ACS Publications. Most Trusted. Most Cited. Most Read
My Activity
CONTENT TYPES

Single-Crystal Nanostructure Arrays Forming Epitaxially through Thermomechanical Nanomolding

  • Guannan Liu
    Guannan Liu
    Department of Mechanical Engineering and Materials Science, Yale University, New Haven, Connecticut 06511, United States
    More by Guannan Liu
  • Sungwoo Sohn
    Sungwoo Sohn
    Department of Mechanical Engineering and Materials Science, Yale University, New Haven, Connecticut 06511, United States
    More by Sungwoo Sohn
  • Naijia Liu
    Naijia Liu
    Department of Mechanical Engineering and Materials Science, Yale University, New Haven, Connecticut 06511, United States
    More by Naijia Liu
  • Arindam Raj
    Arindam Raj
    Department of Mechanical Engineering and Materials Science, Yale University, New Haven, Connecticut 06511, United States
    More by Arindam Raj
  • Udo D. Schwarz
    Udo D. Schwarz
    Department of Mechanical Engineering and Materials Science, Yale University, New Haven, Connecticut 06511, United States
    Department of Chemical and Environmental Engineering, Yale University, New Haven, Connecticut 06511, United States
  • , and 
  • Jan Schroers*
    Jan Schroers
    Department of Mechanical Engineering and Materials Science, Yale University, New Haven, Connecticut 06511, United States
    *Email: [email protected]
    More by Jan Schroers
Cite this: Nano Lett. 2021, 21, 23, 10054–10061
Publication Date (Web):November 22, 2021
https://doi.org/10.1021/acs.nanolett.1c03744

Copyright © 2021 The Authors. Published by American Chemical Society. This publication is licensed under

CC-BY-NC-ND 4.0.
  • Open Access

Article Views

1780

Altmetric

-

Citations

LEARN ABOUT THESE METRICS
PDF (5 MB)
Supporting Info (1)»

Abstract

For nanostructures in advanced electronic and plasmonic systems, a single-crystal structure with controlled orientation is essential. However, the fabrication of such devices has remained challenging, as current nanofabrication methods often suffer from either polycrystalline growth or the difficulty of integrating single crystals with substrates in desired orientations and locations to create functional devices. Here we report a thermomechanical method for the controlled growth of single-crystal nanowire arrays, which enables the simultaneous synthesis, alignment, and patterning of nanowires. Within such diffusion-based thermomechanical nanomolding (TMNM), the substrate material diffuses into nanosized cavities under an applied pressure gradient at a molding temperature of ∼0.4 times the material’s melting temperature. Vertically grown face-centered cubic (fcc) nanowires with the [110] direction in an epitaxial relationship with the (110) substrate are demonstrated. The ability to control the crystal structure through the substrate takes TMNM a major step further, potentially allowing all fcc and body-centered cubic (bcc) materials to be integrated as single crystals into devices.

This publication is licensed under

CC-BY-NC-ND 4.0.
  • cc licence
  • by licence
  • nc licence
  • nd licence
High-quality single-crystal nanostructures of controlled orientation are critical for realizing desirable electronic and optical properties in nanomaterials and devices, as grain boundaries are often detrimental to these properties. In electronics, single-crystal nanostructures exhibit higher electrical conductivity than polycrystalline structures, (1) which suffer from grain boundary scattering. (2,3) In plasmonics, metallic nanostructures exhibit a multitude of optical resonances associated with localized surface plasmon excitations, (4) which enable a negative refractive index, (5) subwavelength resolution imaging, (6) nanolasing, (7) and field manipulation. (8) However, grain boundaries and surface roughness in plasmonic nanostructures can cause plasmon damping, which significantly increases losses in device performance. (9,10) Crystal orientation and imperfections can also affect the mechanical, (11,12) electrochemical, (13) and thermal properties. (14)
Today’s nanofabrication methods have been exploring the fabrication of single-crystal nanostructures, in response to the high demand for such structures. (15−19) Despite recent advances, these methods are still limited in critical aspects (Figure 1). For example, widely used physical vapor deposition (PVD) methods form polycrystalline structures (Figure 1b), typically with a high defect concentration. (20,21) Chemical vapor deposition (CVD) methods are limited in the choice of precursors as feedstock and often require specific catalysts and substrates for anisotropic crystal growth to form 1D nanostructures. (22,23) Another well-established nanofabrication method, solution-based chemical synthesis, suffers from impurities, a limited selection of precursor compounds, and the challenge of preventing synthesized nanocrystals from aggregation and agglomeration in liquids. (24,25) It is also challenging to integrate solution-grown nanostructures (Figure 1a) onto substrates to build 3D-functional devices. (21) In fact, today’s scalable nanofabrication methods are incapable of producing nanostructures attached to a substrate of the same material and orientation. This is in contrast with nanolithography techniques such as electron beam lithography (EBL) (26) and focused ion beam (FIB), (27) which allow the fabrication of nanowire arrays that are identically oriented and attached to a substrate (Figure 1c) but cannot be scaled to be used for practical nanofabrication. In addition, the surface layers of FIB-created structures may have implanted gallium or argon atoms that would be detrimental to device operation. (28) In summary, the presently existing approaches to nanofabrication remain deficient in the large-scale production of single-crystal nanostructures that are in an epitaxial relation to their substrates, even though such capability is crucial for realizing the advances in plasmonics and nanoelectronics envisioned above.

Figure 1

Figure 1. Current nanofabrication methods and atomic arrangement (in 2D illustration) of the nanostructures they can produce. (a) Schematic illustration of solution-based chemical synthesis and the resulting detached single-crystal nanowire. (b) Physical vapor deposition (PVD) method and the resulting nanostructure containing grain boundaries and often polycrystals. (c) Using a focused ion beam (FIB) inside a scanning electron microscope (SEM) allows the fabrication of single-crystal nanostructures integrated in the crystal of the substrate but lacks scalability.

We have recently developed a controlled and scalable nanofabrication method called thermomechanical nanomolding (TMNM) (29−32) that holds promise for obtaining single-crystal nanostructures. TMNM exhibits high control over the size and aspect ratio of the nanowires. Nanowire arrays with aspect ratios of up to 1000 (see Figure S1, Supporting Information) and diameters down to 5 nm have been formed for a wide range of materials and structures. The nanowires’ length and diameter are limited practically by the mold dimensions, but there is no fundamental limit to these features. We have demonstrated that TMNM can be applied to a wide range of materials, including metals, alloys, and ordered phases. (29,31−33) The underlying mechanism of TMNM is based on diffusion, specifically on interface diffusion (see Figure S2 Supporting Information) for temperatures above ∼0.4TM, where TM is the melting temperature of the material. Notably, functional nanostructures fabricated by TMNM include superconductors, topological insulators, semiconductors, and phase change materials. (29) Nanowires synthesized through TMNM are almost entirely single crystals; (31) however, thus far, they have always exhibited a change in orientation and the associated multiple grains and grain boundaries in the area between the substrate material and the nanowire. The mechanisms for the observed growth direction and grain boundary formation in the nanowires are still unclear. As previously discussed, these associated grain boundaries deteriorate the performance in device applications. (2,3,9,10)
In this work, we introduce a fabrication method based on TMNM that prevents crystal reorientation and the formation of grain boundaries, resulting in the epitaxial growth of single-crystal nanowire arrays on substrates. Hence, we can form single-crystal nanostructures in a highly controllable manner that all exhibit the same crystal structure and orientation as the substrate.

