, J. Goebl, F. Zaera, Y. D. Yin. Energ. Environ. Sci. 5, 6321 (2012).10.1039/C1EE02533CSearch in Google Scholar

[123] Y. X. Hu, Q. Zhang, J. Goebl, T. R. Zhang, Y. D. Yin. Phys. Chem. Chem. Phys. 12, 11836 (2010).10.1039/c0cp00031kSearch in Google Scholar PubMed

[124] J. P. Ge, Q. Zhang, T. R. Zhang, Y. D. Yin. Angew. Chem., Int. Ed. 47, 8924 (2008).10.1002/anie.200803968Search in Google Scholar PubMed

[125] I. Lee, Q. Zhang, J. Ge, Y. Yin, F. Zaera. Nano Res. 4, 115 (2011).10.1007/s12274-010-0059-8Search in Google Scholar

[126] Q. Zhang, D. Q. Lima, I. Lee, F. Zaera, M. F. Chi, Y. D. Yin. Angew. Chem., Int. Ed. 50, 7088 (2011).10.1002/anie.201101969Search in Google Scholar PubMed

[127] B. Oregan, M. Gratzel. Nature 353, 737 (1991).10.1038/353737a0Search in Google Scholar

[128] C. An, S. Peng, Y. Sun. Adv. Mater. 22, 2570 (2010).10.1002/adma.200904116Search in Google Scholar PubMed

[129] T. Xia, X. B. Chen. J. Mater. Chem. A 1, 2983 (2013).10.1039/c3ta01589kSearch in Google Scholar

[130] X. B. Chen, S. H. Shen, L. J. Guo, S. S. Mao. Chem. Rev. 110, 6503 (2010).10.1021/cr1001645Search in Google Scholar PubMed

[131] X. B. Chen, L. Liu, P. Y. Yu, S. S. Mao. Science 331, 746 (2011).10.1126/science.1200448Search in Google Scholar PubMed

[132] J. Graciani, A. Nambu, J. Evans, J. A. Rodriguez, J. F. Sanz. J. Am. Chem. Soc. 130, 12056 (2008).10.1021/ja802861uSearch in Google Scholar PubMed

Published Online: 2014-01-17
Published in Print: 2014-01-22

©2014 IUPAC & De Gruyter Berlin Boston

Downloaded on 1.5.2024 from https://www.degruyter.com/document/doi/10.1515/pac-2014-5000/html
Scroll to top button