Skip to content
BY-NC-ND 3.0 license Open Access Published by De Gruyter October 6, 2018

Glancing angle deposition meets colloidal lithography: a new evolution in the design of nanostructures

  • Bin Ai ORCID logo EMAIL logo and Yiping Zhao
From the journal Nanophotonics

Abstract

The combination of colloidal lithography and glancing angle deposition facilitates a new powerful fabrication technique – shadow sphere lithography (SSL), which can greatly expand the variety and complexity of nanostructures fabricated using simple evaporation and colloidal monolayer templates. Their applications have been widely investigated in plasmonics and associated fields. Here, we present an overview of the principle of SSL, followed by different strategies of utilizing SSL to design various nanostructures by changing the nanosphere monolayer masks, deposition configurations, different ways to combine deposition and etching, etc. Typical nanostructures fabricated by SSL, including nanorods on nanospheres, patchy nanospheres, nanotriangles, nanoring, nanocrescents, etc., are introduced. Recent optical applications of these plasmonic nanostructures are also summarized. It is expected that this review will inspire more ingenious designs of plasmonic nanostructures by SSL for advanced and smart applications.

1 Introduction

The periodical array of a close-packed microsphere monolayer was first used as a shadow mask in 1981 for the deposition of platinum disks by Fischer and Zingsheim [1]. Ever since then, this simple technique, termed as natural lithography [2], nanosphere lithography [3], [4], or colloidal lithography (CL) [5], to make microscale and nanoscale patterns has been widely investigated and has evolved to one of the simple and powerful nanostructure patterning processes. Such a CL method is based on the use of colloidal crystals [three-dimensional (3D) CCs] or monolayers [two-dimensional (2D) CCs] (to be general, later, we refer to both as CCs) as masks. In most cases, CCs used for CL mostly are close-packed; however, sometimes, dispersed colloidal layers have been applied. A CL method normally consists of four main procedures: (1) preparation of CCs; (2) modification of CCs; (3) colloidal mask assisted deposition; and (4) colloidal mask assisted etching. Monodisperse microspheres/nanospheres with sizes ranging from tens of micrometers down to tens of nanometers can easily be synthesized by conventional emulsion polymerization and sol-gel techniques or cheaply purchased from commercial companies [5]. Owing to their good monodispersity in size and shape, they can be self-assembled into 2D and 3D periodic CCs [6], [7]. Both the ordered arrays of microspheres/nanospheres and those of the interstices between the spheres of CCs can be used as templates or masks for surface patterning. The CCs can be either etched or deformed to increase the structural complexity of the template or mask. Such a CL process has paved a simple and low-cost route for nanopatterning with a flexibility of feature size <100 nm. Figure 1 shows some examples of typical CL processes to fabricate various nanostructures. Using 2D CCs as masks during evaporation, triangular nanoparticles [8], nanodots [9], and thin film with nanohole arrays [10] have been realized on the exposed substrates (Figure 1A). While using the 2D CCs as masks for etching, nanodisks [11], nanotips [12], or nanopillars [13] have been fabricated on polymer, silica, silicon, and other substrates (Figure 1B). By metal deposition, infiltration, or imprint using etched nanospheres as molds, ordered arrays of spherical voids [15], [21] and nanoshells [14], [22] have been produced (Figure 1C). Through dewetting around nanospheres, nanorings of polymers [16], carbon nanotubes (CNTs) [17], or nanoparticles [18] can be obtained (Figure 1D). In addition to nanostructure arrays with simple geometric shapes, by combining other nanofabrication techniques to modify or sculpture individual nanospheres, more complicated or 3D nanostructures can be produced. For example, by site-selective deposition or etching, nanospheres with asymmetric shape or functional decoration have been realized [19], [20] (Figure 1E), which is difficult or impossible to be obtained by other synthetic routes.

Figure 1: CL methods to fabricate nanostructures.(A) Colloidal mask-assisted deposition to fabricate triangular nanoparticles [8] (left), nanodots [9] (middle), and nanohole arrays [10] (right). (B) Colloidal mask-assisted etching to fabricate nanodisks (top left) [11], nanotips [12] (top right), and nanopillars [13] (bottom). (C) Deposition or imprinting with nanospheres as molds to fabricate spherical nanoshells [14] (left) and nanovoids [15] (right). (D) Dewetting around nanospheres to fabricate nanorings of PS polymers (left) [16], carbon nanotubes (CNTs) [17] (middle), and CdSe nanoparticles [18] (right). (E) Site-selective deposition or etching to fabricate patchy spheres [19] (upper panel) and silica particle labeled with Au NPs [20] (bottom panel).
Figure 1:

CL methods to fabricate nanostructures.

(A) Colloidal mask-assisted deposition to fabricate triangular nanoparticles [8] (left), nanodots [9] (middle), and nanohole arrays [10] (right). (B) Colloidal mask-assisted etching to fabricate nanodisks (top left) [11], nanotips [12] (top right), and nanopillars [13] (bottom). (C) Deposition or imprinting with nanospheres as molds to fabricate spherical nanoshells [14] (left) and nanovoids [15] (right). (D) Dewetting around nanospheres to fabricate nanorings of PS polymers (left) [16], carbon nanotubes (CNTs) [17] (middle), and CdSe nanoparticles [18] (right). (E) Site-selective deposition or etching to fabricate patchy spheres [19] (upper panel) and silica particle labeled with Au NPs [20] (bottom panel).

Among the four common procedures for CL, material deposition is a key means to design and diversify the resulting nanostructures. Many physical vapor deposition (PVD) methods working in high vacuum or ultra-high vacuum, such as thermal evaporation, electron beam evaporation, and molecular beam epitaxy, as well as other low-vacuum deposition techniques, such as sputtering growth, pulsed laser deposition, and chemical vapor deposition, have been applied in CL. During the deposition, not only the structure of the CC template but also the deposition configuration can be tuned to design different nanostructures. In particular, most PVD deposition techniques can be readily converted into a glancing angle deposition (GLAD) configuration, if the deposition vapors can be intentionally engineered to be highly directional and the substrate temperature is low enough so that limited surface smoothening effects, i.e. low surface diffusion of adatoms, could occur. GLAD alone is another powerful micro-/nano-fabrication technique based on a PVD process where highly collimated incident vapor flux arrives on the substrate at a large angle with respect to the substrate surface normal, causing nanorod-like growth due to a self-shadowing effect [23]. Using GLAD, numerous nanostructures have been built by flexibly controlling the deposition parameters, such as incident angle θ, azimuthal angle φ, substrate rotation, deposition rate, substrate temperature, and deposition materials [24], [25], [26], [27].

A typical GLAD configuration is shown in Figure 2A [28]. Basically, GLAD is an extension of the commonly used oblique angle deposition (OAD) [29] in the thin film deposition community, coupled with the azimuthal rotation of the substrate. In OAD, the incident vapor flux arrives at the substrate at a large angle θ with respect to substrate surface normal, which can be tuned by changing the orientation of the substrate through a stepper motor. The geometric shadowing effect and surface diffusion of the adatoms account for the main mechanisms for the formation of the nanostructures (Figure 2B), revealing two characteristic features of OAD: self-shadow effect and self-alignment [30]. At the beginning of the deposition, islands with different sizes are randomly formed on the substrates. As the deposition proceeds, short islands will be shadowed by the nearby tall ones and cannot receive more vapor, while the tall islands continuously receive the vapor and grow longer and larger. In the meantime, the adatoms move along the surface due to the surface diffusion and intend to smoothen the islands. As long as the surface diffusion cannot dominate the self-shadowing effect, a layer of nanocolumns tilted toward the incident vapor direction is formed (Figure 3A and A′) [31]. When an additional motor is added to control the azimuthal rotation of the substrate holder, a so-called GLAD configuration is formed. By a single OAD deposition, tilted nanorods are fabricated by fixing the incident angle θ at a relatively large value (>75°). When continuously rotating the substrate azimuthally with an appropriate constant speed, vertically aligned nanocolumns are formed (Figure 3B and B′) [31]. If the substrate rotation speed is intentionally slowed down compared to the deposition rate, a continuous formation of nanocolumns along different directions results, since at each rotation interval there will be enough amount of vapor deposited onto the substrate to form nanocolumns with a preferred orientation, this results in a continuous formation of nanocolumns along different directions. Thus, helical nanorod will be formed (Figure 3C and C′) [31]. By controlling the time interval between each revolution and the incident angle, one can control the length and diameter of the rod, the diameter of the helical pitch, and the number of turns of the helical nanorod structure or to even form zigzag nanorods (Figure 3D and D′) [23], C-shape nanorods (Figure 3E and E′) [32], and bead-like nanorods [31] (Figure 3F and F′). In addition to the diverse shapes of nanostructures that can be formed using GLAD, a wide variety of materials, including metals such as Ag, Mg, Ni, V, and Ta, semiconductors such as Si and Ge, and dielectrics such as MgF2, WO3, and TiO2, can be used in GLAD [38]. The only limit for the materials is their ability to be deposited via a PVD technique. By combining the morphological design and the depositions of different materials, composite and heterogeneous nanostructures have been realized (Figure 3G–K and G′–K′) [33], [34], [35], [36], [37]. The heterogeneous nanostructures can be obtained through a series of depositions of different materials by altering the deposition angle as well as the substrate rotation, while composite nanostructures can be fabricated through co-deposition, in which two or more materials are evaporated and deposited on the substrates simultaneously.

Figure 2: Principles of GLAD [28].(A) The deposition configuration of GLAD. (B) The geometric shadowing effect and surface diffusion of the adatoms of GLAD for forming nanostructures.
Figure 2:

Principles of GLAD [28].

(A) The deposition configuration of GLAD. (B) The geometric shadowing effect and surface diffusion of the adatoms of GLAD for forming nanostructures.

Figure 3: SEM images of typical nanorod arrays fabricated by GLAD.Morphological sculpture: (A, A′) tilted Si nanorods deposited at θ=80° [31]; (B, B′) vertically aligned Si nanorods deposited at θ=86° [31]; (C, C′) regular array of helical Si nanorods [31]; (D, D′) zigzag nanorods [23]; (E, E′) C-shape MgF2 nanorod array [32]; and (F, F′) bead-like Si nanorod array [31]. Heterostructure design [33], [34], [35], [36], [37]: (G) multilayer nanorods and (G′) multilayer helical Si/Ni nanorod array, (H) side-coated nanorods and (H′) Ag side-coated L-shaped Si nanorods, (I) sandwiched nanorods and (I′) V/Mg sandwiched nanoblades for hydrogen storage, (J) core-shell nanorods and (J′) WO3/TiO2 quasi-core-shell nanorods for effective photocatalyst and (K) checker board nanorods and (K′) Ta/Si checker board nanorods.
Figure 3:

SEM images of typical nanorod arrays fabricated by GLAD.

Morphological sculpture: (A, A′) tilted Si nanorods deposited at θ=80° [31]; (B, B′) vertically aligned Si nanorods deposited at θ=86° [31]; (C, C′) regular array of helical Si nanorods [31]; (D, D′) zigzag nanorods [23]; (E, E′) C-shape MgF2 nanorod array [32]; and (F, F′) bead-like Si nanorod array [31]. Heterostructure design [33], [34], [35], [36], [37]: (G) multilayer nanorods and (G′) multilayer helical Si/Ni nanorod array, (H) side-coated nanorods and (H′) Ag side-coated L-shaped Si nanorods, (I) sandwiched nanorods and (I′) V/Mg sandwiched nanoblades for hydrogen storage, (J) core-shell nanorods and (J′) WO3/TiO2 quasi-core-shell nanorods for effective photocatalyst and (K) checker board nanorods and (K′) Ta/Si checker board nanorods.

However, it is shown that most nanostructures formed by GLAD on a flat substrate are different variations of nanocolumns. If different 2D or 3D ordered templates/masks can be used in the GLAD deposition system, the variety and complexity of nanostructures fabricated via GLAD can be greatly expanded [39], [40], [41]. Such a strategy really depends on how templates/masks are made through conventional or nonconventional nanolithography methods. Among various nanolithography methods, CL is a good candidate due to its low-cost and efficient fabrication process, as demonstrated in Figure 1. Thus, the combination of GLAD and CL could lead to a new powerful nanofabrication technique, and we call it the shadow sphere lithography (SSL). By replacing the flat substrates by 2D CCs or modified 2D CCs, the shadow effect due to GLAD becomes significantly more controllable. This can lead to sufficient degrees of freedom to design and fabricate a wide variety of nanostructures that incorporate complex geometry with small feature sizes and multiple materials and to generate a large variation of nanostructures for high-throughput screening that may yield unexpected properties. In the following sections, we present an overview of different strategies of utilizing SSL to design various nanostructures, including nanorods on nanospheres, patchy particles, nanotriangles, nanoring, and nanocrescents, or other complex metasurfaces using either CCs or modified CCs as mask or CL-derived masks. The material of nanostructures can be uniform or heterogeneous, including semiconductors, metal oxides, plasmonic noble metals, magnetic materials, and others as long as they can be evaporated via a PVD technique. The material, mask, substrate, and parameters of the typical nanostructures are summarized in Table 1. A brief update of plasmonic applications of the nanostructures with plasmonic noble metals (Au and Ag) is provided. Finally, the challenges and outlook of SSL are summarized to inspire more exciting developments in SSL and other applications.

Table 1:

Parameters of the typical nanostructures fabricated by SSL.