Results and Discussion

ARTICLE SECTIONS
Jump To

The fabrication of epitaxially grown nanowires on a substrate uses pressure and temperature (Figure 2a). TMNM involves pressing a substrate material against a mold with nanometer cavity sizes under an applied pressure p of typically p ≈ 10 to 500 MPa at a temperature of ∼0.4TM. After this molding step, the nanostructures are released by etching the mold. Representative Ag nanowire arrays that are vertically grown through TMNM are shown in Figure 2b,c, with nanowire diameters of 120 and 40 nm.

Figure 2

Figure 2. Overview of thermomechanical nanomolding (TMNM). (a) Schematic illustration of the process flow of TMNM performed in this study. A nanosized mold, for example, anodic aluminum oxide, and a flat feedstock substrate are depicted. (b,c) Morphology of the resulting vertically grown Ag nanowire arrays with nanowire diameters of 120 and 40 nm, respectively, revealed by SEM imaging at a 30° tilted angle. Nanowires with a small aspect ratio are free-standing and precisely aligned and located as shown in panel b, and nanowires with a large aspect ratio may agglomerate due to surface tension, as shown in panel c.

Required for the epitaxial formation of nanowires is the use of a single crystal with a specific orientation as feedstock material. We observed that if this is not the case, for example, when using randomly oriented polycrystals as substrates (Figure 3a), then grains typically reorient, leading to grain boundary formation at the root of the nanowires (Figure 3b). Furthermore, a single crystal, which grows along a specific orientation, ultimately forms at a nanocavity depth that approximately coincides with the nanocavity diameter (Figure 3b). For face-centered cubic (fcc) materials, such orientation is [110]. Surprisingly, such [110] single-crystal growth is independent of the crystal orientation of the adjacent crystal in the substrate. This finding of the tendency for fcc nanowires to grow along [110] sets the requirement to use a (110) single-crystal substrate to avoid grain reorientation and associated grain boundary formation. However, because crystals tend to reorient themselves during plastic deformation under sufficient uniaxial pressure, we carried out experiments to determine under which conditions the single crystal changes its orientation and potentially forms polycrystals (Figure 3c,d and Figure 4a,b).

Figure 3

Figure 3. Orientation of nanowires with grain boundaries fabricated by TMNM. (a–d) Schematics of TMNM using polycrystalline and single-crystal substrates and the resulting microstructure and crystal orientation of the nanowires. (110) directions are denoted with red arrows. (a) Polycrystalline substrate for TMNM. (b) Crystal structure and grain orientation of a representative nanowire and its adjacent substrate area after processing in panel a. The nanowire grows along the (110) direction but forms grain boundaries at the root. (c) (110) single-crystal substrate exposed to a “high”-pressure ph during TMNM, where “high” means that ph is larger than the reorientation pressure of the crystal orientation. (d) Crystal structure and grain orientation of a representative nanowire and its adjacent substrate area after processing in panel c. Because of the acting “high” pressure ph, the (110) substrate reorients to (111). Because the growth direction deeper within the nanocavity is along (110), grain boundaries are forming at the root of the nanowire. (e) Transmission electron microscopy (TEM) image of a 40 nm nanowire with the polycrystalline root region. (f) TEM image from the area marked by the dashed red rectangle in panel e revealing the grain boundaries and multiple grains with different crystallographic orientations at the root of the nanowire. (See the original TEM image in Figure S3, Supporting Information.) (g) High-resolution TEM image from the area indicated in panel f by the yellow square showing lattice fringes with the (110) orientation along the growth direction. (h) Selected area electron diffraction (SAED) pattern obtained within the region enclosed by the dashed white circle in panel e, revealing that the nanowire root is polycrystalline. (i,j) High-magnification TEM images covering the regions within the blue and orange squares in panel f showing well-developed grain boundaries at the root of the nanowire. (See the original and additional TEM images in Figure S3, Supporting Information.)

Figure 4

Figure 4. TEM characterization of single-crystal nanostructures fabricated by TMNM. (a,b) Schematics of TMNM using a single-crystal substrate and the resulting structure of nanowires. The [110] direction is denoted with a red arrow. (a) (110) single-crystal substrate under “low” pressure pl during TMNM. (b) Crystal structure and grain orientation of a representative nanowire and its adjacent substrate area after processing. If pl is below the reorientation threshold, then nanowires of the same [110] orientation grow from the (110) single-crystal substrate in an epitaxial relationship. (c) TEM image of a 40 nm single-crystal nanowire. (d) TEM image from the selected region in yellow in panel c revealing an absence of polycrystals and grain boundaries at the root of the nanowire. (e–h) SAED patterns from four sections of the nanowire in panel c, revealing that the sample is a face-centered cubic (fcc) single crystal. (See Figure S5 in the Supporting Information for the indexing.) Scale bar: 10 1/nm. (i–l) High-resolution TEM images from the regions marked in panel d in blue and orange (i,k) and further magnifications into the areas highlighted with the white dashed square (j,l), showing lattice fringes with (110) orientation along the growth direction of the nanowire.

