Towards infrared reconfigurable plasmonics and metamaterialscj...supporting strong resonances at...

121
Towards Infrared Reconfigurable Plasmonics and Metamaterials A dissertation presented by Kan Yao to the Department of Electrical and Computer Engineering in partial fulfillment of the requirements for the degree of Doctor of Philosophy in the field of Electrical Engineering Northeastern University Boston, Massachusetts July 2017

Transcript of Towards infrared reconfigurable plasmonics and metamaterialscj...supporting strong resonances at...

  • Towards Infrared Reconfigurable

    Plasmonics and Metamaterials

    A dissertation presented

    by

    Kan Yao

    to

    the Department of Electrical and Computer Engineering

    in partial fulfillment of the requirements

    for the degree of

    Doctor of Philosophy

    in the field of

    Electrical Engineering

    Northeastern University

    Boston, Massachusetts

    July 2017

  • ii

    © Copyright 2017

    Kan Yao

    All Rights Reserved

  • iii

    Abstract

    Metamaterials are rationally engineered composite artificial structures with exotic

    properties not easily obtained or altogether unavailable in nature. Since the proof-of-

    concept experiments in the microwave regime in the early 2000s, the operating

    wavelength of metamaterials has been quickly pushed from terahertz (THz) to mid-/near-

    infrared and down to visible, where metamaterials started to interfere with another

    emerging field of plasmonics. The dynamic interplay between them has enabled various

    innovative concepts and applications, which are impossible with either area alone.

    Despite the continuous blending of metamaterials and plasmonics, there are still

    several sub-fields not fully explored and limitations not yet resolved. For instance, metals

    do not support pronounced plasmonic responses at mid-infrared and far-infrared

    frequencies. On the other hand, infrared light has found widespread applications such as

    biomedical sensing, spectroscopy, and surveillance. The lack of suitable platforms

    supporting strong resonances at infrared wavelengths, as a consequence, naturally leads

    to the quest of better alternatives for infrared plasmonics and metamaterials.

    Another restriction is tunability. So far, most metamaterials and plasmonic devices are

    in solid phase and have fixed functionalities once they are fabricated. Since the optical

    properties of metamaterials and plasmonic structures rely on both the constituents and the

    way in which they are shaped and organized, introducing variables into the system, either

    changeable material properties or reconfigurable structures, will open up new

    opportunities for dynamically tunable metamaterials and plasmonic structures.

  • iv

    In this work, aiming at solving both problems mentioned above, I will present several

    solutions to achieve infrared reconfigurable plasmonics and metamaterials. Simulations

    and experiments are conducted respectively for the design and realization of infrared

    optical devices, covering different aspects of this topic. I will show plasmonic core/shell

    particles and their self-assembled clusters as infrared resonators, ultra-compact

    nanoantennas with switchable and unidirectional radiation properties, and tunable

    graphene metasurfaces for infrared light beam shaping. Future work and preliminary

    results on programmable plasmonic metamolecules via magnetic self-assembly will also

    be discussed.

  • v

    TABLE OF CONTENTS

    Table of Contents

    List of Figures

    Acknowledgements

    Chapter 1 Introduction ............................................................................................. 1

    Chapter 2 Plasmonic Core-Shell Particles by Chemical Synthesis .............................. 10

    2.1 Introduction ........................................................................................................... 10

    2.2 Synthesis of Plasmonic Core-shell Particles ........................................................... 13

    2.3 Assembly of Clusters .............................................................................................. 23

    Chapter 3 Self-Assembled Plasmonic Core-Shell Particles as Infrared Resonators..... 28

    3.1 Bright-Field Reflection Measurements .................................................................. 28

    3.2 Simulation of Single Particle Scattering on a Gold Substrate ................................ 35

    3.3 Simulation of Single Cluster Scattering on a Gold Substrate ................................. 40

    3.4 Summary and Outlook ........................................................................................... 44

    Chapter 4 Designer Plasmonic Core-Shell Particles by Microfabrication ..................... 47

    4.1 Introduction ........................................................................................................... 47

    4.2 Self-Assembly of Polystyrene Particle Monolayers ............................................... 49

    4.3 Core-Shell Particles by Microfabrication ................................................................ 52

    4.4 Conclusions ............................................................................................................ 56