StrategyStructureExampleMaterialMask (template)Substrate (support)# of depositionPolar angle θ (°)Azimuthal angle ϕ (°)aRefs.
Nanorod deposited on CCsTilted nanorodsFigure 4APtSelf-maskPS sphere185[42]
Tilted nanorodsFigure 4A″CoPtSelf-maskPS sphere187[43]
Square spiralsFigure 4BSiSelf-maskPS sphereMultiple steps85With/without continuous rotation[36]
Multilayer spiral/column rodsFigure 4B″SiSelf-maskPS sphereMultiple steps85With/without continuous rotation[36]
Y-shaped nanorodsFigure 4CCuSelf-maskPS sphere184Continuous rotation[44]
Two component nanorodsFigure 4C″Si/TaSelf-maskSilica spheres184[45]
Patches deposited on CCSPatchy particlesFigure 6BAuSpherePS sphere1600/15/30/46[46]
Patchy particlesFigure 6CAuSpherePS sphere1800/18/29/40[46]
Patchy particlesFigure 6DAuSpherePS sphere1880/14/28/38[46]
Patchy particlesFigure 7AAgSpherePS sphere180Initial domain orientation=30

with sweep angle=60
[47]
Overlapped patchy particlesFigure 7BAu/AgSpherePS sphere28053/50[48]
Fan structureFigure 7CAgSpherePS sphere3860/120/240[49]
Helically stacked structureFigure 7DAg/SiO2SpherePS sphere7800/180/(–)90/180/(–)90/180/(–)90[50]
Swiss roll nanostructureFigure 7EAg/SiO2SpherePS sphere7700/180/(–)90/180/(–)90/180/(–)90[51]
Nanopatterns on substrates using CCs as maskTriangular NPsFigure 10ACrPS sphereSi110/20/26/40/28/2/16/2[52]
Touching triangular dimers/trimers/quadrumersFigure 10BAuPS sphereGlass2/3/410.8/16/16108/72/72[53]
Triangular NPsFigure 10CAgPS sphereGlass10°/10°/15°/20°Continuous rotation[54]
NanocrescentFigure 11BAuPS sphereCaF2140[55]
Stacked nanocrescentFigure 11CAuPS sphereGlass230φb=120[56]
NanocrescentFigure 11DAuPS sphereSi130Continuous rotation[57]
NanoantennasFigure 12Au/Ag/Pt/Cu/Ti/Cr/Co/Ge/NiPS sphereSiMultiple stepsMultiple anglesMultiple angles[58]
Nanostructures using CL-derived masksNanodiskFigure 14AAuHole maskSi10[59]
Heterodimer arrayFigure 14BAu/AgHole maskSi3N4235/32.5φ=180[60]
Gold trimerFigure 14CAuHole maskGlass310–15φ=120[61]
Silver tetramerFigure 14CAgHole maskGlass410–15φ=90[61]
Split-ringsFigure 14EAuHole maskSi122.5288[62]
Asymmetric double split-ringFigure 14FAuHole maskGlass2166/134[63]
Opposite spindlesFigure 15AAuHole maskGlass230φ=180[64]
NanocrescentsFigure 15BAuHole maskSiN150[65]
Nanoparticle-filled nanodishesFigure 15CAuHole maskSi1Continuous rotation[66]
  1. aThe reference to the azimuthal angle varied in different papers.

  2. bφ is the difference of the azimuthal angle between two deposition steps.

2 Nanostructures using CCs as a template

2.1 Nanostructures on CCs

2.1.1 Nanorods on CCs

The straightforward combination of GLAD and CL is to use CC (referring to close-packed monolayer except where noted) as a template to grow regular arrays of nanorods, as summarized in Figure 4. For example, tilted nanorods can be fabricated on nanospheres by OAD (Figure 4A) [42], [43]. By changing the speed and phase of the azimuthal rotation, the polar rotation, and the deposition rate, nanorods grown on nanospheres can be sculptured into a single layer or multilayer vertical, zigzag, or helical shaped nanocolumn arrays (Figure 4B) [67]. With a dual-source deposition (co-deposition), the Y-shaped nanorods and the two-layer Janus hetero-nanorods have been easily fabricated (Figure 4C) [44], [45]. During the fabrication, different source materials, such as Cu [45], Cr [68], Ta [69], Si [67], SiO2 [70], TiO2 [71], GeSbSe [72], TixSn1−xO2 [73], Sn1−xTixO2 [74], Pt [43], Ag [75], Ni [76], CoPt [42], and Fe3O4 [77], have been used. Although the basic principle and deposition procedure are the same for fabricating nanorods on flat substrates, there are also unique features for the nanorods grown on CC templates. The size and arrangement of the nanorods are significantly more uniform as compared with those fabricated on flat substrates. Especially, the size and arrangement of the nanorods can be controlled by the CC template [67]. This is explained by the shadowing effect of GLAD and the close-packed 2D CCs with a diameter of D as shown in Figure 5A. If the incident flux angle is fixed at θ and the shape of the nanosphere does not change during the deposition, due to the shadowing effect, the effective diameter d of the grown nanorod is given by [67]

Figure 4: Representative examples of nanorods on CCs.(A) Cross-section SEM images of tilted nanorods fabricated on spheres by a single OAD. (A′) Tilted Pt nanorods fabricated at θ=85° [42]. (A″) Tilted CoPt nanorods fabricated at θ=87° [43]. (B) Cross-section SEM images of nanorods fabricated by GLAD with continuous rotation [67]. (B′) Si spirals on a Si substrate covered with 830-nm colloid particles. (B″) Multilayer Si spiral/column rods on a Si substrate covered with 490-nm colloid particles. (C) Nanorods fabricated by dual-source co-deposition. (C′) SEM images of Y-shaped Cu nanorods [44]. (C″) Cross-sectional TEM micrograph from an array of Si-Ta two component nanorods grown on silica spheres [45].
Figure 4:

Representative examples of nanorods on CCs.

(A) Cross-section SEM images of tilted nanorods fabricated on spheres by a single OAD. (A′) Tilted Pt nanorods fabricated at θ=85° [42]. (A″) Tilted CoPt nanorods fabricated at θ=87° [43]. (B) Cross-section SEM images of nanorods fabricated by GLAD with continuous rotation [67]. (B′) Si spirals on a Si substrate covered with 830-nm colloid particles. (B″) Multilayer Si spiral/column rods on a Si substrate covered with 490-nm colloid particles. (C) Nanorods fabricated by dual-source co-deposition. (C′) SEM images of Y-shaped Cu nanorods [44]. (C″) Cross-sectional TEM micrograph from an array of Si-Ta two component nanorods grown on silica spheres [45].

Figure 5: Schematic configuration of the geometrical model and opposite GLAD.(A) The ratio of column diameter d to the colloid particle diameter D as a function of incident flux angle θ [67]. The insert shows the geometrical relationship for these three parameters. (B) Chamber schematic for simultaneous opposite GLAD, from Ref. [44].
Figure 5:

Schematic configuration of the geometrical model and opposite GLAD.

(A) The ratio of column diameter d to the colloid particle diameter D as a function of incident flux angle θ [67]. The insert shows the geometrical relationship for these three parameters. (B) Chamber schematic for simultaneous opposite GLAD, from Ref. [44].

(1)d=Dsin[θarcsin(12cosθ)].

As shown in Figure 5A, the ratio d/D decreases monotonically with the increase of the incident angle θ. Clearly, the diameter of the nanorod is determined by the diameter D of the colloid particle and the flux incident angle θ. Based on this geometrical model, Dolatshahi-Pirouz et al. have controlled the shape of the individual nanorod by varying θ and the amount of deposited materials [43]. In addition, substrate temperature Ts is another important factor to greatly impact the nanorods grown on nanospheres. As shown by Gall et al., when Ta nanorods have been grown on 2D SiO2 nanosphere CCs at θ=84° at different Ts (from 200 to 900°C) [69], [78], the nanorod morphologies have been significantly different. At low Ts (≤300°C), individual nanorods have grown on each sphere, branches of sub-rods are formed near the nanorod tips, and regular hexagonal arrays of nanorods are produced. At high Ts (≥500°C), branching starts to emerge during the nucleation stage, and the fraction of branched nanorods decreases with increasing Ts [79]. These results are attributed to the increased adatom mobility at elevated Ts, leading to stronger inter-rod competition and a larger average separation of growth mounds and, in turn, lower nucleation probabilities for sub-rods.

In addition to experimental studies, the formation mechanisms of Ag nanorods on bare and 2D CC substrates by GLAD have also been investigated via a 3D kinetic Monte Carlo simulation [75]. Jiang et al. have demonstrated three growth stages of the nanorods on a close-packed spherical seed layer. In the first stage, the Ag film is forced to grow on periodically aligned nanospheres instead of random distribution island. A periodic seed layer can enforce the formation of uniform nanorod arrays within a certain film thickness. In the second stage, as nanorod height increases, taller nanorods broaden and short nanorods are extinct. In the final stage, the lateral modulation effect of seed layers on the growth vanishes and random Ag nanorods form on the top, which is similar to the nanorod growth on the bare substrate.

Gall et al. have built a simultaneous opposite glancing-angle deposition (SO-GLAD) system (Figure 5B) to fabricate nanorod arrays on CCs with laterally stacked components (or Janus nanorods) [44]. During SO-GLAD, two different materials are deposited simultaneously from opposite directions both at oblique angles onto a 2D CC template. The shadowing effect caused by the nanospheres allows the selective deposition of only one material on one side of each nanosphere, leading to the formation of nanorods with two laterally separated components. In addition, other nanorod arrays, such as Cu Y-shaped nanorods (Figure 4C′) [45], Cr-Si two-component multi-stack nanorods [68], anisotropic Cu nanorods [80], Si-Ta two-component rods (Figure 4C″) [37], zigzag two-component nanosprings [37], vertical multilayer nanorods [37], and a checkerboard arranged nanorod array [37], have also been fabricated based on this SO-GLAD technique.

2.1.2 Patchy particles

Patchy particles, defined as particles with precisely controlled dual or multiple patches of varying surface and interaction properties, are a special type of particle used to assemble photonic crystals, fabricate targeted drug delivery system, and produce components for nanoelectronics [81], [82]. In SSL, upon deposition, the nanospheres in 2D or 3D CCs will receive evaporated vapors, and patches of different shapes can be formed on the nanospheres. In general, during SSL, the vapor incident angle θ determines the coverage of the patch on a nanosphere; i.e. the patch coverage decreases with increasing θ, and the azimuthal angle φ determines the shape of the patch. As shown in Figure 6A, through a single OAD, partial shells covering half spheres can be fabricated [83], [84], [85]. With increasing θ, the surface coverage (patchiness) of the shells can be reduced from 50% to 3.7% (Figure 6B–D) [46]. The shape of the patches is very sensitive to the azimuthal angle φ and is repeated within a period of ∆φ=60° for the hexagonally arranged CCs. The change in incident and azimuthal angle leads to a large variety of patchy particles (Figure 6B–D) [86], [87], [88], [89], [90]. Thus, the SSL is a powerful technique to prepare patchy particles and can control the shape and surface coverage. When the azimuthal angle is sequentially or continuously changed during patchy particle preparation, patchy particles with complicated shapes can be produced. Zhao et al. have used a swinging GLAD strategy to produce patchy particles with uniform size, shape, and plasmonic response (Figure 7A) [47]. Swinging the colloid substrate azimuthally causes the domain orientation to rotate away from the initial orientation (φ0) through a prescribed angle range (Δφ) during the deposition. Thus, during a deposition, the symmetry inherent in the close-packed colloidal monolayer can be utilized to produce patchy particles with the desired uniformity without the complications of a prefabricated template or other special colloid layer preparation steps.

Figure 6: Patchy particles fabricated by single deposition.(A) Schematic of metal evaporation geometry and monolayer orientation with respect to the incident vapor beam. (a) x-z cross sectional view indicating angle of incidence θ. x-y cross-section indicating monolayer orientation (b) at φ=0° and (c) at φ=30°. Experimental gold patches on 2.4 μm PS particles as a function of the angle of incidence (θ) and monolayer orientation (φ). The images show the top view (x-y view) of the monolayer. (B) Patches obtained at an angle of incidence of θ=60° and monolayer orientations of φ=0, 15, 30, and 46° (from left to right). (C) Patches obtained at an angle of incidence of θ=80° and monolayer orientations of φ=0, 18, 29, and 40° (from left to right). (D) Patches obtained at an angle of incidence θ=88° and monolayer orientations of φ=0, 14, 28, and 38° (from left to right). Scale bars in experimental images correspond to 2 μm, adapted from Ref. [46].
Figure 6:

Patchy particles fabricated by single deposition.

(A) Schematic of metal evaporation geometry and monolayer orientation with respect to the incident vapor beam. (a) x-z cross sectional view indicating angle of incidence θ. x-y cross-section indicating monolayer orientation (b) at φ=0° and (c) at φ=30°. Experimental gold patches on 2.4 μm PS particles as a function of the angle of incidence (θ) and monolayer orientation (φ). The images show the top view (x-y view) of the monolayer. (B) Patches obtained at an angle of incidence of θ=60° and monolayer orientations of φ=0, 15, 30, and 46° (from left to right). (C) Patches obtained at an angle of incidence of θ=80° and monolayer orientations of φ=0, 18, 29, and 40° (from left to right). (D) Patches obtained at an angle of incidence θ=88° and monolayer orientations of φ=0, 14, 28, and 38° (from left to right). Scale bars in experimental images correspond to 2 μm, adapted from Ref. [46].

Figure 7: Patchy particles fabricated by multiple deposition.(A) SEM image of the Ag deposited on PS bead monolayers by swinging deposition with sweep angles (Δφ=60°) and initial domain orientations (φ0=30°) [47]. (B) SEM image of the patchy particles fabricated by two subsequent OADs [48]. The inset shows the patch geometry obtained with the mathematical model. (C) SEM image showing the fan structure on spheres fabricated by three OADs [49]. (D) TEM image of helically stacked structure fabricated by seven subsequent OADs [50]. White and blue dashed lines are artificially added to outline different Ag and SiO2 layers, respectively. Inset is a 3D schematic expected from the deposition sequence. (E) Cross-sectional SEM image of Swiss roll nanostructures. The inset shows the schematic of the corresponding structures [51].
Figure 7:

Patchy particles fabricated by multiple deposition.

(A) SEM image of the Ag deposited on PS bead monolayers by swinging deposition with sweep angles (Δφ=60°) and initial domain orientations (φ0=30°) [47]. (B) SEM image of the patchy particles fabricated by two subsequent OADs [48]. The inset shows the patch geometry obtained with the mathematical model. (C) SEM image showing the fan structure on spheres fabricated by three OADs [49]. (D) TEM image of helically stacked structure fabricated by seven subsequent OADs [50]. White and blue dashed lines are artificially added to outline different Ag and SiO2 layers, respectively. Inset is a 3D schematic expected from the deposition sequence. (E) Cross-sectional SEM image of Swiss roll nanostructures. The inset shows the schematic of the corresponding structures [51].