Crystallographic characterization via transmission electron microscopy (TEM) was used to reveal the structure of the nanowires. When using a (110) single-crystal substrate at high pressure (e.g., 500 MPa), we observed that grain reorientation takes place in both the substrate material and the nanowire. Consequently, an undesired polycrystalline structure at the nanowire’s root forms (Figure 3e–j). We propose that observed polycrystal formation at the root of the nanocavity originates from a pressure-induced reorientation of the substrate material adjacent to the nanocavity from [110] to [111]. Because the material in the nanocavity prefers to grow along the [110] direction, reorientation is required. The reorientation is accomplished by the formation of multiple grains, which gradually change their orientation such that their [110] direction, indicated by the red arrow, is eventually lined up with the nanocavity. Because the reorientation of the substrate toward [111] is pressure-induced, the use of a pressure below the material’s reorientation pressure is required to prevent reorientation. Using such a “low” pressure of ∼80 MPa, nanowires of the same [110] orientation as the (110) single-crystal substrate grow in an epitaxial manner and relationship (Figure 4c,d). Selected area electron diffraction (SAED) patterns (Figure 4e,h) from four different locations of the nanowire reveal the same [110] patterns and orientations, confirming that the nanowire is an fcc single crystal. High-resolution TEM images (Figure 4i–l) of the Ag nanowire body show clear lattice fringes representing (110) planes along the nanowire’s length. The spacing between fringes is 0.289 nm (±5% error), which precisely matches the expected spacing between (110) planes in the fcc Ag lattice. Additional high-resolution TEM images of the nanowire root and the indexed SAED patterns can be found in Figures S4 and S5 of the Supporting Information.
We explain the preferred [110] growth direction of Ag nanowires by a lower surface energy γ of the nanowire when growing along [110] (Figure 5b). Because nanowires have a large surface-to-volume ratio, the surface energy plays an important role. (34) Surface energies for different crystallographic planes vary on an fcc crystal structure, and for its three low-index surfaces (100), (111), and (110), its relative strengths are γ(111) < γ(100) < γ(110). (35,36) For example, for Ag, the surface energies are 0.713 J/m2 for {111} planes, 0.752 J/m2 for {100} planes, and 0.833 J/m2 for {110} planes. (35) Thereby, {111}-type planes have the lowest surface energy among all existing planes in an fcc crystal due to their close-packed atomic arrangement. Hence, to reduce the surface energy, an fcc nanowire prefers to orient itself so that as much of its surface area as possible represents {111} planes, which requires growth along a ⟨110⟩ direction, the close-packed direction on {111} planes. Note that to form a “prism”-like overall shape (Figure 5e), at least two of the terminating surfaces need to be of {100} type, which is the surface with the second-lowest surface energy in an fcc crystal. During TMNM, however, the nanowires are more likely trying to grow as cylinders to adopt to the largely spherical shape of the mold. To reduce the surface energy, the nanowire can then form alternating {111} and {100} surfaces as side walls to achieve the appearance of a “round” cross-section (Figure 5f). In contrast, nanowires formed with (100)- or (111)-terminated feedstock substrates would need to have, at least partially, walls with higher Miller indices and thus higher surface energies. Another advantage of terminating as much of the nanowire’s surface with {111}-type planes as possible is that they have the lowest activation energy barrier of single adatom hopping among different surface facets of fcc materials, (37) which better facilitates interface diffusion as compared with any other surface termination and consequently promotes nanowire growth during TMNM. To realize such orientation of the nanowire surfaces formed through TMNM, crystal reorientation and hence grain boundary formation are required unless the substrate is already of {110} type. To estimate when the formation of a grain boundary to enable the reorientation of the nanowire toward [110] is energetically favored, we calculate the energy of a nanowire–substrate system with the nanowire growing along a random non-⟨110⟩ direction (Figure 5a), which gives
(1)
where γav is the average surface energy that the many different (hkl) planes that must form the nanowire’s round walls would feature (with γav > γ(111)), r is the nanowire radius, and L is the nanowire length. We compare this energy to the energy of a system where the nanowire grows along a ⟨110⟩ direction, including a grain boundary (Figure 5b), to consider the required change of orientation
(2)
Here γ(111) is the surface energy of the (111) plane, and γGB is the grain boundary energy. Note that this calculation will slightly underestimate the nanowire’s surface energy, as some {100}-type surfaces will also be present. However, this underestimation is not too significant, as {100} planes feature the second lowest surface energies of all surface terminations possible in an fcc crystal, with γ(100) being only 5.5% larger than γ(111) for Ag. Consequently, the overall deviation of E2 from the actual value can be estimated to be <2.75%. An energetic preference for the [110] growth direction requires E2 < E1. This is the case when L > Lc, where Lc is the critical length. We can obtain Lc from eq 1 = eq 2
(3)
Going forward, we now use values for γ(111) of 0.713 J/m2, γ(100) of 0.752 J/m2, γ(210) of 0.87 J/m2, and γGB of 0.21 J/m2. (35,38) (See Table S1, Supporting Information.) If the nanowire does not already grow in a ⟨110⟩ direction, then γav cannot be lower than γ(100). Identifying γav with γ(100), we get Lc ≈ 4.4r. It is, however, more likely that γav will be higher than that, so choosing γav = γ(210) as a more realistic value causes Lc to already drop to 0.3r. This comparison of energies explains the polycrystalline structure at the nanowire roots as a consequence of the nanowire’s tendency to adjust its original crystal orientation, given by the crystal orientation in the substrate to the energetically preferred ⟨110⟩ growth direction. Therefore, for the formation of ⟨110⟩ single-crystal nanowires epitaxially grown from the substrate, it is imperative that the substrate also exhibits a ⟨110⟩ orientation to avoid polycrystallinity and associated grain boundary formation.

Figure 5

Figure 5. Mechanisms of grain reorientation in the substrate and in the nanowire. (a) Schematics of two unknown crystallographic orientations in the feedstock substrate and the nanowire of a typical nanowire-substrate system under uniaxial pressure. (b) Surface energy comparison among three different substrate–nanowire systems: randomly oriented substrate and nanowire along a random growth direction, randomly oriented substrate and nanowire along the [110] growth direction including a grain boundary, and (110) substrate and nanowire along the [110] growth direction. Surface I refers to a termination with a combination of various mostly non-{111} and non-{100} (hkl) planes, and surface II refers to a termination by a combination of {111} and {100} planes. (c) Rotation of a (110) plane under compression, which results in a polycrystalline substrate with {111} planes as the dominant orientation. (d) X-ray diffraction characterization of the Ag substrate before and after TMNM (experimental conditions: p = 1.3 GPa, T = 0.6TM, t = 30 s), revealing a structural change from a (110) single-crystal to a (111)-dominant polycrystal. (e) Perspective representation of a hexagonal prism-shaped fcc single-crystal nanowire that grows along the [110] direction with {111} and {100} surfaces as the side walls. (f) Schematic illustration of the atomic arrangement on the cross-section of a cylinder-shaped fcc (110) single-crystal nanowire. Zoom on the outer surface of the nanowire shows alternating {111}/{100} surfaces as side walls to achieve the appearance of a “round” shape.

It is also important to consider that in general, a single-crystal substrate is prone to become polycrystalline with grains of different orientations during the application of a uniaxial pressure (Figure 5c). This behavior originates from plastic deformation owing to dislocation slip, grain boundary formation, and grain rotation. (39,40) This phenomenon has been confirmed by X-ray diffraction (XRD) measurements (Figure 5d) on the Ag substrate material before and after TMNM, revealing the structural change from a (110) single-crystal to a polycrystal with {111} planes as the dominant crystallographic orientation. This preferred orientation may also originate from lowering of the surface energy. For the experimental conditions under which TMNM is carried out, reorientation in the substrate is kinetically enabled by the combination of pressure and temperature providing the energy required for atoms to rearrange in significant numbers toward a more low-energy state, and hence lowering the processing pressure can prevent such reorientation from occurring. Therefore, to enable the epitaxial growth of a nanowire on the substrate, the processing pressure must be below the reorientation pressure to prevent crystal reorientation toward {111} crystals in the substrate. (See Figure S6 in the Supporting Information for the XRD characterization on (110) Ag substrate before and after TMNM under “low” pressure.) We identified the reorientation pressure to be approximately three times the yield stress in the case of Ag nanowires.
Within this work, we have focused on epitaxially grown single-crystal Ag nanostructures on a single-crystal Ag substrate. We choose Ag because it is a representative fcc material. Because our mechanistic descriptions only use information about the atomic structure, the same behavior should manifest for other fcc materials. As a result, we expect that the technique can be applied to the nanofabrication of all fcc materials.
Finally, a similar mechanism should be present for body-centered cubic (bcc) materials. (See Figure S7, Supporting Information.) In a bcc structure, {110} planes have the lowest surface energy. (41) Therefore, to reduce the surface energy of the system, a bcc nanowire orients itself so that its surfaces are {110} surfaces. This requires the nanowire to grow along a ⟨111⟩ direction, the close-packed direction on {110} planes. Therefore, epitaxial growth of any bcc single-crystal nanowire array via TMNM can be expected when a pressure below the reorientation pressure and a (111) substrate are used.

Conclusions

ARTICLE SECTIONS
Jump To

The controlled fabrication of vertically aligned single-crystal nanowire arrays epitaxially grown on the substrate material is demonstrated via diffusion-based TMNM. The requirement for the epitaxial growth of single-crystal nanowires on a substrate is that the crystal orientations in the substrate and wire are identical. The crystal orientation in fcc nanowires prefers to be along the [110] direction, which is motivated by achieving the lowest possible surface energy of the associated side walls, which for [110] is {111} with some {100} in between to achieve an overall round shape of the wire. As a consequence, the formation of an fcc single crystal that continues from the substrate into the nanowire requires the use of a (110)-oriented single-crystal substrate, but because (110) single-crystal substrates tend to reorient toward the {111}-type under high pressure, the pressure exerted during formation must be limited to below the reorientation pressure to avoid reorientation and grain boundary formation. Because the formation motifs and orientation selections only depend on crystallographic-specific symmetries, we expect the method to be applicable to any fcc material. Our estimations also predict vertically aligned single-crystal bcc nanowire arrays along [111] epitaxially grown on a (111) substrate material. This suggests that most materials can be fabricated through this method as single-crystal nanowire arrays including technologically interesting materials such as magnetic materials, heterosystems, and a wide range of quantum materials. This very practical and self-regulating technique provides new insights and a new route for the nanofabrication of single-crystal structures and holds promise for the large-scale fabrication of high-performance nanoelectronic and plasmonic devices.