    Chapter 5 Ultracompact Nanoantennas with Switchable Directionality ..................... 57

    5.1 Introduction ........................................................................................................... 57

    5.2 Design of Ultracompact Unidirectional Nanoantennas ......................................... 59

    5.3 Analysis and Discussions ........................................................................................ 70

  • vi

    5.4 Conclusions ............................................................................................................ 73

    Chapter 6 Tunable Graphene Metasurfaces.................................................................. 75

    6.1 Introduction ........................................................................................................... 75

    6.2 Infrared Light Beam Shaping by Graphene Metasurfaces ..................................... 78

    6.2.1 Anomalous Reflection .................................................................................... 82

    6.2.2 Reflective Focusing Lens................................................................................. 84

    6.2.3 Non-diffracting Airy Beams ............................................................................ 86

    6.3 Infrared Circular Dichroism by Graphene Metamirrors ......................................... 88

    6.4 Conclusions ............................................................................................................ 90

    Chapter 7 Summary and Future work ........................................................................... 92

    7.1 Summary ................................................................................................................ 92

    7.2 Future Work ................................................................................................... 93

    6.2.1 Programmable Plasmonic Metamolecules via Magnetic Self-Assembly ....... 93

    6.2.2 Core-Shell Particles for Sensing Applications ................................................. 98

    References .............................................................................................................. 100

  • vii

    LIST OF FIGURES

    Figure 1.1: Schematics of some representative metamaterials. (a) A lattice of metallic

    wires to achieve negative permittivity at low-frequency. (b) Metallic split ring

    resonators (SRRs) to produce artificial magnetism. (c) Metallic helices as building

    blocks for chiral metamaterials. ....................................................................................1 Figure 1.2: (a) Illustration of SPPs and oscillation of surface charges at a dielectric/metal

    interface. (b) The typical electromagnetic field profile of SPPs, reflecting their

    bound, non-radiant nature. ............................................................................................2

    Figure 1.3: Reconfigurable metamaterials and metadevices by material property changes.

    (a) Functional devices can be written in or erased from a chalcogenide glass (GST)

    phase-change film by trains of femtosecond laser pulses at different conditions. (b)

    The optical image of two Fresnel zone patterns written by light. (c) The second

    Fresnel zone pattern on the right is erased by light. (d) A hybrid metamaterial

    consisting of a vanadium oxide (VO2) film sandwiched by gold meshes and a ground

    plane. Applying electrical currents will induce the insulator-to-metal transition of

    VO2 and change the reflectance. (e) Top left: Microscopic image of the metamaterial.

    Letters are bare VO2 areas without gold meshes. Rest panels in clockwise order:

    Infrared images at 2.67 μm wavelength with 0 A, 2.03 A, and 2.20 A current,

    respectively. (f) An infrared light beam excites tunable plasmon resonances across

    the graphene nanoribbons biased at varying gate voltages. (g) Plasmon-enhanced

    light-matter interactions improve detection sensitivity of molecules adsorbed on

    graphene. (a-c) are adopted from Ref. [13], (d, e) are adopted from Ref. [14], (f) is

    adopted from Ref. [17], and (g) is adopted from Ref. [18], respectively. ....................4

    Figure 1.4: Reconfigurable metamaterials and metamolecules by finite structural changes.

    (a) Split ring resonator (SRR)-bar on a stretchable film. Mechanical deformation of

    the elastomeric substrate changes the gap size and thus the coupling, which causes

    noticeable resonance shift in (b). (c) A metamaterial consisting of “meander near the

    wire” gold nanostructures on silicon nitride strings. (d) Applying voltages to the

    neighboring strings induces an attractive electrostatic force in the plane, which

    reduces the gap size, enabling modulation of transmission and reflection. (e, f) Two

    gold nanorods are hosted on a DNA template to form a chiral metamolecule

    exhibiting tunable chiroptical properties as in (g). (e) The relative angle between the

    nanorods can be actively controlled by adding certain DNA strands to pair the locks

    on the sides of the template. (f) Alternatively, the nanorods are assembled

    perpendicular to each other, and the DNA locks function as binding sites that allow

    the rod on the top (yellow) to “walk” stepwise between the two ends of the template.