Through two subsequent OAD depositions with different φ, the shape of the patches on the patchy particles has been further widened [91], [92], [93]. For example, Lei et al. used stepwise OAD of two metal layers on a centimeter-scale self-assembled dielectric microsphere array to produce large-area stacked-patch plasmonic chiral metamaterials (SPPCMs) [94]. Kretzschmar et al. used two sequential vapor depositions with ∆φ=60° to fabricate overlapping patches on the same hemisphere of the particle, as shown in Figure 7B [48]. In such a way, L-shaped Ag nanostructures composed of two rectangular shaped stripes with different thicknesses are prepared by controlling substrate azimuth angle and deposition time [95]. These ordered patchy particles show strong optical chirality in visible and near-infrared (IR) regions. If three subsequent OADs are performed at three different φ, three patches will be deposited on the nanospheres to form fan-shaped structures. For example, through three OADs with θ=86° and ∆φ=120°, three identical patches are overlapped onto hexagonal close-packed CCs, forming fan structures (Figure 7C) [49], [96], [97]], [98]], [99]. The Ag fan structures on 500-nm-diameter polystyrene (PS) spheres demonstrated a strong, visible, light circular dichroism response. In addition, during multiple OAD processes of patch formation, if the deposition materials can be alternated, the chemical ordering of the resulting patch can be modulated, and even more complicated patchy particles can be designed. By seven OADs and controlling the azimuthal rotations of the 2D CC substrate, Ag and SiO2 layers can be helically stacked in left-handed and right-handed fashions to form a continuous helically stacked structure (Figure 7D) [50]. The wavelength range of the circular dichroism response of these helical structures can be tuned by the diameter of the nanospheres. Furthermore, more complicated 3D patchy structures, such as conical Swiss roll structures on nanospheres, can also be fabricated by combining swing GLAD and multiple alternating depositions of Ag and SiO2 (Figure 7E) [51]. The resulting chiral plasmonic structures show a broadband circular dichroism response.

If the nanosphere template is manipulated or altered during the multiple depositions, additional types of patchy particles can be fabricated. One method is to use polydimethylsiloxane (PDMS) stamping to inverse the CC monolayers for two subsequent depositions onto the upper and lower half surfaces of the nanospheres (Figure 8A) [100]. Hollow spheres with windows of controlled size and number have been fabricated, which can be used as nano-carrier, for drug transport, and as nano-research containers. In addition, by changing the lattice structure of the CCs, the patches on the nanospheres can be altered. For example, linear chains of nanospheres can be self-assembled by the V-shaped or square-shaped grooves in a Si wafer [101], [103]. The patch shape and size on those nanospheres can be controlled by the vapor incident angle θ, nanosphere diameter D, the dimensions of the grooves, the vapor flux azimuthal direction with respect to the groove direction (α), and the number of deposition and evaporated materials (Figure 8B). The dimensions of the grooves add another degree of freedom to design a large variety of asymmetric patchy particles. The nanosphere monolayer can also be replaced by two-layered or multilayered CCs to form different patches on lower layered nanospheres. Zhang et al. used this strategy to produce diverse Au nanosized dots on the lower layer spheres by using the upper layer nanospheres as masks for deposition (Figure 8C) [102]. Such a strategy uses the top nanosphere monolayer(s) as a mask and the lower nanosphere monolayer(s) as a template.

Figure 8: Patchy particles fabricated by modifying the monolayers.(A) Schematic of the fabricating process of two-pole patchy particles [100]. (B) Schematic of metal vapor flow in template-assisted patchy particle fabrication using the glancing angle deposition method with a template at various angles [101]. (C) Schematic depictions of angle-resolved nanoembossment of Au patterns on microspheres by using their crystals as masks during Au vapor deposition [102].
Figure 8:

Patchy particles fabricated by modifying the monolayers.

(A) Schematic of the fabricating process of two-pole patchy particles [100]. (B) Schematic of metal vapor flow in template-assisted patchy particle fabrication using the glancing angle deposition method with a template at various angles [101]. (C) Schematic depictions of angle-resolved nanoembossment of Au patterns on microspheres by using their crystals as masks during Au vapor deposition [102].

The main advantage of fabricating patchy particles by SSL technique is that it can accurately position patches onto the surfaces of nanospheres controlled by simple geometric rules and the patch materials can be widely altered as long as the material can be evaporated. However, there are still many challenges for using SSL to fabricate high-quality patchy particles. The biggest challenge is the domain orientation and alignment of the nanosphere monolayer. The nanosphere monolayer used in most studies has a hexagonal lattice, which possesses a sixfold symmetry. The patch shape depends on the domain orientation, leading to diverse nanostructures. Figure 9 shows a true-color optical image captured for a stacked-patch nanosphere monolayer under unpolarized white light illumination [94]. The micro-sized regions of different reflection colors represent microscale domains with different lattice orientations and geometric morphologies, which means that the nanostructures on the nanospheres are different. According to Figure 6, as long as one performs a single OAD, since different domains have different φ with respect to the azimuthal vapor incident direction, the shape of the patches on nanospheres in different domains will be different. Especially if the θ is closer to 90°, the significance of the difference in patch shape will become greater. One way to resolve this issue is to make large-area, single-domain monolayers and pre-align the domain orientation with respect to the azimuthal direction of the incident vapor flux. However, large-area, single-domain monolayers are not easy to achieve experimentally and the alignment needs extra work. There are also alternative ways that have been developed to reduce the influence of domain orientation. A template can be used to assist the self-assembly of nanosphere monolayers so that small patches of monolayers with the same orientation can be formed, thus leading to the same shadow effect [101], [103]. In addition, if the substrate with 2D CCs is swung (rotated azimuthally back and forth) azimuthally during the deposition, symmetry inherent in the close-packed nanosphere monolayer can be utilized to produce patchy particles with uniform size, shape, and plasmonic response (Figure 7D) [47]. Finally, if subsequent multiple depositions can cover a 360° azimuthal rotation, since different shaped patches can be produced at different depositions at various monolayer domains, the resulting patchy particles inherently have the same morphology and thus demonstrate the same property. For example, the conical Swiss roll structures (Figure 7E) have properties independent of the orientations of monolayer domains, which leads to a very uniform array [51]. Recently, with multiple depositions of Ag and SiO2 at different relative φ, a uniform large area broadband optical absorber has been designed based on the plasmonic response of different shapes of Ag patches on nanospheres [104]. While the presence of multiple domains complicates the fabrication of large-area homogeneous patchy particles, it is useful in high-throughput screening of patchy particles when a diverse set of patch features on a single sample is required. For instance, the multiple domains created on a single SSL-fabricated surface with multiple domains can be characterized locally to establish a patch shape-property relationship via a one-step deposition so that a desired patch shape can be easily identified for specific property optimization.

Figure 9: Top-view reflection image of a stacked-patch nanosphere monolayer illuminated under an unpolarized white light [94].
Figure 9:

Top-view reflection image of a stacked-patch nanosphere monolayer illuminated under an unpolarized white light [94].

2.2 Nanopatterns on substrates using CCs as mask

2.2.1 Nanotriangles

Using CCs as a shadow mask, different nanopatterns can be designed and produced on substrates through direct deposition. Like the formation of different patchy particles, multiple depositions or multilayer deposition can also be used to design complex nanopatterns on the substrates. One typical structure that forms through such a strategy is a nanotriangle array, which can be fabricated by direct vertical deposition through close-packed nanosphere monolayers [3]. The shape and size of the nanotriangles strongly depend on vapor incident angle θ and azimuthal angle φ (with respect to the orientation of monolayer domains) due to a simple geometric shadowing effect [52]. By a single OAD deposition, elongated triangles are formed (Figure 10A). Such a deposition not only can change the size of the resulting nanotriangles but also introduce anisotropy in produced nanostructures, leading to anisotropic physical and chemical properties [52]. By two subsequent depositions with different relative θ and φ, nanotriangles with overlap, nanocontacts, and nanogaps were fabricated [105]. When the θ was the same, but with two opposite deposition directions (+θ/−θ), bow-tie-like triangles with small gaps have been fabricated [106]. By three subsequent depositions with different relative θ, chains of nanotriangles are formed [105]. By four subsequent depositions and carefully adjusting of φ and the deposition thickness, complicated structures such as chiral plasmonic oligomers composed of four achiral triangles have been successfully produced and demonstrated [107]. Examples of the triangular dimers, trimers, and quadrumers can be seen in Figure 10B [53]. In addition, the triangular shape can be modulated when the substrate is continuously rotated azimuthally at different vapor incident angle θ during deposition. A transition from triangles to polygonal shapes then to triangle network-like structures has been observed when θ changes from 0° to 20° during the deposition of Ag by continuously rotating the substrates (Figure 10C) [54], [108], [109]. For the triangle network-like structures with θ=20°, if a second deposition was implemented, nanoparticle-in-ring arrays would be fabricated [110].

Figure 10: SEM images of the nanotriangles fabricated by SSL.(A) SEM images of triangular NPs and images with simulated geometry superimposed, respectively. (A1, A2) θ=10°, φ=28°, (B1, B2) θ=20°, φ=2°, (C1, C2) θ=26°, φ=16°, and (D1, D2) θ=40°, φ=2° [52]. All samples are Cr deposited onto Si (111) substrates. (B) SEM images of the touching triangular dimers, trimers, and quadrumers fabricated by two, three, and four OADs, respectively. The evaporation and rotation angles θ and φ are depicted below each image [53]. (C) SEM images of the triangular NPs fabricated by GLAD with continuous rotation at θ=0°, 10°, 15°, and 20°, respectively [54].
Figure 10:

SEM images of the nanotriangles fabricated by SSL.

(A) SEM images of triangular NPs and images with simulated geometry superimposed, respectively. (A1, A2) θ=10°, φ=28°, (B1, B2) θ=20°, φ=2°, (C1, C2) θ=26°, φ=16°, and (D1, D2) θ=40°, φ=2° [52]. All samples are Cr deposited onto Si (111) substrates. (B) SEM images of the touching triangular dimers, trimers, and quadrumers fabricated by two, three, and four OADs, respectively. The evaporation and rotation angles θ and φ are depicted below each image [53]. (C) SEM images of the triangular NPs fabricated by GLAD with continuous rotation at θ=0°, 10°, 15°, and 20°, respectively [54].

2.2.2 Nanorings and nanocrescents

Another typical structure fabricated on substrates by SSL is nanocrescent. Nanocrescent is a symmetry-broken structure and has shown a range of interesting phenomena due to the reduction of symmetry. Figure 11A shows a typical process to fabricate nanocrescents based on SSL [55], [112], [114], [115]], [116]. First, randomly dispersed or ordered nanospheres are formed on substrates to act as template [or for close-packed 2D CCs, the nanospheres are usually etched via a reactive ion etching (RIE) process to introduce the separations between nanospheres]. Then, metal is deposited on the nanosphere-covered substrate with an adjusted incident angle θ. After deposition, ion beam milling is performed vertically to remove the metal layer that is not masked by the nanospheres. Here, the nanospheres are acting as masks that prevent anything inside the projected areas being etched. Finally, the nanospheres are removed and the metal layers left under the projected areas of the nanospheres form nanocrescents on substrates (Figure 11B). In the process, the size and shape of nanocrescents can be controlled by the size of nanospheres, deposited thickness of metal, incident angle θ, as well as the azimuthal rotation. In fact, the desired nanocrescent shapes can be modeled by a 3D profile simulation [56], which can provide the guidance in fabrication of nanocrescents.

Figure 11: Nanocrescents fabricated by SSL.(A) Scheme for the typical fabrication process of crescents [111]. (B) SEM image of a nanocrescent fabricated by a single OAD. The red and blue arrows show the long and short axis, respectively [112]. (C) SEM image of stacked nanocrescent fabricated by two OADs [111]. (D) SEM image of a nanocrescent fabricated by continuous rotation [113].
Figure 11:

Nanocrescents fabricated by SSL.

(A) Scheme for the typical fabrication process of crescents [111]. (B) SEM image of a nanocrescent fabricated by a single OAD. The red and blue arrows show the long and short axis, respectively [112]. (C) SEM image of stacked nanocrescent fabricated by two OADs [111]. (D) SEM image of a nanocrescent fabricated by continuous rotation [113].

The typical process has been extended to fabricate various nanocrescent-derived nanostructures. When the substrate is rotated with respect to the surface normal by an azimuthal angle φ (φ<180°) with subsequent multi-depositions, stacked nanocrescents from an open crescent to a fully connected ring are obtained (Figure 11C) [111]. When φ=180°, opposing nanocrescent dimers with sub-10-nm gap between the tips of the final crescent can be fabricated [117]. If a dielectric layer is deposited prior to the second deposition, one can obtain stacked double-crescent arrays [118]. In such a way, 3D densely packed arrays of nanocrescents have been fabricated by applying the above process twice with a SiO2 space layer at two different φ [119]. In addition to changing θ and φ, the etching process can also be varied; e.g. perpendicular metal etching can be expanded to tilted etching to provide more opportunities to produce nanocrescents in various shapes [120]. The final profiles after perpendicular or tilted etching can also be calculated as references for the fabrications [57].

Sutherland et al. have also developed an alternative method to fabricate nanocrescents by continuously rotating the substrate, eliminating the metal milling step [113]. The concept is to use the nanospheres to deposit a structured sacrificial layer (e.g. SiO2) around the nanoparticle to define an adhesive region (e.g. Ti) on the surface. Then, OAD is used to deposit a metal nanostructure onto the adhesive domain. The shape is determined by the gap between the nanospheres and the sacrificial layer. Nanostructures of nanorings or nanocrescents are formed after the removal of the nanospheres and the sacrificial material (Figure 11D). The process step of changing the deposition angle of the sacrificial layer gives the opportunity to systematically control the shape of the final pattern. In addition, other processes have also been developed to fabricate nanoscrescents based on SSL. For example, SSL and parallel imprint have been combined to fabricate long-range ordered crescent structure arrays [121]. First, silicon crescent nanohole arrays have been fabricated based on SSL, whose complementary structure can be replicated by PDMS. After metal deposition on the PDMS model, nanocrescent particle arrays can be duplicated onto various substrates by parallel imprinting. The parallel printing method greatly improves the reproducibility of the fabrication process and also reduces the cost. In addition, by incorporating a sacrificial copper layer in SSL, aluminum plasmonic nanocrescent antennas have been fabricated over large substrate areas [122]. The addition of the copper mask eliminates the argon ion milling step, which is challenging to implement for aluminum nanostructures because of the robust native oxide layer.

In addition to SSL, other templates can also be used to create nanoring and nanocrescent structures. Jiang et al. have reported a scalable bottom-up approach for fabricating periodic arrays of metal nanorings and nanocrescents [123]. First, nanoring-shaped trenches between templated polymer posts and sacrificial nanoholes have been created. Then directional deposition of metals in the trenches, followed by liftoff of the polymer posts and the sacrificial nanoholes, results in ordered metal nanorings. By simply controlling the vapor deposition angle θ of the metal beams, continuous geometric transition from concentric nanorings to eccentric nanorings and even to nanocrescents has been achieved. Giersig et al. have fabricated nanorings and nanocrescents based on the deposition through the circular aperture formed using the close-packed CCs [124], [125]. In addition to the nanocrescent particles, crescent nanohole arrays have also been realized by varying the SSL fabrication processes [115], [121].