Methods

ARTICLE SECTIONS
Jump To

Sample Preparation

Polycrystalline and single-crystal Ag samples were prepared for nanomolding. High-purity polycrystal Ag (99.99+% in purity, from Alfa Aesar) and (110) single-crystal Ag substrates (99.999% in purity, one side optical polished <30 Å, from MTI Corporation) were cut into 2.5 mm × 2.5 mm × 0.5 mm rectangular plates. The samples were then finely polished as needed.

Thermomechanical Nanomolding

Prepared feedstock samples were pressed against anodic aluminum oxide (AAO) molds (from InRedox) at a certain pressure and processing temperature using an Instron universal testing system equipped with heating plates. The load was applied over a period of 10 min and then held for a fixed period of time. After molding, AAO molds can be etched away in 20 wt % KOH or 10 wt % phosphoric acid at room temperature for 10 h to obtain free-standing nanowire arrays.

Characterization

The length and morphology of vertically grown Ag nanowire arrays were characterized by scanning electron microscopy (SEM) using a Hitachi SU-70 analytical field emission SEM at a voltage of 10 kV. TEM was performed using an FEI Tecnai Osiris 200 kV TEM to characterize the structure of Ag nanowires. Before TEM characterization, a portion of the sample (including the substrate material, nanowires, and mold) was lifted out using a Helios G4 UX DualBeam FIB/SEM system secured on a copper-based TEM grid. The sample was thinned to ∼50 nm. The crystal orientation of the Ag substrates before and after TMNM was evaluated by X-ray diffraction with a Rigaku SmartLab X-ray diffractometer using Bragg–Brentano focusing, Cu Kα radiation, and a 2 mm beam mask.

Supporting Information

ARTICLE SECTIONS
Jump To

The Supporting Information is available free of charge at https://pubs.acs.org/doi/10.1021/acs.nanolett.1c03744.

  • Supplementary Text Sections 1−3: Diffusion Mechanisms of TMNM, Preferred Growth Direction Calculation for bcc Materials, and Reorientation Mechanism of the Substrate. Supplementary Figures S1–S7. Supplementary Table S1. Supplementary References (PDF)

Terms & Conditions

Most electronic Supporting Information files are available without a subscription to ACS Web Editions. Such files may be downloaded by article for research use (if there is a public use license linked to the relevant article, that license may permit other uses). Permission may be obtained from ACS for other uses through requests via the RightsLink permission system: http://pubs.acs.org/page/copyright/permissions.html.

Author Information

ARTICLE SECTIONS
Jump To

  • Corresponding Author
    • Jan Schroers - Department of Mechanical Engineering and Materials Science, Yale University, New Haven, Connecticut 06511, United States Email: [email protected]
  • Authors
    • Guannan Liu - Department of Mechanical Engineering and Materials Science, Yale University, New Haven, Connecticut 06511, United StatesOrcidhttps://orcid.org/0000-0002-7976-1739
    • Sungwoo Sohn - Department of Mechanical Engineering and Materials Science, Yale University, New Haven, Connecticut 06511, United States
    • Naijia Liu - Department of Mechanical Engineering and Materials Science, Yale University, New Haven, Connecticut 06511, United States
    • Arindam Raj - Department of Mechanical Engineering and Materials Science, Yale University, New Haven, Connecticut 06511, United StatesOrcidhttps://orcid.org/0000-0001-7277-6770
    • Udo D. Schwarz - Department of Mechanical Engineering and Materials Science, Yale University, New Haven, Connecticut 06511, United StatesDepartment of Chemical and Environmental Engineering, Yale University, New Haven, Connecticut 06511, United StatesOrcidhttps://orcid.org/0000-0002-5361-0342
  • Author Contributions

    J.S. and G.L. conceived the project. G.L. designed and conducted the experiments. G.L. conducted the SEM, XRD, and FIB characterization. S.S. and G.L. performed the TEM characterization. All authors contributed to the discussion of the results and the writing of the manuscript. All authors have given approval to the final version of the manuscript.

  • Notes
    The authors declare no competing financial interest.

Acknowledgments

ARTICLE SECTIONS
Jump To

We thank Dr. Min Li and Dr. Yujun Xie for their help with the FIB. This work was supported by the National Science Foundation through the Advanced Manufacturing Program (CMMI 1901613).

References

ARTICLE SECTIONS
Jump To

This article references 41 other publications.