2.2.3 Other nanostructures fabricated on substrates

Besides the two representative structures, the nanotriangles and nanocrescents, other complex nanopatterns can also be generated by controlling the parameters in the SSL process [126], [127]. Kostinski et al. considered a set of possible 2D patterns accessible via periodic SSL [128]. A topological characterization based on overlap of shadows has been proposed and a variety of patterns is reduced to 4+1 topologically distinct groups. The authors have developed phase diagrams linking the five topological groups from any lattice geometry to experimental rotation and tilt angles. The phase diagrams are readily constructed for a variety of configurations by plotting contours of zero-valued Cayley-Menger determinants containing the desired sphere radii and lattice vectors and can guide CL experiments.

Experimentally, by using O2-plasma etched bilayers of hexagonally packed spheres as templates, Wang et al. have fabricated highly ordered binary arrays of gold nanoparticles with varied shapes, such as a shuttlecock-like shape composed of a small crescent-shaped nanoparticle and a big fan-shaped one [129], and two different triangular nanoparticles interspaced by a tiny gap [130]. The size and shape of both small and big nanoparticles obtained are manipulated by the plasma etching time and the incidence angle θ of Au vapor. The subsequent thermal annealing leads to binary arrays of round Au nanoparticles with a rather narrow distribution in terms of size and shape. Based on a similar strategy, they have successfully demonstrated a stepwise SSL process to grow highly ordered multiplex quasi-3D grids of metallic one-dimensional nanostructures, e.g. nanowires and nanorods [131]. In addition, elliptical holes can be created by simply tilting the substrate during the metal deposition [10], [132]. The ellipticity of the holes depends on the size of the nanospheres and the vapor incident angle θ; smaller nanospheres with a larger θ produce ellipses with higher aspect ratios.

Recently, Whitesides’s group presented a detailed and systematic work on producing an extensive variety of complex 2D metasurfaces based on SSL (Figure 12) [58]. They showed that rationalizing the vast and almost entirely unexplored parameter space of shadow-derived shapes enables a substantial expansion of the spectrum of structures that can be generated by projection through 2D CCs. Their approach uses custom-designed software to predict the shadows that guide multiangled deposition of one or multiple materials through a plasma-etched 2D CC. Through a comprehensive exploration of the space of available shadow-derived shapes, they have found that the patterns that result from different combinations of θ and φ generally belong to a set of five classes: (1) interconnected lines, (2) asymmetric bars, (3) symmetric bars, (4) triangular islands, or (5) an interconnected, honeycomb-like lattice. Each feature can be duplicated at φ=60° intervals due to the sixfold (C6) rotation symmetry of the hexagonal lattice. Varying the azimuthal angle φ produces continuous transitions between the different classes of shadows and offers many intermediate positions and shapes; varying the polar (incident) angle θ or the gap between the spheres controls the position, length, and width of each feature.

Figure 12: SEM images of nanoantennas designed and fabricated by SSL.Patterns composed of two to three (A), four (B), and five to six (C) different angles of deposition. (D) Subset of possible unit cells categorized according to optical function as designed (left) and fabricated (right), adapted from Ref. [58].
Figure 12:

SEM images of nanoantennas designed and fabricated by SSL.

Patterns composed of two to three (A), four (B), and five to six (C) different angles of deposition. (D) Subset of possible unit cells categorized according to optical function as designed (left) and fabricated (right), adapted from Ref. [58].

Although numerous structures have been fabricated on substrates based on SSL, challenges still exist. The major obstacle encountered when using SSL to fabricate patterns on substrates is the inability to control mask registry φ, i.e. the orientation of the 2D CC domains. Because 2D CC growth results in multiple hexagonally close-packed domains with random orientations (Figure 9), the shape, size, and spacing of the resulting structures could vary in different domains since the detailed shadowing effect is different for different deposition configurations. While the presence of multiple domains introduces the inhomogeneity in the fabrication of large-area homogeneous structures, it could be useful in high-throughput screening a diverse set of nanostructures. For instance, if a local property characterization technique can be used and the location of a substrate can be registered, then a quick structure-property relationship can be established using the structures fabricated via a multi-domain 2D CC. If the resulting CC is a single domain, the alignment of the domain orientation with respect to vapor deposition direction φ becomes important. However, for small-diameter nanospheres such as diameter <500 nm, it is a big technical challenge to achieve such an alignment. Moreover, for multiple depositions, in most simulation programs, only the shadowing effect due to bare nanospheres is considered, and the additional shadowing effect due to the deposited layers on top of the nanospheres or side walls of previous deposition is usually not considered. The additional shadowing effect could produce structures with distorted shape/pattern compared to those predicted by the simulations. To minimize the effect, sometimes, multiple depositions with smaller thickness at different θ and φ conditions can be used [49].

3 Nanostructures using CL-derived masks

3.1 Using a CL-derived hole as mask

In SSL, not only the configuration of the deposition and the locations of materials deposited on CCs are important, but also the masks or templates used are critical. As discussed above, the nanosphere monolayers or multilayers (3D CCs) can serve as templates or masks to design various nanostructures. They can also be used as templates or masks to design other templates or masks to fabricate more complicated nanostructures using the shadowing effect. For example, the nanohole arrays fabricated by CL constitute one of the most used masks, and nanofabrication using holes as mask can be called as hole-mask CL (HCL) [62], [133], [134]. The basic fabrication steps in HCL are shown in Figure 13 [134]: (1) A sacrificial thin film [e.g. polymer such as poly (methyl methacrylate) (PMMA) or dielectric such as SiO2] is spin-coated or deposited onto a flat surface (Figure 13A). (2) Nanoholes are fabricated by a typical CL process [10] on the sacrificial layer (Figure 13B and C). It should be noted that the nanoholes are usually dispersed. (3) Reactive plasma etching is applied to selectively and isotropically remove the sacrificial polymer or dielectric layer exposed underneath the nanoholes (Figure 13D). The etching selectivity of the double-layer film (the mask material and the sacrificial film) leads to a partial undercut of the sacrificial film. (4) GLADs with varied configuration parameters are carried out (Figure 13E). (5) The sacrificial layer and the mask are removed to leave the nanostructures on the substrates (Figure 13F). The essential feature of HCL is the presence of a sacrificial layer combined with a thin film mask with nanoholes. The hole mask is used in shadowing evaporation and/or etching to define a nanopattern. The sacrificial layer is used not only to define the projected pattern and boundary but also to remove the hole mask in the last processing step. By using the process with only one simple OAD, nanodisks (Figure 14A) [133] and tilted pillars [138] have been fabricated. If two OADs have been performed at opposite azimuthal directions, nanodisk dimers with overlaps or gaps can be designed (Figure 14B) [60], [61], [139]. Heterotrimers (Figure 14C) and heterotetramers (Figure 14D) composed of closely spaced silver and/or gold nanodisks of different heights have also been prepared through HCL by three and four OAD depositions, respectively [136]. The parameters of the disk can be well controlled by the hole size and the OAD configuration parameters θ and φ. If the hole mask can be rotated in both the polar and azimuthal directions, ring or split-ring structures can be produced (Figure 14E) [134], [140], [141]. Clearly with the HCL, the materials are not limited to evaporation but they can also be deposited via electroplating or other methods [63]. In addition, heterostructures such as Ag/Au [139], Au/Co [141], and Pt/Ru [60] hetero-nanodisk dimers can be produced with multilayer depositions.

Figure 13: Fabrication scheme.(A) Deposition of sacrificial PMMA layer; (B) deposition of polystyrene spheres in an arbitrarily distributed arrangement; (C) evaporation of oxygen plasma resistant Au-mask; (D) removing polystyrene spheres with ultrasonic bath and subsequent isotropic dry etching leads to extended holes in PMMA layer; (E) evaporation of the structures by tilted angle evaporation (polar rotation angle θ, azimuthal tilt angle φ); (F) lift-off of the sacrificial PMMA layer, adapted from Ref. [134].
Figure 13:

Fabrication scheme.

(A) Deposition of sacrificial PMMA layer; (B) deposition of polystyrene spheres in an arbitrarily distributed arrangement; (C) evaporation of oxygen plasma resistant Au-mask; (D) removing polystyrene spheres with ultrasonic bath and subsequent isotropic dry etching leads to extended holes in PMMA layer; (E) evaporation of the structures by tilted angle evaporation (polar rotation angle θ, azimuthal tilt angle φ); (F) lift-off of the sacrificial PMMA layer, adapted from Ref. [134].

Figure 14: Nanostructures fabricated by using holes as the mask.(A) SEM image of nanodisks fabricated by single OAD based on HCL [135]. (B) SEM image of an Au-Ag heterodimer array fabricated by two OADs [61]. Inset shows elemental mapping of one dimer, with red representing Au and green representing Ag. SEM images of a (C) gold trimer and (D) silver tetramer by three and four depositions, respectively [136]. (E) SEM image of split-rings fabricated by continuous rotation [134]. (F) SEM image of asymmetric double split-ring fabrication by two depositions with continuous rotation [137].
Figure 14:

Nanostructures fabricated by using holes as the mask.

(A) SEM image of nanodisks fabricated by single OAD based on HCL [135]. (B) SEM image of an Au-Ag heterodimer array fabricated by two OADs [61]. Inset shows elemental mapping of one dimer, with red representing Au and green representing Ag. SEM images of a (C) gold trimer and (D) silver tetramer by three and four depositions, respectively [136]. (E) SEM image of split-rings fabricated by continuous rotation [134]. (F) SEM image of asymmetric double split-ring fabrication by two depositions with continuous rotation [137].

The shape and related geometric parameters of the nanostructures fabricated by HCL can be calculated based on the simple geometric shadow model. The control of the azimuthal angle φ and the incident angle θ in combination with the sacrificial layer thickness h0 determines the outer radius of the ring (r) and the gap size s (opening of the ring), with the parameters defined in Figure 13. The radius r is given by

(2)r=h0tanθ,

and the gap size s is determined by

(3)s=2πr360°φ360°=2πh0tanθ360°φ360°.

The width w is determined by the mask-hole diameter. The structure height h can be calculated by the equation

(4)h=h145°wφh0tanθ.

Here, h1 is the evaporated material thickness. By applying two rotated depositions based on this geometric model, asymmetric double split-rings have been fabricated (Figure 14F) [137]. By varying the polar rotating speed of the samples during gold evaporation, Giessen et al. have fabricated spiral-type ramp split-ring structures [142]. Also, a series of split-ring nanoantenna placed directly on top of a magneto-plasmonic ring element have also been produced using a similar strategy [143]. Other complex plasmonic nanostructures (e.g. split-ring with different orientations) can be fabricated using multiple repetitions of HCL [59].

The typical fabrication process shown in Figure 13 can be modified by dynamically changing the hole size or sacrificial layer thickness during the deposition. For example, the diameter of the hole can be shrunk by evaporating Ti at a high θ angle (≥70°) with continuous azimuthal rotation [135]. Due to the shadowing effect around the hole, Ti will only be deposited around the side wall of the holes, and such a lateral growth continuously makes the hole diameter smaller. Such a strategy could ensure a continuum modification of the mask during the deposition. In addition, Sutherland et al. have reported an extension of HCL to fabricate gradients of nanostructures with systematically varied nanostructure features to screen the properties of plasmonic arrays by modulating the polymer sacrificial layer thickness across the sample prior to creation of the hole mask [144]. A gradient in polymer film thickness can be translated into a gradient of the nanostructure feature size by using OAD [see Equations (2) and (3)]. Käll et al. introduced another SSL methodology – evaporation on a partially exposed rotating substrate with hole mask – to generate continuous and precise gradients in nanoscale structure parameters, including vertical thickness, lateral positioning, and orientation, based on controlling the relative movement between a supporting substrate and a partial shutter during material evaporation [64]. Nanodisks with continuously varying thicknesses, sandwich structures with a gradient in the spacer thickness, nanoparticle dimers with continuously increasing gap distances, and split-ring nanoparticles with continuously changing orientations have been fabricated based on this strategy. In addition, SSL using hole as mask can also be used to fabricate an array of plasmonic opposite metal spindles in nanowells (Figure 15A) [65], nanopores containing nanocrescent particles (Figure 15B) [66], and 3D nanodishes that serve as small plasmonic containers to host guest nanoparticles (Figure 15C) [145].

Figure 15: Nanostructures fabricated by using the CL-derived mask.SEM images of (A) opposite metal spindles in nanowells [65], (B) nanopores containing nanocrescent particles [66], and (C) nanoparticle-filled nanodishes [145].
Figure 15:

Nanostructures fabricated by using the CL-derived mask.

SEM images of (A) opposite metal spindles in nanowells [65], (B) nanopores containing nanocrescent particles [66], and (C) nanoparticle-filled nanodishes [145].

3.2 Using other CL-derived masks

In addition to using holes as templates or masks in SSL, other different kinds of templates and masks derived from nanosphere monolayers have also been investigated. By using inverted hemispheres as a template for SSL, Jin et al. have fabricated a periodic array of 3D crescent-like holes [146]. First, A PS film perforated with a hexagonally non-close-packed dimple array is generated by heating a silica nanosphere monolayer on top of the PS film. Then, an OAD is employed to deposit a layer of silver on the patterned PS film to form a 3D crescent-like hole array. The silver covered area within each individual dimple is controlled by the vapor incident angle θ. If a second OAD has been employed from the opposite azimuthal direction, an array of non-planar nano-arc-gaps will be formed [147], [148]. The gap width of the nano-arc-gaps can be tuned via the angle θ. If θ becomes bigger, the width of the gap can be larger.

Using nanodisks (e.g. Figure 14A) or the triangular shaped nanodots (e.g. Figure 10A) formed via SSL as templates, ordered arrays of Si nanorods and nanospirals have been produced [149], [150], [151], [152]. Prepatterned honeycomb-like and hexagonal closed-packed arranged Au nanodots have been obtained by CL using self-assembled mono- and double layers of PS nanospheres, and Si nanostructures have been deposited onto these nanopatterns via GLAD. The morphology of the Si nanostructures deposited on the nanodots depends strongly on the deposition angle θ [151]. A continuous film with honeycomb-like arranged hillocks has been produced at θ=0°. With the increased θ (≤70°), the nanostructure changes to dense columnar films with a mesh of hexagonally arranged pores. Finally, when θ becomes larger than 85°, separated rod-like structures with a triangular cross-section are formed. The nanodots or nanodisks can also serve as an etching mask (or nanostructured seed layer) for GLAD via an RIE or a chemical etching process [153]. The nanodots or nanodisks formed after RIE can have a much higher aspect ratio (i.e. depth) compared to similar structures formed via direct SSL, so that when a Si GLAD is performed on these templates, a better self-alignment effect can be achieved and less growth of Si nanostructures in between the nanodots could occur. Similarly, inverted pyramid hole patterns have been created as templates for GLAD via anisotropic etching of Si (001) substrates, leading to the formation of a network of nanochannels surrounded by arrays of vertically aligned nanorods after a subsequent GLAD [150], [154]. Here, nano-triangular-shaped Cr masks have been fabricated onto the Si (100) substrates via a conventional CL, and inverted-pyramidal pit arrays are formed using an anisotropic chemical etching process. Many other nanostructures have been also designed using similar strategies. For example, nanocones and nanopillars formed via an RIE process by CL have been used as mask in SSL to fabricate half-cone shells [155], [156] and Janus nanopillars [157], [158], [159], respectively. All these results demonstrate that substrate patterning by CL is very useful in SSL to design more complicated nanostructures.

4 Plasmonic properties of the nanostructures fabricated by SSL

The nanostructures fabricated by SSL have been widely exploited in various fields, including nanomotors based on Janus nanorods on spheres [70], [71], self-assembly based on patchy spheres [81], [160], magneto-optical activity [143], anisotropic wettability [156], [157], cell growth behaviors on nanotopographies [144], electrocatalytic properties [60], [144], plasmonics, etc. Among these applications, plasmonics is one of the most widely investigated fields due to the fact that numerous noble metal nanostructures can be easily fabricated by SSL. In the following, we will mainly show some examples of the applications of the above structures (as classified in patchy particles, nanotriangles, nanocrescents, and structures fabricated by HCL) in plasmonics.

Patchy particles have been designed to show strong chiral optical properties by controlling the patch shape and material deposited on the spheres, which can be used for chiral optics and sensors. Fan-shaped Ag nanostructures (Figure 7C) have been designed on nanospheres by three subsequent OADs with a relative difference of 120° in azimuthal angle based on the shape of the patch [49]. The relative rotation direction during OADs breaks the racemic symmetry, leading to a different distribution of the amount of Ag on three patches of the fan shape, which results in the chirality of the optical response. However, structural handedness varies locally due to the sixfold symmetry of the lattice and different orientations of the monolayer domains (see discussions in Section 2.2.3). Thus it is very hard to realize the same handedness within a large substrate area [97]. To address this issue, other improved methods have been developed. By controlling the azimuthal rotation of substrates as well as the deposition materials, Ag and SiO2 layers have been alternatively deposited by OADs on to nanospheres and helically stacked in left-handed and right-handed fashions to form helices of Ag patches (Figure 7D), called helically stacked plasmonic layers [50]. The pitch and diameter of the stacked helices are determined by the diameter of nanospheres, the vapor incident angle θ, and the layer thicknesses of Ag and SiO2. These plasmonic helices exhibit strong chiroptical responses in the visible to near-IR region, which can be tuned by changing the diameter of nanospheres. With such flexibility in the design, helically stacked plasmonic layers may act as tunable chiral metamaterials, as well as serve as different building blocks for chiral assemblies. Based on stepwise GLADs of two metal layers, 3D SPPCMs with high structural degree of freedom have been fabricated, as shown in Figure 16 [94]. Sophisticated modifications based on symmetry considerations to the deposition parameters used in previous GLAD recipes eliminate the undesirable yet unavoidable nature of racemic mixture induced by the long-range disorder of the monolayer domains, resulting in plasmonic chiral metamaterials with definite handedness over macroscopic areas and consequently an order of magnitude enhancement of circular dichroism (CD) over nearly the whole visible spectrum. The CD response can be flexibly controlled by adjusting the deposition conditions and used in refractive index sensing [99].

Figure 16: Stacked-patch plasmonic chiral particles.(A) Left: Schematic diagram of two-step GLAD on a self-assembled microsphere monolayer. The black dashed lines labeled with “Projection 1”, “Projection 2”, and “Projection 2′” are the in-plane projections of the material vapor beams 1, 2, and 2′ in the substrate plane, respectively; θ is the deposition angle of the vapor beam with respect to the normal of the substrate (red dashed line); φ is the azimuthal angle with respect to the reference direction of the microsphere array (black dashed line labeled by “Reference line” overlapped with “Projection 1”); and Δφ=φ1−φ2 is the relative change in azimuthal orientation of the substrate between the first (φ1) and second (φ2) material deposition processes. Right: left-hand (LH) (upper-right) and right-hand (RH) SPPCMs (lower-right) can be prepared, respectively, by two sequential coatings of Ag films following either a clockwise (Δφ=+45°) or anticlockwise (Δφ=−45°) direction. (B) Top-view SEM image of a 90 nm-thick LH-SPPCM prepared at θ=60°, Δφ=+45°. The scale bar in all panels is 1 μm. (C) Measured CD spectra for an LH- (black) and an RH- (red) SPPCM prepared at θ=60° and φ1=0° with Δφ=±45°, respectively. The deposition thickness of two Ag films is 90 nm and the diameter of PS spheres used here is 430 nm, adapted from Ref. [94].
Figure 16:

Stacked-patch plasmonic chiral particles.

(A) Left: Schematic diagram of two-step GLAD on a self-assembled microsphere monolayer. The black dashed lines labeled with “Projection 1”, “Projection 2”, and “Projection 2′” are the in-plane projections of the material vapor beams 1, 2, and 2′ in the substrate plane, respectively; θ is the deposition angle of the vapor beam with respect to the normal of the substrate (red dashed line); φ is the azimuthal angle with respect to the reference direction of the microsphere array (black dashed line labeled by “Reference line” overlapped with “Projection 1”); and Δφ=φ1φ2 is the relative change in azimuthal orientation of the substrate between the first (φ1) and second (φ2) material deposition processes. Right: left-hand (LH) (upper-right) and right-hand (RH) SPPCMs (lower-right) can be prepared, respectively, by two sequential coatings of Ag films following either a clockwise (Δφ=+45°) or anticlockwise (Δφ=−45°) direction. (B) Top-view SEM image of a 90 nm-thick LH-SPPCM prepared at θ=60°, Δφ=+45°. The scale bar in all panels is 1 μm. (C) Measured CD spectra for an LH- (black) and an RH- (red) SPPCM prepared at θ=60° and φ1=0° with Δφ=±45°, respectively. The deposition thickness of two Ag films is 90 nm and the diameter of PS spheres used here is 430 nm, adapted from Ref. [94].

2D plasmonic structures fabricated on substrates by SSL have been investigated extensively during the past decades. Fano resonance and chiral response have been obtained by the nanotriangles fabricated by SSL, either improving figure of merit (FOM) or showing great potentials for sensing chiral molecules. Giessen et al. utilized triangular building blocks fabricated by SSL to build large-area (10×10 mm2), high-quality dimers, trimers, and quadrumers (Figure 10B) with reproducibly small gaps and desired well-defined structural gapless shapes [53]. The triangle oligomers can excite “self-induced” coupling between fundamental dipolar modes and higher-order modes within the same structure that lead to Fano resonances. Fundamental modes, higher-order modes, as well as Fano resonances due to coupling between bright and dark modes within the same complex structure depend on the polarization of incident light and the geometry of the plasmonic structures. Fano resonances occur, in particular, for the case when the plasmonic triangles touch each other by controlling the θ and φ. For this situation, structural symmetry breaking allows for tuning the modes and the shapes of the Fano resonances. These plasmonic triangle oligomers have been used as a refractive index sensor with a theoretical FOM exceeding 15. The triangles can also be used as building blocks to form chiral plasmonic oligomers based on the stepwise SSL [107]. Their chirality is dependent on the anticlockwise or clockwise deposition sequence of the achiral particles. A large chiroptical resonance in the visible region was achieved, and this resonance can be easily adjusted by using CCs with different nanosphere diameters.

Nanocrescents fabricated by SSL can exhibit strong polarization-dependent resonances in the visible and IR regions of the electromagnetic spectrum. The spectra can be well controlled by the structural parameters, which are good for optical filter and localized surface plasmon resonance (LSPR) sensors. Due to the sharp corners of these structures, the crescents are expected to produce a highly enhanced optical near-field and behave as optical antennas [114]. These resonances are strongly dependent on the electric-field polarization of the incident light and were of the same type as required for magnetic resonators. As shown in Figure 17, four extinction peak maxima have been measured for crescents. Most of the extinction peaks are highly sensitive to the orientation of the particle in the polarized light beam. When the long axis of the crescent is orthogonal to the electrical-field polarization, the particles exhibit a strong resonance at 830 nm. When the long axis of the crescent is parallel to the electric field of the polarized beam, the resonance at 830 nm disappears and resonances appear at 1300 nm and 1958 nm. All four resonances have been excited when the crescent orientation is rotated by 45° with respect to the electric-field polarization, as expected for dipolar resonances parallel and perpendicular to the symmetry plane of the crescents. The plasmonic properties can be controlled by the parameters of size, film thickness, opening, and metal deposition angles of the structures [115]. Significant shifts in resonance position are observed when the size or the opening angle of the structures was varied, allowing for a fine tuning of specific resonances in a desired wavelength range. Fundamentally, different responses can be seen for crescents that have clearly separated tips, tips separated by only a small gap, and overlapping tips. One additional peak in the visible range has been identified as the transverse resonance of a rod. The main resonances change upon variations in metal layer thickness and diameter of the masking colloid in a way that was in qualitative agreement with this interpretation. Unique plasmonic properties are also observed from the nanocrescent variations of opposing nanocrescent dimers with sub-10-nm gap [117] and stacked double crescent arrays [118].

Figure 17: UV-vis near-IR polarized extinction spectra for gold crescents fabricated on glass substrate.The SEM images of the 400-nm-diameter particle indicate the orientation of the particle-coated substrate relative to the electrical field (double arrow) of the linearly polarized light, adapted from Ref. [114].
Figure 17:

UV-vis near-IR polarized extinction spectra for gold crescents fabricated on glass substrate.

The SEM images of the 400-nm-diameter particle indicate the orientation of the particle-coated substrate relative to the electrical field (double arrow) of the linearly polarized light, adapted from Ref. [114].

Greatly enhanced color display, local electric fields, surface-enhanced IR absorption, and chiral responses have been realized by fabricating nanostructures with controlled nanogaps using HCL. Shegai et al. have explored the application of the bimetallic nanodisks fabricated by SSL using hole as mask in directional color routing [139]. The pairs of closely spaced gold and silver disks show phase accumulation through material-dependent plasmon resonances, leading to exotic optical properties, in particular scattering of red and blue light in opposite directions, in spite of being as compact as ~λ3/100 (λ is the wavelength). These spectral and spatial photon-sorting nanodevices can be fabricated on a wafer scale and offered a versatile platform for manipulating optical response through polarization, choice of materials, and geometrical parameters, thereby opening possibilities for a wide range of practical applications. Split-rings fabricated by SSL using hole as mask have shown polarization-dependent optical properties and various plasmonic modes. This novel substrate has been employed for antenna-assisted surface-enhanced IR absorption measurements using octadecanethiol (ODT) and deuterated ODT [134]. The method has the potential to make resonant plasmon enhanced IR spectroscopy a standard lab tool in biology, pharmacology, and medicine. Moreover, large-area 3D chiral plasmonic structures can be fabricated by SSL using hole as mask [142]. The optical properties show chiroptical resonances in the 100–400 THz frequency region (750–3000 nm), with CD values of up to 13%. This method offers a simple, low-cost manufacturing method of cm2-sized chiral plasmonic templates for chiroptical applications such as stereochemical enantiomer sensors.

5 Conclusions and outlook

This review summarizes the principle of SSL and gives some exciting design strategies to produce the various nanostructures as well as their promising plasmonic performances. The understanding of the fundamental principle and configuration of SSL is a foundation to flexibly create various nanostructures. As discussed, different and complex nanostructures can be constructed by adjusting parameters such as vapor incident angle θ, azimuthal rotation angle φ, how to rotate θ and φ, deposition materials, the number of deposition, and template or mask. Nanostructures with various shapes or topologies, such as nanorods on nanospheres, patchy nanospheres, nanotriangles, nanodisks, nanocrescents, etc., as well as their assemblies such as dimers, trimers, quadrumers, split-rings, metasurfaces, etc., have been realized. Simple geometrical models based on shadowing effect can be used to predict the formed nanostructures and assemblies and can serve as a designing guide for SSL fabrication. These nanostructures demonstrate different plasmonic properties and can be used in different applications such as LSPR index sensors, extraordinary optical transmission sensors, chiral sensors, SERS, metamaterials, plasmonic heating, optics, light trapping, and color display. It is expected that by varying the SSL parameters and combining SSL with other nanofabrication processes, new and advanced nanostructures can be designed and their novel plasmonic properties and applications will be exploited.

Based on above discussions, the advantages of SSL can be summarized as the following: (1) patterning large areas at low cost and in parallel fashion; (2) fabricating sub-100-nm nanostructures with high resolutions, which are hard to be implemented sometimes by conventional nanolithography techniques; (3) having the flexibility in designing ordered or non-ordered 2D and 3D nanostructures; and (4) depositing materials with almost no limit as long as the materials can be evaporated.

Although investigations have progressed rapidly, challenges in this field still exist. The major obstacle encountered when using SSL is the inability to control mask registry φ. As discussed in Section 2.2.3, in most 2D CC growth, there exists multiple hexagonally close-packed domains with random orientations, and the shape, size, and spacing of nanostructures produced by SSL in different domains could vary significantly at different locations of the substrates. More elaborate nanosphere mask preparation techniques should be developed to allow fabrication of large-area, single-domain nanosphere monolayers. However, multiple domains provide an advantage to simultaneously fabricate multiple nanostructures so that a high throughput screening of nanostructures is possible. Second, the crystallinity of the nanostructure fabricated by SSL may be poor. In order to form the nanostructure with defined geometries predicted by the shadowing effect, the substrate temperature during the deposition should be much lower than the bulk melting temperature of the evaporated material. The films grown at near room temperature are usually amorphous for most semiconductor and compound materials and polycrystalline for metals. For applications requiring high-quality crystalline nanostructures, post-deposition annealing is usually performed (if nanospheres are kept with the nanostructures, the nanospheres should also have high melting temperature such as silica nanospheres). However, the improvement in crystallinity is typically insignificant. Third, high-quality interfaces are lacking for hetero-structures. The low-temperature grown nanostructures usually have fairly roughened surfaces. Many defects could exist at the interfaces that can impair the charge transport properties or other properties across the interfaces. The adatom diffusion has to be increased to improve the quality of the nanostructures.