  1. 1
    Wu, Y.; Xiang, J.; Yang, C.; Lu, W.; Lieber, C. M. Single-crystal metallic nanowires and metal/semiconductor nanowire heterostructures. Nature 2004, 430, 6165,  DOI: 10.1038/nature02674
  2. 2
    Zhang, Q. G.; Zhang, X.; Cao, B. Y.; Fujii, M.; Takahashi, K.; Ikuta, T. Influence of grain boundary scattering on the electrical properties of platinum nanofilms. Appl. Phys. Lett. 2006, 89, 114102,  DOI: 10.1063/1.2338885
  3. 3
    Sciacca, B.; van de Groep, J.; Polman, A.; Garnett, E. C. Solution-Grown Silver Nanowire Ordered Arrays as Transparent Electrodes. Adv. Mater. 2016, 28, 905909,  DOI: 10.1002/adma.201504045
  4. 4
    Henzie, J.; Lee, J.; Lee, M. H.; Hasan, W.; Odom, T. W. Nanofabrication of Plasmonic Structures. Annu. Rev. Phys. Chem. 2009, 60, 147165,  DOI: 10.1146/annurev.physchem.040808.090352
  5. 5
    Liu, N.; Guo, H. C.; Fu, L. W.; Kaiser, S.; Schweizer, H.; Giessen, H. Three-dimensional photonic metamaterials at optical frequencies. Nat. Mater. 2008, 7, 3137,  DOI: 10.1038/nmat2072
  6. 6
    Zhang, X.; Liu, Z. W. Superlenses to overcome the diffraction limit. Nat. Mater. 2008, 7, 435441,  DOI: 10.1038/nmat2141
  7. 7
    Zhou, W.; Dridi, M.; Suh, J. Y.; Kim, C. H.; Co, D. T.; Wasielewski, M. R.; Schatz, G. C.; Odom, T. W. Lasing action in strongly coupled plasmonic nanocavity arrays. Nat. Nanotechnol. 2013, 8, 506511,  DOI: 10.1038/nnano.2013.99
  8. 8
    Pendry, J. B.; Schurig, D.; Smith, D. R. Controlling electromagnetic fields. Science 2006, 312, 17801782,  DOI: 10.1126/science.1125907
  9. 9
    Bosman, M.; Zhang, L.; Duan, H. G.; Tan, S. F.; Nijhuis, C. A.; Qiu, C. W.; Yang, J. K. W. Encapsulated Annealing: Enhancing the Plasmon Quality Factor in Lithographically-Defined Nanostructures. Sci. Rep. 2015, 4, 16,  DOI: 10.1038/srep05537
  10. 10
    Park, J. H.; Ambwani, P.; Manno, M.; Lindquist, N. C.; Nagpal, P.; Oh, S. H.; Leighton, C.; Norris, D. J. Single-Crystalline Silver Films for Plasmonics. Adv. Mater. 2012, 24, 39883992,  DOI: 10.1002/adma.201200812
  11. 11
    Cao, A.; Wei, Y.; Ma, E. Grain boundary effects on plastic deformation and fracture mechanisms in Cu nanowires: Molecular dynamics simulations. Phys. Rev. B: Condens. Matter Mater. Phys. 2008, 77, 195429,  DOI: 10.1103/PhysRevB.77.195429
  12. 12
    Greer, J. R.; Jang, D.; Gu, X. W. Exploring deformation mechanisms in nanostructured materials. JOM 2012, 64, 12411252,  DOI: 10.1007/s11837-012-0438-6
  13. 13
    Borkowska, Z.; Tymosiak-Zielinska, A.; Shul, G. Electrooxidation of methanol on polycrystalline and single crystal gold electrodes. Electrochim. Acta 2004, 49, 12091220,  DOI: 10.1016/j.electacta.2003.09.046
  14. 14
    Zhang, Q. G.; Cao, B. Y.; Zhang, X.; Fujii, M.; Takahashi, K. Influence of grain boundary scattering on the electrical and thermal conductivities of polycrystalline gold nanofilms. Phys. Rev. B: Condens. Matter Mater. Phys. 2006, 74, 134109,  DOI: 10.1103/PhysRevB.74.134109
  15. 15
    Butburee, T.; Bai, Y.; Wang, H. J.; Chen, H. J.; Wang, Z. L.; Liu, G.; Zou, J.; Khemthong, P.; Lu, G. Q. M.; Wang, L. Z. 2D Porous TiO2 Single-Crystalline Nanostructure Demonstrating High Photo-Electrochemical Water Splitting Performance. Adv. Mater. 2018, 30, 1705666,  DOI: 10.1002/adma.201705666
  16. 16
    Choi, S. H.; Kim, H. J.; Song, B.; Kim, Y. I.; Han, G.; Nguyen, H. T. T.; Ko, H.; Boandoh, S.; Choi, J. H.; Oh, C. S.; Cho, H. J.; Jin, J. W.; Won, Y. S.; Lee, B. H.; Yun, S. J.; Shin, B. G.; Jeong, H. Y.; Kim, Y. M.; Han, Y. K.; Lee, Y. H.; Kim, S. M.; Kim, K. K. Epitaxial Single-Crystal Growth of Transition Metal Dichalcogenide Monolayers via the Atomic Sawtooth Au Surface. Adv. Mater. 2021, 33, 2006601,  DOI: 10.1002/adma.202006601
  17. 17
    Xie, D. G.; Nie, Z. Y.; Shinzato, S.; Yang, Y. Q.; Liu, F. X.; Ogata, S.; Li, J.; Ma, E.; Shan, Z. W. Controlled growth of single-crystalline metal nanowires via thermomigration across a nanoscale junction. Nat. Commun. 2019, 10, 18,  DOI: 10.1038/s41467-019-12416-x
  18. 18
    Xiong, Y. J.; Xia, Y. N. Shape-controlled synthesis of metal nanostructures: The case of palladium. Adv. Mater. 2007, 19, 33853391,  DOI: 10.1002/adma.200701301
  19. 19
    Zhu, Y. C.; Bando, Y.; Xue, D. F.; Golberg, D. Oriented assemblies of ZnS one-dimensional nanostructures. Adv. Mater. 2004, 16, 831834,  DOI: 10.1002/adma.200305486
  20. 20
    Grayli, S. V.; Zhang, X.; MacNab, F. C.; Kamal, S.; Star, D.; Leach, G. W. Scalable, Green Fabrication of Single-Crystal Noble Metal Films and Nanostructures for Low-Loss Nanotechnology Applications. ACS Nano 2020, 14, 75817592,  DOI: 10.1021/acsnano.0c03466
  21. 21
    Zhang, H. Y.; Kinnear, C.; Mulvaney, P. Fabrication of Single-Nanocrystal Arrays. Adv. Mater. 2020, 32, 1904551,  DOI: 10.1002/adma.201904551
  22. 22
    Chou, Y. C.; Hillerich, K.; Tersoff, J.; Reuter, M. C.; Dick, K. A.; Ross, F. M. Atomic-Scale Variability and Control of III-V Nanowire Growth Kinetics. Science 2014, 343, 281284,  DOI: 10.1126/science.1244623
  23. 23
    Jacobsson, D.; Panciera, F.; Tersoff, J.; Reuter, M. C.; Lehmann, S.; Hofmann, S.; Dick, K. A.; Ross, F. M. Interface dynamics and crystal phase switching in GaAs nanowires. Nature 2016, 531, 317322,  DOI: 10.1038/nature17148
  24. 24
    Ditlbacher, H.; Hohenau, A.; Wagner, D.; Kreibig, U.; Rogers, M.; Hofer, F.; Aussenegg, F. R.; Krenn, J. R. Silver nanowires as surface plasmon resonators. Phys. Rev. Lett. 2005, 95, 257403,  DOI: 10.1103/PhysRevLett.95.257403
  25. 25
    Kim, F.; Sohn, K.; Wu, J. S.; Huang, J. X. Chemical Synthesis of Gold Nanowires in Acidic Solutions. J. Am. Chem. Soc. 2008, 130, 1444214443,  DOI: 10.1021/ja806759v
  26. 26
    Hicks, E. M.; Zou, S. L.; Schatz, G. C.; Spears, K. G.; Van Duyne, R. P.; Gunnarsson, L.; Rindzevicius, T.; Kasemo, B.; Kall, M. Controlling plasmon line shapes through diffractive coupling in linear arrays of cylindrical nanoparticles fabricated by electron beam lithography. Nano Lett. 2005, 5, 10651070,  DOI: 10.1021/nl0505492
  27. 27
    Tseng, A. A. Recent developments in nanofabrication using focused ion beams. Small 2005, 1, 924939,  DOI: 10.1002/smll.200500113
  28. 28
    Rubanov, S.; Munroe, P. R. FIB-induced damage in silicon. J. Microsc. 2004, 214, 213221,  DOI: 10.1111/j.0022-2720.2004.01327.x
  29. 29
    Liu, N. J.; Xie, Y. J.; Liu, G. N.; Sohn, S.; Raj, A.; Han, G. X.; Wu, B. Z.; Cha, J. J.; Liu, Z.; Schroers, J. General Nanomolding of Ordered Phases. Phys. Rev. Lett. 2020, 124, 036102  DOI: 10.1103/PhysRevLett.124.036102
  30. 30
    Liu, Z. One-step fabrication of crystalline metal nanostructures by direct nanoimprinting below melting temperatures. Nat. Commun. 2017, 8, 17,  DOI: 10.1038/ncomms14910
  31. 31
    Liu, Z.; Han, G. X.; Sohn, S.; Liu, N. J.; Schroers, J. Nanomolding of Crystalline Metals: The Smaller the Easier. Phys. Rev. Lett. 2019, 122, 036101  DOI: 10.1103/PhysRevLett.122.036101
  32. 32
    Liu, N.; Liu, G.; Raj, A.; Sohn, S.; Morales, M.; Liu, J.; Schroers, J. Unleashing Nanofabrication through Thermomechanical Nanomolding. Science Advances 2021, 7, eabi8795
  33. 33
    Raj, A.; Liu, N.; Liu, G.; Sohn, S.; Xiang, J.; Liu, Z.; Schroers, J. Nanomolding of Gold and Gold-Silicon Heterostructures at Room Temperature. ACS Nano 2021, 15, 1427514284,  DOI: 10.1021/acsnano.1c02636
  34. 34
    Li, S.; Ding, X.; Li, J.; Ren, X.; Sun, J.; Ma, E. High-efficiency mechanical energy storage and retrieval using interfaces in nanowires. Nano Lett. 2010, 10, 17741779,  DOI: 10.1021/nl100263p
  35. 35
    Wen, Y. N.; Zhang, H. M. Surface energy calculation of the fcc metals by using the MAEAM. Solid State Commun. 2007, 144, 163167,  DOI: 10.1016/j.ssc.2007.07.012
  36. 36
    Hoefelmeyer, J. D.; Niesz, K.; Somorjai, G. A.; Tilley, T. D. Radial anisotropic growth of rhodium nanoparticles. Nano Lett. 2005, 5, 435438,  DOI: 10.1021/nl048100g
  37. 37
    Sun, J.; He, L. B.; Lo, Y. C.; Xu, T.; Bi, H. C.; Sun, L. T.; Zhang, Z.; Mao, S. X.; Li, J. Liquid-like pseudoelasticity of sub-10-nm crystalline silver particles. Nat. Mater. 2014, 13, 10071012,  DOI: 10.1038/nmat4105
  38. 38
    Barg, A. I.; Rabkin, E.; Gust, W. Faceting Transformation and Energy of a Sigma-3 Grain-Boundary in Silver. Acta Metall. Mater. 1995, 43, 40674074,  DOI: 10.1016/0956-7151(95)00094-C
  39. 39
    Yang, G.; Park, S. J. Deformation of Single Crystals, Polycrystalline Materials, and Thin Films: A Review. Materials 2019, 12, 2003,  DOI: 10.3390/ma12122003
  40. 40
    Wang, L.; Teng, J.; Liu, P.; Hirata, A.; Ma, E.; Zhang, Z.; Chen, M.; Han, X. Grain rotation mediated by grain boundary dislocations in nanocrystalline platinum. Nat. Commun. 2014, 5, 17,  DOI: 10.1038/ncomms5402
  41. 41
    Wen, Y. N.; Zhang, J. M. Surface energy calculation of the bcc metals by using the MAEAM. Comput. Mater. Sci. 2008, 42, 281285,  DOI: 10.1016/j.commatsci.2007.07.016