It is worth noting that simulation models play an important role in SSL due to the fact that nanostructures could become complicated when multiple parameters are changed so that direct visualization of resulted nanostructures is not straightforward. Models based on the geometrical shadowing effect can disclose the basic relationship between the fabricated nanostructure profiles and the process parameters during the deposition and provide a design guide to the fabrication (or potentially could be used to predict the plasmonic properties). One has to be aware that most simulation models only count for the shadowing effect due to the initial CC lattice and do not take into account the growing shadow effect during the deposition. Therefore, it could introduce discrepancies between the model and real structures. In addition, compared to the experimentally resulting nanostructures, other effects such as the surface diffusion of adatom, nucleation, polycrystallinity and roughness of the films, etc., can play a significant role in determining the final structures. To better predict the resulting nanostructures, a kinematic Monte Carlo or molecular dynamic simulation model that accounts for different growth effects could be developed [54]. Clearly, a better simulation model to predict the real structures would be useful.

There is much room left to use SSL to design and fabricate complicated nanostructures and explore different applications. In order to unravel the full power of SSL, a close collaboration among the researchers from different disciplines is crucial. Further research on SSL is believed to mainly focus on four aspects: first, further improvement on the preparation of large-area, high-quality 2D/3D CCs; second, better control of adatom diffusion in the deposition process to better utilize the shadow effect and simultaneously achieve smooth interfaces; third, successful advancement in simulation models to reflect the real fabricated structure; and last, ingenious strategies on designing novel structures. Further exploration on the applications based on the novel nanostructures fabricated by SSL will be carried out. We envision that SSL will continue to evolve and that an increasing number of novel nanostructures with surprising effects/properties will be designed and fabricated. This simple, inexpensive, large-scale nanofabrication and design tool will greatly benefit the field of plasmonics and impact other nanotechnology-related fields such as medical diagnostics, environmental monitoring, food safety, security, etc.

Award Identifier / Grant number: CMMI-1435309

Award Identifier / Grant number: ECCS-1609815

Funding statement: This work is supported by The National Science Foundation under grant nos. CMMI-1435309 and ECCS-1609815, Funder Id: http://dx.doi.org/10.13039/501100008982.

References

[1] Fischer UC, Zingsheim H. Submicroscopic Pattern Replication with Visible Light. J Vacuum Sci Technol 1981;19:881–5.10.1116/1.571227Search in Google Scholar

[2] Deckman H, Dunsmuir J. Natural lithography. Appl Phys Lett 1982;41:377–9.10.1063/1.93501Search in Google Scholar

[3] Hulteen JC, Van duyne RP. Nanosphere lithography – a materials general fabrication process for periodic particle array surfaces. J Vacuum Sci Technol A Vacuum Surf Films 1995;13:1553–8.10.1116/1.579726Search in Google Scholar

[4] Hulteen JC, Treichel DA, Smith MT, Duval ML, Jensen TR, Van Duyne RP. Nanosphere lithography: size-tunable silver nanoparticle and surface cluster arrays. J Phys Chem B 1999;103:3854–63.10.1021/jp9904771Search in Google Scholar

[5] Zhang G, Wang DY. Colloidal lithography – the art of nanochemical patterning. Chem Asian J 2009;4:236–45.10.1002/asia.200800298Search in Google Scholar PubMed

[6] López C. Materials aspects of photonic crystals. Adv Mater 2003;15:1679–704.10.1002/adma.200300386Search in Google Scholar

[7] Xia Y, Gates B, Yin Y, Lu Y. Monodispersed colloidal spheres: old materials with new applications. Adv Mater 2000;12: 693–713.10.1002/(SICI)1521-4095(200005)12:10<693::AID-ADMA693>3.0.CO;2-JSearch in Google Scholar

[8] Dickreuter S, Gleixner J, Kolloch A, Boneberg J, Scheer E, Leiderer P. Mapping of plasmonic resonances in nanotriangles. Beilstein J Nanotechnol 2013;4:588–602.10.3762/bjnano.4.66Search in Google Scholar PubMed

[9] Haynes CL, Van Duyne RP. Nanosphere lithography: a versatile nanofabrication tool for studies of size-dependent nanoparticle optics. J Phys Chem B 2001;5:27.10.1021/jp010657mSearch in Google Scholar

[10] Lee SH, Bantz KC, Lindquist NC, Oh SH, Haynes CL. Self-assembled plasmonic nanohole arrays. Langmuir 2009;25:13685–93.10.1021/la9020614Search in Google Scholar PubMed

[11] Walter R, Tittl A, Berrier A, Sterl F, Weiss T, Giessen H. Large-area low-cost tunable plasmonic perfect absorber in the near infrared by colloidal etching lithography. Adv Opt Mater 2015;3:398–403.10.1002/adom.201400545Search in Google Scholar

[12] Ai B, Yu Y, Mohwald H, Wang LM, Zhang G. Resonant optical transmission through topologically continuous films. ACS Nano 2014;8:1566–75.10.1021/nn4058177Search in Google Scholar PubMed

[13] Naureen S, Sanatinia R, Shahid N, Anand S. High optical quality INP-based nanopillars fabricated by a top-down approach. Nano Lett 2011;11:4805–11.10.1021/nl202628mSearch in Google Scholar PubMed

[14] Yao Y, Yao J, Narasimhan VK, et al. Broadband light management using low-Q whispering gallery modes in spherical nanoshells. Nature Commun 2012;3:664.10.1038/ncomms1664Search in Google Scholar PubMed

[15] Cole RM, Baumberg JJ, Garcia de Abajo FJ, Mahajan S, Abdelsalam M, Bartlett PN. Understanding plasmons in nanoscale voids. Nano Lett 2007;7:2094–100.10.1021/nl0710506Search in Google Scholar

[16] Sun ZQ, Li Y, Zhang JH, et al. A universal approach to fabricate various nanoring arrays based on a colloidal-crystal-assisted-lithography strategy. Adv Funct Mater 2008;18:4036–42.10.1002/adfm.200801103Search in Google Scholar

[17] Kim MH, Choi JY, Choi HK, et al. Carbon nanotube network structuring using two-dimensional colloidal crystal templates. Adv Mater 2008;20:457–61.10.1002/adma.200700956Search in Google Scholar

[18] Chen JX, Liao WS, Chen X, et al. Evaporation-induced assembly of quantum dots into nanorings. ACS Nano 2009;3:173–80.10.1021/nn800568tSearch in Google Scholar PubMed PubMed Central

[19] Zhang G, Wang DY, Mohwald H. Patterning microsphere surfaces by templating colloidal crystals. Nano Lett 2005;5:143–6.10.1021/nl048121aSearch in Google Scholar PubMed

[20] Lin CC, Liao CW, Chao YC, Kuo CS. Fabrication and characterization of asymmetric Janus and ternary particles. ACS Appl Mater Inter 2010;2:3185–91.10.1021/am1006589Search in Google Scholar PubMed

[21] Li Y, Ye XZ, Ma YR, Qi LM. Interfacial nanosphere lithography toward Ag2s-Ag heterostructured nanobowl arrays with effective resistance switching and enhanced photoresponses. Small 2015;11:1183–8.10.1002/smll.201402078Search in Google Scholar PubMed

[22] Liu GQ, Li Y, Duan GT, Wang JJ, Liang CH, Cai WP. Tunable surface plasmon resonance and strong Sers performances of Au opening-nanoshell ordered arrays. ACS Appl Mater Inter 2012;4:1–5.10.1021/am201455xSearch in Google Scholar

[23] Hawkeye MM, Brett MJ. Glancing angle deposition: fabrication, properties, and applications of micro-and nanostructured thin films. J Vac Sci Technol A 2007;25:1317–35.10.1116/1.2764082Search in Google Scholar

[24] Schmidt D, Kjerstad A, Hofmann T, Skomski R, Schubert E, Schubert M. Optical, structural, and magnetic properties of cobalt nanostructure thin films. J Appl Phys 2009;105:113508.10.1063/1.3138809Search in Google Scholar

[25] Briley C, Schmidt D, Hofmann T, Schubert E, Schubert M. Anisotropic magneto-optical hysteresis of permalloy slanted columnar thin films determined by vector magneto-optical generalized ellipsometry. Appl Phys Lett 2015;106:133104.10.1063/1.4916669Search in Google Scholar

[26] Mock A, Korlacki R, Briley C, et al. Anisotropy, band-to-band transitions, phonon modes, and oxidation properties of cobalt-oxide core-shell slanted columnar thin films. Appl Phys Lett 2016;108:051905.10.1063/1.4941399Search in Google Scholar

[27] Kilic U, Mock A, Feder R, et al. Tunable Plasmonic resonances in highly porous nano-bamboo Si-Au superlattice-type thin films. arXiv preprint arXiv:1807.06166 2018.Search in Google Scholar

[28] Takadoum J. Nanomatériaux, traitement et fonctionnalisation des surfaces. Hermes, 2008.Search in Google Scholar

[29] Abelmann L, Lodder C. Oblique evaporation and surface diffusion. Thin Solid Films 1997;305:1–21.10.1016/S0040-6090(97)00095-3Search in Google Scholar

[30] van Kranenburg H, Lodder C. Tailoring growth and local composition by oblique-incidence deposition: a review and new experimental data. Mat Sci Eng R 1994;11:295–354.10.1016/0927-796X(94)90021-3Search in Google Scholar

[31] Zhao YP, Ye DX, Wang GC, Lu TM. Designing nanostructures by glancing angle deposition. Nanotubes and Nanowires 2003;5219:59–73.10.1117/12.505253Search in Google Scholar

[32] Robbie K, Brett MJ. Sculptured thin films and glancing angle deposition: growth mechanics and applications. J Vac Sci Technol A 1997;15:1460–5.10.1116/1.580562Search in Google Scholar

[33] He YP, Fu JX, Zhang Y, et al. Multilayered Si/Ni nanosprings and their magnetic properties. Small 2007;3:153–60.10.1002/smll.200600375Search in Google Scholar PubMed

[34] He YP, Wu JS, Zhao YP. Designing catalytic nanomotors by dynamic shadowing growth. Nano Lett 2007;7:1369–75.10.1021/nl070461jSearch in Google Scholar PubMed

[35] He YP, Zhao YP. Improved hydrogen storage properties of a V decorated Mg nanoblade array. Phys Chem Chem Phys 2009;11:255–8.10.1039/B815924FSearch in Google Scholar

[36] Smith W, Zhao Y-P. Superior photocatalytic performance by vertically aligned core-shell Tio 2/Wo 3 Nanorod arrays. Catalysis Communications 2009;10:1117–21.10.1016/j.catcom.2009.01.010Search in Google Scholar

[37] Zhou CM, Li HF, Gall D. Multi-component nanostructure design by atomic shadowing. Thin Solid Films 2008;517:1214–8.10.1016/j.tsf.2008.05.049Search in Google Scholar

[38] He Y, Zhao Y. Advanced multi-component nanostructures designed by dynamic shadowing growth. Nanoscale 2011;3:2361–75.10.1039/c1nr10103jSearch in Google Scholar PubMed

[39] Yang A, Huntington MD, Cardinal MF, Masango SS, Van Duyne RP, Odom TW. Hetero-oligomer nanoparticle arrays for plasmon-enhanced hydrogen sensing. ACS Nano 2014;8:7639–47.10.1021/nn502502rSearch in Google Scholar PubMed

[40] Suh JY, Kim CH, Zhou W, et al. Plasmonic bowtie nanolaser arrays. Nano Lett 2012;12:5769–74.10.1021/nl303086rSearch in Google Scholar PubMed

[41] Suh JY, Huntington MD, Kim CH, Zhou W, Wasielewski MR, Odom TW. Extraordinary nonlinear absorption in 3D bowtie nanoantennas. Nano Lett 2012;12:269–74.10.1021/nl2034915Search in Google Scholar PubMed PubMed Central

[42] Kitai S, Zhang ZJ, Ji S, Nakamura Y. Characterization of COPT nanowire fabricated by glancing angle deposition. Chinese Phys B 2015;24:056201.10.1088/1674-1056/24/5/056201Search in Google Scholar

[43] Dolatshahi-Pirouz A, Sutherland DS, Foss M, Besenbacher F. Growth characteristics of inclined columns produced by glancing angle deposition (glad) and colloidal lithography. Appl Surf Sci 2011;257:2226–30.10.1016/j.apsusc.2010.09.079Search in Google Scholar

[44] Zhou C, Gall D. Two-component nanorod arrays by glancing-angle deposition. Small 2008;4:1351–4.10.1002/smll.200701289Search in Google Scholar PubMed

[45] Wang J, Huang H, Kesapragada SV, Gall D. Growth of Y-Shaped nanorods through physical vapor deposition. Nano Lett 2005;5:2505–8.10.1021/nl0518425Search in Google Scholar PubMed

[46] Pawar AB, Kretzschmar I. Patchy particles by glancing angle deposition. Langmuir 2008;24:355–8.10.1021/la703005zSearch in Google Scholar PubMed

[47] Bradley L, Zhao Y. Uniform plasmonic response of colloidal Ag Patchy particles prepared by swinging oblique angle deposition. Langmuir 2016;32:4969–74.10.1021/acs.langmuir.6b00980Search in Google Scholar PubMed

[48] Pawar AB, Kretzschmar I. Multifunctional patchy particles by glancing angle deposition. Langmuir 2009;25:9057–63.10.1021/la900809bSearch in Google Scholar PubMed

[49] Larsen GK, He Y, Ingram W, Zhao Y. Hidden chirality in superficially racemic patchy silver films. Nano Lett 2013;13:6228–32.10.1021/nl4036687Search in Google Scholar PubMed

[50] He Y, Larsen GK, Ingram W, Zhao Y. Tunable Three-dimensional helically stacked plasmonic layers on nanosphere monolayers. Nano Lett 2014;14:1976–81.10.1021/nl404823zSearch in Google Scholar PubMed