Cited By

ARTICLE SECTIONS
Jump To

This article is cited by 4 publications.

  1. Bozhao Wu, Yupeng Wu, Yangyang Pan, Ze Liu. Nanoscale deformation of crystalline metals: Experiments and simulations. International Journal of Plasticity 2023, 161 , 103501. https://doi.org/10.1016/j.ijplas.2022.103501
  2. Jun-Xiang Xiang, Ze Liu. Observation of enhanced nanoscale creep flow of crystalline metals enabled by controlling surface wettability. Nature Communications 2022, 13 (1) https://doi.org/10.1038/s41467-022-35703-6
  3. Mehrdad T. Kiani, Judy J. Cha. Nanomolding of topological nanowires. APL Materials 2022, 10 (8) , 080904. https://doi.org/10.1063/5.0096400
  4. Ze Liu, Naijia Liu, Jan Schroers. Nanofabrication through molding. Progress in Materials Science 2022, 125 , 100891. https://doi.org/10.1016/j.pmatsci.2021.100891
  • Abstract

    Figure 1

    Figure 1. Current nanofabrication methods and atomic arrangement (in 2D illustration) of the nanostructures they can produce. (a) Schematic illustration of solution-based chemical synthesis and the resulting detached single-crystal nanowire. (b) Physical vapor deposition (PVD) method and the resulting nanostructure containing grain boundaries and often polycrystals. (c) Using a focused ion beam (FIB) inside a scanning electron microscope (SEM) allows the fabrication of single-crystal nanostructures integrated in the crystal of the substrate but lacks scalability.

    Figure 2

    Figure 2. Overview of thermomechanical nanomolding (TMNM). (a) Schematic illustration of the process flow of TMNM performed in this study. A nanosized mold, for example, anodic aluminum oxide, and a flat feedstock substrate are depicted. (b,c) Morphology of the resulting vertically grown Ag nanowire arrays with nanowire diameters of 120 and 40 nm, respectively, revealed by SEM imaging at a 30° tilted angle. Nanowires with a small aspect ratio are free-standing and precisely aligned and located as shown in panel b, and nanowires with a large aspect ratio may agglomerate due to surface tension, as shown in panel c.

    Figure 3

    Figure 3. Orientation of nanowires with grain boundaries fabricated by TMNM. (a–d) Schematics of TMNM using polycrystalline and single-crystal substrates and the resulting microstructure and crystal orientation of the nanowires. (110) directions are denoted with red arrows. (a) Polycrystalline substrate for TMNM. (b) Crystal structure and grain orientation of a representative nanowire and its adjacent substrate area after processing in panel a. The nanowire grows along the (110) direction but forms grain boundaries at the root. (c) (110) single-crystal substrate exposed to a “high”-pressure ph during TMNM, where “high” means that ph is larger than the reorientation pressure of the crystal orientation. (d) Crystal structure and grain orientation of a representative nanowire and its adjacent substrate area after processing in panel c. Because of the acting “high” pressure ph, the (110) substrate reorients to (111). Because the growth direction deeper within the nanocavity is along (110), grain boundaries are forming at the root of the nanowire. (e) Transmission electron microscopy (TEM) image of a 40 nm nanowire with the polycrystalline root region. (f) TEM image from the area marked by the dashed red rectangle in panel e revealing the grain boundaries and multiple grains with different crystallographic orientations at the root of the nanowire. (See the original TEM image in Figure S3, Supporting Information.) (g) High-resolution TEM image from the area indicated in panel f by the yellow square showing lattice fringes with the (110) orientation along the growth direction. (h) Selected area electron diffraction (SAED) pattern obtained within the region enclosed by the dashed white circle in panel e, revealing that the nanowire root is polycrystalline. (i,j) High-magnification TEM images covering the regions within the blue and orange squares in panel f showing well-developed grain boundaries at the root of the nanowire. (See the original and additional TEM images in Figure S3, Supporting Information.)

    Figure 4

    Figure 4. TEM characterization of single-crystal nanostructures fabricated by TMNM. (a,b) Schematics of TMNM using a single-crystal substrate and the resulting structure of nanowires. The [110] direction is denoted with a red arrow. (a) (110) single-crystal substrate under “low” pressure pl during TMNM. (b) Crystal structure and grain orientation of a representative nanowire and its adjacent substrate area after processing. If pl is below the reorientation threshold, then nanowires of the same [110] orientation grow from the (110) single-crystal substrate in an epitaxial relationship. (c) TEM image of a 40 nm single-crystal nanowire. (d) TEM image from the selected region in yellow in panel c revealing an absence of polycrystals and grain boundaries at the root of the nanowire. (e–h) SAED patterns from four sections of the nanowire in panel c, revealing that the sample is a face-centered cubic (fcc) single crystal. (See Figure S5 in the Supporting Information for the indexing.) Scale bar: 10 1/nm. (i–l) High-resolution TEM images from the regions marked in panel d in blue and orange (i,k) and further magnifications into the areas highlighted with the white dashed square (j,l), showing lattice fringes with (110) orientation along the growth direction of the nanowire.