[51] He Y, Larsen G, Li X, Ingram W, Chen F, Zhao Y. Nanoscale conical Swiss roll with broadband visible and NIR circular dichroism. Adv Opt Mater 2015;3:342–6.10.1002/adom.201400435Search in Google Scholar

[52] Haynes CL, McFarland AD, Smith MT, Hulteen JC, Van Duyne RP. Angle-resolved nanosphere lithography: manipulation of nanoparticle size, shape, and interparticle spacing. J Phys Chem B 2002;106:1898–902.10.1021/jp013570+Search in Google Scholar

[53] Zhao J, Frank B, Burger S, Giessen H. Large-area high-quality plasmonic oligomers fabricated by angle-controlled colloidal nanolithography. ACS Nano 2011;5:9009–16.10.1021/nn203177sSearch in Google Scholar PubMed

[54] Ingram W, He Y, Stone K, Dennis WM, Ye D, Zhao Y. Tuning the plasmonic properties of silver nanopatterns fabricated by shadow nanosphere lithography. Nanotechnology 2016;27:385301.10.1088/0957-4484/27/38/385301Search in Google Scholar PubMed

[55] Cooper CT, Rodriguez M, Blair S, Shumaker-Parry JS. Polarization anisotropy of multiple localized plasmon resonance modes in noble metal nanocrescents. J Phys Chem C 2014;118:1167–73.10.1021/jp4107876Search in Google Scholar

[56] Zhou X, Virasawmy S, Knoll W, Liu KY, Tse MS, Yen LW. Profile simulation and fabrication of gold nanostructures by separated nanospheres with oblique deposition and perpendicular etching. Plasmonics 2007;2:217–30.10.1007/s11468-007-9040-0Search in Google Scholar

[57] Zhou XD, Knoll W, Zhang N, Liu H. Profile calculation of gold nanostructures by dispersed-nanosphere lithography through oblique etching for LSPR applications. J Nanopart Res 2009;11:1065–74.10.1007/s11051-008-9508-7Search in Google Scholar

[58] Nemiroski A, Gonidec M, Fox JM, Jean-Remy P, Turnage E, Whitesides GM. Engineering Shadows to fabricate optical metasurfaces. ACS Nano 2014;8:11061–70.10.1021/nn504214bSearch in Google Scholar PubMed

[59] Zhao J, Jaber S, Mulvaney P, Braun PV, Giessen H. Repetitive hole-mask colloidal lithography for the fabrication of large-area low-cost plasmonic multishape single-layer metasurfaces. Adv Opt Mater 2015;3:680–6.10.1002/adom.201400561Search in Google Scholar

[60] Wickman B, Seidel YE, Jusys Z, Kasemo B, Behm RJ. Fabrication of Pt/Ru nanoparticle pair arrays with controlled separation and their electrocatalytic properties. ACS Nano 2011;5:2547–58.10.1021/nn1021692Search in Google Scholar PubMed

[61] Zhang M, Large N, Koh AL, et al. High-density 2D homo-and hetero-plasmonic dimers with universal sub-10-nm gaps. ACS Nano 2015;9:9331–9.10.1021/acsnano.5b03090Search in Google Scholar PubMed

[62] Tang Z, Wei A. Fabrication of anisotropic metal nanostructures using innovations in template-assisted lithography. ACS Nano 2012;6:998–1003.10.1021/nn300375rSearch in Google Scholar PubMed PubMed Central

[63] Cho K, Loget G, Corn RM. Lithographically Patterned nanoscale electrodeposition of plasmonic, bimetallic, semiconductor, magnetic, and polymer nanoring arrays. J Phys Chem C 2014;118:28993–9000.10.1021/jp501783zSearch in Google Scholar PubMed PubMed Central

[64] Ogier R, Shao L, Svedendahl M, Kall M. Continuous-gradient plasmonic nanostructures fabricated by evaporation on a partially exposed rotating substrate. Adv Mater 2016;28: 4658–64.10.1002/adma.201600112Search in Google Scholar PubMed

[65] Liu L, Wu F, Xiao D, et al. Fabrication of Plasmonic opposite metal spindles in nanowells by shadow deposition for sensing. RSC Adv 2017;7:4759–62.10.1039/C6RA25332FSearch in Google Scholar

[66] Mazzotta F, Hook F, Jonsson MP. High throughput fabrication of plasmonic nanostructures in nanofluidic pores for biosensing applications. Nanotechnology 2012;23:415304.10.1088/0957-4484/23/41/415304Search in Google Scholar PubMed

[67] Zhao Y-P, Ye D-X, Wang P-I, Wang G-C, Lu T-M. Fabrication of Si nanocolumns and Si square spirals on self-assembled monolayer colloid substrates. Int J Nanosci 2002;1:87–97.10.1142/S0219581X02000073Search in Google Scholar

[68] Kesapragada SV, Gall D. Two-component nanopillar arrays grown by glancing angle deposition. Thin Solid Films 2006;494:234–9.10.1016/j.tsf.2005.08.128Search in Google Scholar

[69] Zhou CM, Gall D. The structure of Ta nanopillars grown by glancing angle deposition. Thin Solid Films 2006;515:1223–7.10.1016/j.tsf.2006.07.136Search in Google Scholar

[70] Ghosh A, Fischer P. Controlled propulsion of artificial magnetic nanostructured propellers. Nano Lett 2009;9:2243–5.10.1021/nl900186wSearch in Google Scholar PubMed

[71] Gibbs JG, Zhao YP. Design and characterization of rotational multicomponent catalytic nanomotors. Small 2009;5:2304–8.10.1002/smll.200900686Search in Google Scholar PubMed

[72] Martin-Palma RJ, Manso-Silvan M, Lakhtakia A, Pantano CG. Ordered arrays of nanocolumns grown by the oblique angle deposition technique on a self-assembled layer of polystyrene spheres. Mater Lett 2009;63:197–9.10.1016/j.matlet.2008.09.055Search in Google Scholar

[73] Chen ST, Li ZC, Zhang ZJ. Anisotropic Tixsn1-Xo2 Nanostructures prepared by magnetron sputter deposition. Nanoscale Res Lett 2011;6:326.10.1186/1556-276X-6-326Search in Google Scholar PubMed PubMed Central

[74] Chen ST, Li ZC, Zhang ZJ. Fabrication and optical property of periodic Sn1-Xtixo2 nanostructures patterned by the polystyrene microsphere templates. J Nanomater 2011;2011:21.10.1155/2011/510237Search in Google Scholar

[75] Liang JS, Chen SH, Lin EY, Luo DF, Jiang SJ. Morphology evolution of glancing angle deposition Ag films on nanosphere-array substrates: kinetic Monte Carlo simulation. Comp Mater Sci 2013;79:31–5.10.1016/j.commatsci.2013.06.014Search in Google Scholar

[76] Cheng R, Huang WJ, Huang LJ, et al. Acceleration of tissue plasminogen activator-mediated thrombolysis by magnetically powered nanomotors. ACS Nano 2014;8:7746–54.10.1021/nn5029955Search in Google Scholar PubMed PubMed Central

[77] Hu J, Huang W, Huang S, ZhuGe Q, Jin K, Zhao Y. Magnetically active Fe3O4 nanorods loaded with tissue plasminogen activator for enhanced thrombolysis. Nano Research 2016;9:2652–61.10.1007/s12274-016-1152-4Search in Google Scholar

[78] Zhou CM, Gall D. Competitive growth of Ta nanopillars during glancing angle deposition: effect of surface diffusion. J Vac Sci Technol A 2007;25:312–8.10.1116/1.2539328Search in Google Scholar

[79] Zhou CM, Gall D. Branched Ta nanocolumns grown by glancing angle deposition. Appl Phys Lett 2006;88:203117.10.1063/1.2204759Search in Google Scholar

[80] Kesapragada SV, Gall D. Anisotropic broadening of Cu nanorods during glancing angle deposition. Appl Phys Lett 2006;89:203121.10.1063/1.2388861Search in Google Scholar

[81] Pawar AB, Kretzschmar I. Fabrication, Assembly, and application of patchy particles. Macromol Rapid Comm 2010;31:150–68.10.1002/marc.200900614Search in Google Scholar PubMed

[82] Glotzer SC, Solomon MJ. Anisotropy of building blocks and their assembly into complex structures. Nature Mater 2007;6:557–62.10.1038/nmat1949Search in Google Scholar PubMed

[83] Zhang Y, Barhoumi A, Lassiter JB, Halas NJ. Orientation-preserving transfer and directional light scattering from individual light-bending nanoparticles. Nano Lett 2011;11:1838–44.10.1021/nl2008357Search in Google Scholar PubMed

[84] Yu Y, Gan L, Zhang G, Yang B. Asymmetric microparticles and heterogeneous microshells via angled colloidal lithography. Colloid Surface A 2012;405:51–8.10.1016/j.colsurfa.2012.04.035Search in Google Scholar

[85] Singh DP, Nagar R, Singh JP. Synthesis of Siox-Ag core-shell nanostructures by oblique angle deposition. J Appl Phys 2010;107:074306.10.1063/1.3366714Search in Google Scholar

[86] Imura Y, Kato M, Kondo T, Kawai T. Strings of metal half-shells fabricated using colloidal particle monolayer as a template. Langmuir 2010;26:11314–8.10.1021/la100626vSearch in Google Scholar PubMed

[87] Zhang N, Su XD, Free P, et al. Plasmonic metal nanostructure array by glancing angle deposition for biosensing application. Sensor Actuat B-Chem 2013;183:310–8.10.1016/j.snb.2013.03.088Search in Google Scholar

[88] Ingram WM, Han C, Zhang Q, Zhao Y. Optimization of Ag-coated polystyrene nanosphere substrates for quantitative surface-enhanced Raman spectroscopy analysis. J Phys Chem C 2015;119:27639–48.10.1021/acs.jpcc.5b06896Search in Google Scholar

[89] Chen JY, Hou YD, Du JL, Zhu JH, Gao FH. Fabrication and microanalysis of 3D-shell-like chiral nanostructure based on micro-sphere assembled technology. Microelectron Eng 2016;153:137–41.10.1016/j.mee.2016.03.056Search in Google Scholar

[90] Giordano MC, Foti A, Messina E, Gucciardi PG, Comoretto D, Buatier de Mongeot F. Sers Amplification from self-organized arrays of plasmonic nanocrescents. ACS Appl Mater Interfaces 2016;8:6629–38.10.1021/acsami.5b11843Search in Google Scholar PubMed

[91] Hou YD, Li SH, Su YR, et al. Design and fabrication of three-dimensional chiral nanostructures based on stepwise glancing angle deposition technology. Langmuir 2013;29:867–72.10.1021/la304122fSearch in Google Scholar PubMed

[92] Su YR, Ye S, Hou YD, et al. Design and fabrication of diverse three-dimensional shell-like nano-structures. Microelectron Eng 2014;115:6–12.10.1016/j.mee.2013.10.013Search in Google Scholar

[93] Fu T, Qu Y, Wang T, et al. Tunable chiroptical response of chiral plasmonic nanostructures fabricated with chiral templates through oblique angle deposition. J Phys Chem C 2017;121:1299–304.10.1021/acs.jpcc.6b10833Search in Google Scholar

[94] Hou Y, Leung HM, Chan CT, Du J, Chan HL-W, Lei DY. Ultrabroadband optical superchirality in a 3D stacked-patch plasmonic metamaterial designed by two-step glancing angle deposition. Adv Funct Mater 2016;26:7807–16.10.1002/adfm.201602800Search in Google Scholar

[95] Wang YK, Deng JC, Wang G, Fu T, Qu Y, Zhang ZY. Plasmonic chirality of L-shaped nanostructure composed of two slices with different thickness. Opt Express 2016;24:2307–17.10.1364/OE.24.002307Search in Google Scholar PubMed

[96] Larsen GK, He YZ, Wang J, Zhao YP. Scalable fabrication of composite Ti/Ag plasmonic helices: controlling morphology and optical activity by tailoring material properties. Adv Opt Mater 2014;2:245–9.10.1002/adom.201300478Search in Google Scholar

[97] He Y, Lawrence K, Ingram W, Zhao Y. Strong local chiroptical response in racemic patchy silver films: enabling a large-area chiroptical device. ACS Photonics 2015;2:1246–52.10.1021/acsphotonics.5b00196Search in Google Scholar

[98] Larsen GK, Stom N, Zhao Y. Continuously tuning the spectral response of chiral plasmonic patchy particles through galvanic replacement reaction. RSC Adv 2015;5:101257–61.10.1039/C5RA16939ASearch in Google Scholar

[99] He Y, Lawrence K, Ingram W, Zhao Y. Circular dichroism based refractive index sensing using chiral metamaterials. Chem Commun 2016;52:2047–50.10.1039/C5CC07742GSearch in Google Scholar

[100] Xie D, Hou YD, Su YR, Gao FH, Du JL. Templated fabrication of hollow nanospheres with ‘windows’ of accurate size and tunable number. Nanoscale Res Lett 2015;10:153.10.1186/s11671-015-0795-5Search in Google Scholar PubMed PubMed Central

[101] He Z, Kretzschmar I. Template-assisted glad: approach to single and multipatch patchy particles with controlled patch shape. Langmuir 2013;29:15755–61.10.1021/la404592zSearch in Google Scholar PubMed

[102] Zhang G, Wang DY, Moehwald H. Nanoembossment of Au patterns on microspheres. Chem Mater 2006;18:3985–92.10.1021/cm060358fSearch in Google Scholar

[103] He Z, Kretzschmar I. Template-assisted fabrication of patchy particles with uniform patches. Langmuir 2012;28:9915–9.10.1021/la3017563Search in Google Scholar PubMed

[104] Wang H-F, Shi J-X, Qian L-Y, Yan C-C, Han C-Q, Zhao Y-P. Large-area broadband optical absorber fabricated by shadowing sphere lithography. Opt Express 2018;26:7507–15.10.1364/OE.26.007507Search in Google Scholar PubMed

[105] Haynes CL, Van Duyne RP. Dichroic optical properties of extended nanostructures fabricated using angle-resolved nanosphere lithography. Nano Lett 2003;3:939–43.10.1021/nl0342287Search in Google Scholar

[106] Hoffmann JM, Janssen H, Chigrin DN, Taubner T. Enhanced infrared spectroscopy using small-gap antennas prepared with two-step evaporation nanosphere lithography. Opt Express 2014;22:14425–32.10.1364/OE.22.014425Search in Google Scholar PubMed

[107] Xie SW, Yang JZ, Xiao X, et al. Scalable fabrication of quasi-three-dimensional chiral plasmonic oligomers based on stepwise colloid sphere lithography technology. Nanoscale Res Lett 2015;10:393.10.1186/s11671-015-1102-1Search in Google Scholar PubMed PubMed Central

[108] Kosiorek A, Kandulski W, Chudzinski P, Kempa K, Giersig M. Shadow nanosphere lithography: simulation and experiment. Nano Lett 2004;4:1359–63.10.1021/nl049361tSearch in Google Scholar

[109] Ingram W, Larson S, Carlson D, Zhao Y. Ag-Cu mixed phase plasmonic nanostructures fabricated by shadow nanosphere lithography and glancing angle Co-deposition. Nanotechnology 2017;28:015301.10.1088/0957-4484/28/1/015301Search in Google Scholar PubMed

[110] Ai B, Song C, Bradley L, Zhao Y. Strong Fano resonance excited in an array of nanoparticle-in-ring nanostructures for dual plasmonic sensor applications. J Phys Chem C 2018;122:20935–44.10.1021/acs.jpcc.8b05154Search in Google Scholar

[111] Rochholz H, Bocchio N, Kreiter M. Tuning resonances on crescent-shaped noble-metal nanoparticles. New J Phys 2007;9:53.10.1088/1367-2630/9/3/053Search in Google Scholar

[112] Cooper CT, Rodriguez M, Blair S, Shumaker-Parry JS. Mid-Infrared localized plasmons through structural control of gold and silver nanocrescents. J Phys Chem C 2015;119:11826–32.10.1021/acs.jpcc.5b01529Search in Google Scholar

[113] Bochenkov VE, Sutherland DS. From rings to crescents: a novel fabrication technique uncovers the transition details. Nano Lett 2013;13:1216–20.10.1021/nl304675wSearch in Google Scholar PubMed

[114] Shumaker-Parry JS, Rochholz H, Kreiter M. Fabrication of crescent-shaped optical antennas. Adv Mater 2005;17:2131–4.10.1002/adma.200500063Search in Google Scholar

[115] Wu LY, Ross BM, Lee LP. Optical Properties of the Crescent-Shaped Nanohole Antenna. Nano Lett 2009;9:1956–61.10.1021/nl9001553Search in Google Scholar PubMed PubMed Central

[116] Lewicka ZA, Li Y, Bohloul A, Yu WW, Colvin VL. Nanorings and nanocrescents formed via shaped nanosphere lithography: a route toward large areas of infrared metamaterials. Nanotechnology 2013;24:115303.10.1088/0957-4484/24/11/115303Search in Google Scholar PubMed

[117] Fischer J, Vogel N, Mohammadi R, et al. Plasmon hybridization and strong near-field enhancements in opposing nanocrescent dimers with tunable resonances. Nanoscale 2011;3:4788–97.10.1039/c1nr10952aSearch in Google Scholar PubMed

[118] Vogel N, Fischer J, Mohammadi R, et al. Plasmon hybridization in stacked double crescents arrays fabricated by colloidal lithography. Nano Lett 2011;11:446–54.10.1021/nl103120sSearch in Google Scholar PubMed

[119] Retsch M, Tamm M, Bocchio N, et al. Parallel preparation of densely packed arrays of 150-nm gold-nanocrescent resonators in three dimensions. Small 2009;5:2105–10.10.1002/smll.200900162Search in Google Scholar PubMed

[120] Choi Y, Hong S, Lee LP. Shadow overlap ion-beam lithography for nanoarchitectures. Nano Lett 2009;9:3726–31.10.1021/nl901911pSearch in Google Scholar PubMed

[121] Li ZB, Zhang XM, Ye SS, et al. The fabrication of long-range ordered nanocrescent structures based on colloidal lithography and parallel imprinting. Nanotechnology 2013;24:105307.10.1088/0957-4484/24/10/105307Search in Google Scholar PubMed

[122] Swartz M, Rodriguez M, Quast AD, Cooper CT, Blair S, Shumaker-Parry JS. Aluminum nanocrescent plasmonic antennas fabricated by copper mask nanosphere template lithography. J Phys Chem C 2016;120:20597–603.10.1021/acs.jpcc.6b00396Search in Google Scholar

[123] Liu XF, Choi B, Gozubenli N, Jiang P. Periodic arrays of metal nanorings and nanocrescents fabricated by a scalable colloidal templating approach. J Colloid Interf Sci 2013;409:52–8.10.1016/j.jcis.2013.07.018Search in Google Scholar PubMed

[124] Kosiorek A, Kandulski W, Glaczynska H, Giersig M. Fabrication of nanoscale rings, dots, and rods by combining shadow nanosphere lithography and annealed polystyrene nanosphere masks. Small 2005;1:439–44.10.1002/smll.200400099Search in Google Scholar PubMed

[125] Gwinner MC, Koroknay E, Fu L, et al. Periodic large-area metallic split-ring resonator metamaterial fabrication based on shadow nanosphere lithography. Small 2009;5:400–6.10.1002/smll.200800923Search in Google Scholar PubMed

[126] Brodoceanu D, Fang C, Voelcker NH, et al. Fabrication of metal nanoparticle arrays by controlled decomposition of polymer particles. Nanotechnology 2013;24:085304.10.1088/0957-4484/24/8/085304Search in Google Scholar PubMed

[127] Saracut V, Giloan M, Gabor M, Astilean S, Farcau C. Polarization-Sensitive linear plasmonic nanostructures via colloidal lithography with uniaxial colloidal arrays. ACS Appl Mater Inter 2013;5:1362–9.10.1021/am3026745Search in Google Scholar PubMed

[128] Kostinski SV, Chen ER, Brenner MP. Characterization of patterns formed by shadows of spheres. Phys Rev Lett 2014;112:235502.10.1103/PhysRevLett.112.235502Search in Google Scholar PubMed

[129] Zhang G, Wang DY, Mohwald H. Ordered binary arrays of au nanoparticles derived from colloidal lithography. Nano Lett 2007;7:127–32.10.1021/nl062284cSearch in Google Scholar PubMed

[130] Zhang G, Wang DY. Fabrication of heterogeneous binary arrays of nanoparticles via colloidal lithography. J Am Chem Soc 2008;130:5616–7.10.1021/ja710771jSearch in Google Scholar PubMed

[131] Zhang G, Wang DY, Mohwald H. Fabrication of multiplex quasi-three-dimensional grids of one-dimensional nanostructures via stepwise colloidal lithography. Nano Lett 2007;7:3410–3.10.1021/nl071820dSearch in Google Scholar PubMed

[132] Ai B, Basnet P, Larson S, Ingram W, Zhao Y. Plasmonic sensor with high figure of merit based on differential polarization spectra of elliptical nanohole array. Nanoscale 2017;9:14710–21.10.1039/C7NR06067JSearch in Google Scholar PubMed

[133] Fredriksson H, Alaverdyan Y, Dmitriev A, et al. Hole-mask colloidal lithography. Adv Mater 2007;19:4297–302.10.1002/adma.200700680Search in Google Scholar

[134] Cataldo S, Zhao J, Neubrech F, et al. Hole-mask colloidal nano lithography for large-area low-cost metamaterials and antenna-assisted surface-enhanced infrared absorption substrates. ACS Nano 2012;6:979–85.10.1021/nn2047982Search in Google Scholar PubMed

[135] Frederiksen M, Sutherland DS. Direct Modification of colloidal hole-masks for locally ordered hetero-assemblies of nanostructures over large areas. Nanoscale 2014;6:731–5.10.1039/C3NR03871HSearch in Google Scholar PubMed

[136] Ogier R, Fang Y, Svedendahl M, Johansson P, Käll M. Macroscopic layers of chiral plasmonic nanoparticle oligomers from colloidal lithography. ACS Photonics 2014;1:1074–81.10.1021/ph500293uSearch in Google Scholar

[137] Zhao J, Zhang C, Braun PV, Giessen H. Large-Area low-cost plasmonic nanostructures in the NIR for Fano resonant sensing. Adv Mater 2012;24:OP247–52.10.1002/adma.201202109Search in Google Scholar PubMed

[138] Verre R, Svedendahl M, Lank NO, et al. Directional light extinction and emission in a metasurface of tilted plasmonic nanopillars. Nano Lett 2016;16:98–104.10.1021/acs.nanolett.5b03026Search in Google Scholar PubMed

[139] Shegai T, Chen S, Miljkovic VD, Zengin G, Johansson P, Kall M. A bimetallic nanoantenna for directional colour routing. Nat Commun 2011;2:481.10.1038/ncomms1490Search in Google Scholar PubMed PubMed Central

[140] Zhao J, Frank B, Neubrech F, Zhang C, Braun PV, Giessen H. Hole-mask colloidal nanolithography combined with tilted-angle-rotation evaporation: a versatile method for fabrication of low-cost and large-area complex plasmonic nanostructures and metamaterials. Beilstein J Nanotechnol 2014;5:577–86.10.3762/bjnano.5.68Search in Google Scholar PubMed PubMed Central

[141] Feng HY, Luo F, Meneses-Rodriguez D, Armelles G, Cebollada A. From Disk to ring: aspect ratio control of the magnetoplasmonic response in Au/Co/Au nanostructures fabricated by hole-mask colloidal lithography. Appl Phys Lett 2015;106:083105.10.1063/1.4913621Search in Google Scholar

[142] Frank B, Yin X, Schaferling M, et al. Large-area 3D chiral plasmonic structures. ACS Nano 2013;7:6321–9.10.1021/nn402370xSearch in Google Scholar PubMed

[143] Feng HY, Luo F, Arenal R, et al. Active magnetoplasmonic split-ring/ring nanoantennas. Nanoscale 2017;9:37–44.10.1039/C6NR07864HSearch in Google Scholar PubMed

[144] Boggild T, Runager K, Sutherland DS. Nanopattern gradients for cell studies fabricated using hole-mask colloidal lithography. ACS Appl Mater Interfaces 2016;8:14975–9.10.1021/acsami.5b08315Search in Google Scholar PubMed

[145] Wi JS, Son JG, Han SW, Lee TG. Nanoparticles inside nanodishes for plasmon excitations. Appl Phys Lett 2015;107:203102.10.1063/1.4935860Search in Google Scholar

[146] Shen Y, Liu M, Li J, et al. Extraordinary transmission of three-dimensional crescent-like holes arrays. Plasmonics 2012;7:221–7.10.1007/s11468-011-9297-1Search in Google Scholar PubMed PubMed Central

[147] Shen Y, Chen X, Dou Z, et al. Tuning the plasmon resonance of a nano-mouth array. Nanoscale 2012;4:5576–80.10.1039/c2nr31516eSearch in Google Scholar PubMed

[148] Shen Y, Liu M, Wang Q, et al. Fabrication of Non-planar silver nano-Arc-gap arrays. Nanoscale 2012;4:2255–9.10.1039/c2nr30152kSearch in Google Scholar PubMed

[149] Wang PY, Bennetsen DT, Foss M, Ameringer T, Thissen H, Kingshott P. Modulation of human mesenchymal stem cell behavior on ordered tantalum nanotopographies fabricated using colloidal lithography and glancing angle deposition. ACS Appl Mater Inter 2015;7:4979–89.10.1021/acsami.5b00107Search in Google Scholar PubMed

[150] Zhou CM, Gall D. Surface patterning by nanosphere lithography for layer growth with ordered pores. Thin Solid Films 2007;516:433–7.10.1016/j.tsf.2007.05.069Search in Google Scholar

[151] Patzig C, Karabacak T, Fuhrmann B, Rauschenbach B. Glancing angle sputter deposited nanostructures on rotating substrates: experiments and simulations. J Appl Phys 2008;104:094318.10.1063/1.3018145Search in Google Scholar

[152] Khare C, Patzig C, Gerlach JW, Rauschenbach B, Fuhrmann B. Influence of substrate temperature on glancing angle deposited Ag nanorods. J Vac Sci Technol A 2010;28:1002–9.10.1116/1.3447231Search in Google Scholar

[153] Patzig C, Rauschenbach B, Fuhrmann B, Leipner HS. Growth of Si nanorods in honeycomb and hexagonal-closed-packed arrays using glancing angle deposition. J Appl Phys 2008;103:024313.10.1063/1.2836962Search in Google Scholar

[154] Zhou CM, Gall D. Development of two-level porosity during glancing angle deposition. J Appl Phys 2008;103:014307.10.1063/1.2828174Search in Google Scholar

[155] Ai B, Wang LM, Mohwald H, Yu Y, Zhang G. Asymmetric half-cone/nanohole array films with structural and directional reshaping of extraordinary optical transmission. Nanoscale 2014;6:8997–9005.10.1039/C4NR01385ASearch in Google Scholar PubMed

[156] Ai B, Wang LM, Moehwald H, et al. Real-Time control of uni-directional liquid spreading on a half-cone nanoshell array. Sci Rep-Uk 2014;4:6751.10.1038/srep06751Search in Google Scholar PubMed PubMed Central

[157] Wang T, Chen H, Liu K, et al. Anisotropic Janus Si nanopillar arrays as a microfluidic one-way valve for gas-liquid separation. Nanoscale 2014;6:3846–53.10.1039/C3NR05865DSearch in Google Scholar

[158] Wang T, Chen H, Liu K, et al. Janus Si Micropillar arrays with thermal-responsive anisotropic wettability for manipulation of microfluid motions. ACS Appl Mater Interfaces 2015;7:376–82.10.1021/am5063647Search in Google Scholar PubMed

[159] Ge P, Wang S, Liu W, et al. Unidirectional wetting of liquids on “Janus” nanostructure arrays under various media. Langmuir 2017;33:2177–84.10.1021/acs.langmuir.7b00034Search in Google Scholar PubMed

[160] Yi GR, Pine DJ, Sacanna S. Recent progress on patchy colloids and their self-assembly. J Phys-Condens Mat 2013;25:193101.10.1088/0953-8984/25/19/193101Search in Google Scholar PubMed

Received: 2018-07-24
Revised: 2018-09-10
Accepted: 2018-09-23
Published Online: 2018-10-06

©2018 Bin Ai et al., published by De Gruyter, Berlin/Boston

This work is licensed under the Creative Commons Attribution-NonCommercial-NoDerivatives 3.0 License.

Downloaded on 24.5.2024 from https://www.degruyter.com/document/doi/10.1515/nanoph-2018-0105/html
Scroll to top button