    Figure 5

    Figure 5. Mechanisms of grain reorientation in the substrate and in the nanowire. (a) Schematics of two unknown crystallographic orientations in the feedstock substrate and the nanowire of a typical nanowire-substrate system under uniaxial pressure. (b) Surface energy comparison among three different substrate–nanowire systems: randomly oriented substrate and nanowire along a random growth direction, randomly oriented substrate and nanowire along the [110] growth direction including a grain boundary, and (110) substrate and nanowire along the [110] growth direction. Surface I refers to a termination with a combination of various mostly non-{111} and non-{100} (hkl) planes, and surface II refers to a termination by a combination of {111} and {100} planes. (c) Rotation of a (110) plane under compression, which results in a polycrystalline substrate with {111} planes as the dominant orientation. (d) X-ray diffraction characterization of the Ag substrate before and after TMNM (experimental conditions: p = 1.3 GPa, T = 0.6TM, t = 30 s), revealing a structural change from a (110) single-crystal to a (111)-dominant polycrystal. (e) Perspective representation of a hexagonal prism-shaped fcc single-crystal nanowire that grows along the [110] direction with {111} and {100} surfaces as the side walls. (f) Schematic illustration of the atomic arrangement on the cross-section of a cylinder-shaped fcc (110) single-crystal nanowire. Zoom on the outer surface of the nanowire shows alternating {111}/{100} surfaces as side walls to achieve the appearance of a “round” shape.

  • References

    ARTICLE SECTIONS
    Jump To

    This article references 41 other publications.

    1. 1
      Wu, Y.; Xiang, J.; Yang, C.; Lu, W.; Lieber, C. M. Single-crystal metallic nanowires and metal/semiconductor nanowire heterostructures. Nature 2004, 430, 6165,  DOI: 10.1038/nature02674
    2. 2
      Zhang, Q. G.; Zhang, X.; Cao, B. Y.; Fujii, M.; Takahashi, K.; Ikuta, T. Influence of grain boundary scattering on the electrical properties of platinum nanofilms. Appl. Phys. Lett. 2006, 89, 114102,  DOI: 10.1063/1.2338885
    3. 3
      Sciacca, B.; van de Groep, J.; Polman, A.; Garnett, E. C. Solution-Grown Silver Nanowire Ordered Arrays as Transparent Electrodes. Adv. Mater. 2016, 28, 905909,  DOI: 10.1002/adma.201504045
    4. 4
      Henzie, J.; Lee, J.; Lee, M. H.; Hasan, W.; Odom, T. W. Nanofabrication of Plasmonic Structures. Annu. Rev. Phys. Chem. 2009, 60, 147165,  DOI: 10.1146/annurev.physchem.040808.090352
    5. 5
      Liu, N.; Guo, H. C.; Fu, L. W.; Kaiser, S.; Schweizer, H.; Giessen, H. Three-dimensional photonic metamaterials at optical frequencies. Nat. Mater. 2008, 7, 3137,  DOI: 10.1038/nmat2072
    6. 6
      Zhang, X.; Liu, Z. W. Superlenses to overcome the diffraction limit. Nat. Mater. 2008, 7, 435441,  DOI: 10.1038/nmat2141
    7. 7
      Zhou, W.; Dridi, M.; Suh, J. Y.; Kim, C. H.; Co, D. T.; Wasielewski, M. R.; Schatz, G. C.; Odom, T. W. Lasing action in strongly coupled plasmonic nanocavity arrays. Nat. Nanotechnol. 2013, 8, 506511,  DOI: 10.1038/nnano.2013.99
    8. 8
      Pendry, J. B.; Schurig, D.; Smith, D. R. Controlling electromagnetic fields. Science 2006, 312, 17801782,  DOI: 10.1126/science.1125907
    9. 9
      Bosman, M.; Zhang, L.; Duan, H. G.; Tan, S. F.; Nijhuis, C. A.; Qiu, C. W.; Yang, J. K. W. Encapsulated Annealing: Enhancing the Plasmon Quality Factor in Lithographically-Defined Nanostructures. Sci. Rep. 2015, 4, 16,  DOI: 10.1038/srep05537
    10. 10
      Park, J. H.; Ambwani, P.; Manno, M.; Lindquist, N. C.; Nagpal, P.; Oh, S. H.; Leighton, C.; Norris, D. J. Single-Crystalline Silver Films for Plasmonics. Adv. Mater. 2012, 24, 39883992,  DOI: 10.1002/adma.201200812
    11. 11
      Cao, A.; Wei, Y.; Ma, E. Grain boundary effects on plastic deformation and fracture mechanisms in Cu nanowires: Molecular dynamics simulations. Phys. Rev. B: Condens. Matter Mater. Phys. 2008, 77, 195429,  DOI: 10.1103/PhysRevB.77.195429
    12. 12
      Greer, J. R.; Jang, D.; Gu, X. W. Exploring deformation mechanisms in nanostructured materials. JOM 2012, 64, 12411252,  DOI: 10.1007/s11837-012-0438-6
    13. 13
      Borkowska, Z.; Tymosiak-Zielinska, A.; Shul, G. Electrooxidation of methanol on polycrystalline and single crystal gold electrodes. Electrochim. Acta 2004, 49, 12091220,  DOI: 10.1016/j.electacta.2003.09.046
    14. 14
      Zhang, Q. G.; Cao, B. Y.; Zhang, X.; Fujii, M.; Takahashi, K. Influence of grain boundary scattering on the electrical and thermal conductivities of polycrystalline gold nanofilms. Phys. Rev. B: Condens. Matter Mater. Phys. 2006, 74, 134109,  DOI: 10.1103/PhysRevB.74.134109
    15. 15
      Butburee, T.; Bai, Y.; Wang, H. J.; Chen, H. J.; Wang, Z. L.; Liu, G.; Zou, J.; Khemthong, P.; Lu, G. Q. M.; Wang, L. Z. 2D Porous TiO2 Single-Crystalline Nanostructure Demonstrating High Photo-Electrochemical Water Splitting Performance. Adv. Mater. 2018, 30, 1705666,  DOI: 10.1002/adma.201705666
    16. 16
      Choi, S. H.; Kim, H. J.; Song, B.; Kim, Y. I.; Han, G.; Nguyen, H. T. T.; Ko, H.; Boandoh, S.; Choi, J. H.; Oh, C. S.; Cho, H. J.; Jin, J. W.; Won, Y. S.; Lee, B. H.; Yun, S. J.; Shin, B. G.; Jeong, H. Y.; Kim, Y. M.; Han, Y. K.; Lee, Y. H.; Kim, S. M.; Kim, K. K. Epitaxial Single-Crystal Growth of Transition Metal Dichalcogenide Monolayers via the Atomic Sawtooth Au Surface. Adv. Mater. 2021, 33, 2006601,  DOI: 10.1002/adma.202006601
    17. 17
      Xie, D. G.; Nie, Z. Y.; Shinzato, S.; Yang, Y. Q.; Liu, F. X.; Ogata, S.; Li, J.; Ma, E.; Shan, Z. W. Controlled growth of single-crystalline metal nanowires via thermomigration across a nanoscale junction. Nat. Commun. 2019, 10, 18,  DOI: 10.1038/s41467-019-12416-x
    18. 18
      Xiong, Y. J.; Xia, Y. N. Shape-controlled synthesis of metal nanostructures: The case of palladium. Adv. Mater. 2007, 19, 33853391,  DOI: 10.1002/adma.200701301
    19. 19
      Zhu, Y. C.; Bando, Y.; Xue, D. F.; Golberg, D. Oriented assemblies of ZnS one-dimensional nanostructures. Adv. Mater. 2004, 16, 831834,  DOI: 10.1002/adma.200305486
    20. 20
      Grayli, S. V.; Zhang, X.; MacNab, F. C.; Kamal, S.; Star, D.; Leach, G. W. Scalable, Green Fabrication of Single-Crystal Noble Metal Films and Nanostructures for Low-Loss Nanotechnology Applications. ACS Nano 2020, 14, 75817592,  DOI: 10.1021/acsnano.0c03466
    21. 21
      Zhang, H. Y.; Kinnear, C.; Mulvaney, P. Fabrication of Single-Nanocrystal Arrays. Adv. Mater. 2020, 32, 1904551,  DOI: 10.1002/adma.201904551
    22. 22
      Chou, Y. C.; Hillerich, K.; Tersoff, J.; Reuter, M. C.; Dick, K. A.; Ross, F. M. Atomic-Scale Variability and Control of III-V Nanowire Growth Kinetics. Science 2014, 343, 281284,  DOI: 10.1126/science.1244623
    23. 23
      Jacobsson, D.; Panciera, F.; Tersoff, J.; Reuter, M. C.; Lehmann, S.; Hofmann, S.; Dick, K. A.; Ross, F. M. Interface dynamics and crystal phase switching in GaAs nanowires. Nature 2016, 531, 317322,  DOI: 10.1038/nature17148
    24. 24
      Ditlbacher, H.; Hohenau, A.; Wagner, D.; Kreibig, U.; Rogers, M.; Hofer, F.; Aussenegg, F. R.; Krenn, J. R. Silver nanowires as surface plasmon resonators. Phys. Rev. Lett. 2005, 95, 257403,  DOI: 10.1103/PhysRevLett.95.257403
    25. 25
      Kim, F.; Sohn, K.; Wu, J. S.; Huang, J. X. Chemical Synthesis of Gold Nanowires in Acidic Solutions. J. Am. Chem. Soc. 2008, 130, 1444214443,  DOI: 10.1021/ja806759v
    26. 26
      Hicks, E. M.; Zou, S. L.; Schatz, G. C.; Spears, K. G.; Van Duyne, R. P.; Gunnarsson, L.; Rindzevicius, T.; Kasemo, B.; Kall, M. Controlling plasmon line shapes through diffractive coupling in linear arrays of cylindrical nanoparticles fabricated by electron beam lithography. Nano Lett. 2005, 5, 10651070,  DOI: 10.1021/nl0505492
    27. 27
      Tseng, A. A. Recent developments in nanofabrication using focused ion beams. Small 2005, 1, 924939,  DOI: 10.1002/smll.200500113
    28. 28
      Rubanov, S.; Munroe, P. R. FIB-induced damage in silicon. J. Microsc. 2004, 214, 213221,  DOI: 10.1111/j.0022-2720.2004.01327.x
    29. 29
      Liu, N. J.; Xie, Y. J.; Liu, G. N.; Sohn, S.; Raj, A.; Han, G. X.; Wu, B. Z.; Cha, J. J.; Liu, Z.; Schroers, J. General Nanomolding of Ordered Phases. Phys. Rev. Lett. 2020, 124, 036102  DOI: 10.1103/PhysRevLett.124.036102
    30. 30
      Liu, Z. One-step fabrication of crystalline metal nanostructures by direct nanoimprinting below melting temperatures. Nat. Commun. 2017, 8, 17,  DOI: 10.1038/ncomms14910
    31. 31
      Liu, Z.; Han, G. X.; Sohn, S.; Liu, N. J.; Schroers, J. Nanomolding of Crystalline Metals: The Smaller the Easier. Phys. Rev. Lett. 2019, 122, 036101  DOI: 10.1103/PhysRevLett.122.036101
    32. 32
      Liu, N.; Liu, G.; Raj, A.; Sohn, S.; Morales, M.; Liu, J.; Schroers, J. Unleashing Nanofabrication through Thermomechanical Nanomolding. Science Advances 2021, 7, eabi8795
    33. 33
      Raj, A.; Liu, N.; Liu, G.; Sohn, S.; Xiang, J.; Liu, Z.; Schroers, J. Nanomolding of Gold and Gold-Silicon Heterostructures at Room Temperature. ACS Nano 2021, 15, 1427514284,  DOI: 10.1021/acsnano.1c02636
    34. 34
      Li, S.; Ding, X.; Li, J.; Ren, X.; Sun, J.; Ma, E. High-efficiency mechanical energy storage and retrieval using interfaces in nanowires. Nano Lett. 2010, 10, 17741779,  DOI: 10.1021/nl100263p
    35. 35
      Wen, Y. N.; Zhang, H. M. Surface energy calculation of the fcc metals by using the MAEAM. Solid State Commun. 2007, 144, 163167,  DOI: 10.1016/j.ssc.2007.07.012
    36. 36
      Hoefelmeyer, J. D.; Niesz, K.; Somorjai, G. A.; Tilley, T. D. Radial anisotropic growth of rhodium nanoparticles. Nano Lett. 2005, 5, 435438,  DOI: 10.1021/nl048100g
    37. 37
      Sun, J.; He, L. B.; Lo, Y. C.; Xu, T.; Bi, H. C.; Sun, L. T.; Zhang, Z.; Mao, S. X.; Li, J. Liquid-like pseudoelasticity of sub-10-nm crystalline silver particles. Nat. Mater. 2014, 13, 10071012,  DOI: 10.1038/nmat4105
    38. 38
      Barg, A. I.; Rabkin, E.; Gust, W. Faceting Transformation and Energy of a Sigma-3 Grain-Boundary in Silver. Acta Metall. Mater. 1995, 43, 40674074,  DOI: 10.1016/0956-7151(95)00094-C
    39. 39
      Yang, G.; Park, S. J. Deformation of Single Crystals, Polycrystalline Materials, and Thin Films: A Review. Materials 2019, 12, 2003,  DOI: 10.3390/ma12122003
    40. 40
      Wang, L.; Teng, J.; Liu, P.; Hirata, A.; Ma, E.; Zhang, Z.; Chen, M.; Han, X. Grain rotation mediated by grain boundary dislocations in nanocrystalline platinum. Nat. Commun. 2014, 5, 17,  DOI: 10.1038/ncomms5402
    41. 41
      Wen, Y. N.; Zhang, J. M. Surface energy calculation of the bcc metals by using the MAEAM. Comput. Mater. Sci. 2008, 42, 281285,  DOI: 10.1016/j.commatsci.2007.07.016
  • Supporting Information

    Supporting Information

    ARTICLE SECTIONS
    Jump To

    The Supporting Information is available free of charge at https://pubs.acs.org/doi/10.1021/acs.nanolett.1c03744.

    • Supplementary Text Sections 1−3: Diffusion Mechanisms of TMNM, Preferred Growth Direction Calculation for bcc Materials, and Reorientation Mechanism of the Substrate. Supplementary Figures S1–S7. Supplementary Table S1. Supplementary References (PDF)


    Terms & Conditions

    Most electronic Supporting Information files are available without a subscription to ACS Web Editions. Such files may be downloaded by article for research use (if there is a public use license linked to the relevant article, that license may permit other uses). Permission may be obtained from ACS for other uses through requests via the RightsLink permission system: http://pubs.acs.org/page/copyright/permissions.html.

Pair your accounts.

Export articles to Mendeley

Get article recommendations from ACS based on references in your Mendeley library.

Pair your accounts.

Export articles to Mendeley

Get article recommendations from ACS based on references in your Mendeley library.

You’ve supercharged your research process with ACS and Mendeley!

STEP 1:
Click to create an ACS ID

Please note: If you switch to a different device, you may be asked to login again with only your ACS ID.

Please note: If you switch to a different device, you may be asked to login again with only your ACS ID.

Please note: If you switch to a different device, you may be asked to login again with only your ACS ID.

MENDELEY PAIRING EXPIRED
Your Mendeley pairing has expired. Please reconnect