    (a, b) are adopted from Ref. [20], (c, d) are adopted from Ref. [21], (e) is adopted

    from Ref. [23], and (f, g) are adopted from Ref. [24]. ..................................................6

    Figure 2.1: Optical resonators at different wavelengths. (a) A high-index dielectric (lead

    telluride (PbTe)) particle can support a series of Mie resonance modes in the mid

    infrared (MIR) region. (b) High-index dielectric (silicon) particle clusters exhibit

    limited spectral tunability at visible frequencies. (c) Plasmonic core-shell

    nanoparticle clusters show complex spectral features in the near infrared (NIR)

    region. Figures are adopted from Refs. [31], [35], and [39], respectively. ................12

  • viii

    Figure 2.2: Flow chart of the synthesis of SiO2/Au core-shell particles. (a) Synthesis of

    gold nanoparticle seeds (i) and modification of SiO2 particles (ii). (b) Preparation of

    precursor particles, where gold nanoparticles are attached to the surface of the

    modified silica beads. (c) Continuous reduction of gold ions in the planting solution

    causes the nanoparticles to increase in size, coalesce, and eventually form complete

    shells covering the silica cores. ..................................................................................14

    Figure 2.3: Typical ultraviolet-visible (UV-Vis) extinction spectra of the THPC gold

    nanoparticle solution. Different curves are collected from different batches of

    samples. All of them exhibit a vague bump close to 500 nm wavelength, indicating

    the sizes of the synthesized nanoparticles are 1-3 nm. Inset shows the image of an

    aged solution, which is dark brown in color. ..............................................................16

    Figure 2.4: Scanning Electron Microscope (SEM) images of a 1 μm (a) and a 2 μm (b)

    core-shell particle (monomer) on a gold-coated substrate. The shell thicknesses are

    estimated to be around 30 and 10 nm, respectively. ...................................................19

    Figure 2.5: Zoom-in SEM images for shell inspection. (a) The surface of a 2 μm particle

    coated by a complete and smooth gold shell. Occasional cracks such as pinhole

    defects and excess gold nanoparticles are smaller than 40 nm and have negligible

    impact on the infrared optical properties of the core-shell particles. (b) Perspective

    view (~70°) of a 2 μm core-shell particle at a rectangular cut on its surface. The cut

    is made by focused ion beam (FIB). Measurements at different locations on the

    cutting edges suggest a shell thickness of ~10 nm. ....................................................20

    Figure 2.6: (a,c) SEM images of a 1 μm (a) and a 2 μm (c) particle after 5 cycles of gold

    reduction with a small amount of planting solution. The gold nanoparticles increase

    in size (~10 nm) but are isolated. (b) Zoom-in image of a 1 μm particle after 12

    cycles of reduction. The grain size increases to 20 nm and larger, whereas the

    nanoparticles still do not coalesce into a complete shell. (d) Zoom-in images of a 2

    μm particle after 5 cycles of reduction with a small amount of planting solution and

    an additional cycle with a large amount (5 times in volume) of planting solution. A

    complete shell is formed whereas the smoothness degrades slightly. ........................22

    Figure 2.7: SEM image of a 1 μm particle after a single cycle of reduction with a

    condensed planting solution. The attached seed nanoparticles increase significantly

    in size but are clearly isolated. ....................................................................................23

    Figure 2.8: Self-assembly of core-shell particles. A droplet of diluted particle suspension

    is pipetted on a gold-coated substrate. During total evaporation, the core-shell

    particles are assembled by the capillary force into different planar structures. ..........24

    Figure 2.9: Self-assembled 1 μm core-shell clusters. Representative geometries include

    but are not limited to: (a) dimers, (b-c) trimers, and (d) chains. The shell thicknesses

    are estimated to be about 30 nm. Scale bars are 200 nm. ...........................................24

    Figure 2.10: Self-assembled 2 μm core-shell clusters. Representative configurations

    include but are not limited to: (a) dimers, (b) trimers, (c) quadrumers, (d) chains, and

    (e) pentamers. The shell thicknesses are estimated to be about 10 nm. Scale bars are

    500 nm. .......................................................................................................................25

    Figure 2.11: Two clusters consisting of particles with incomplete coating. Some of the

    attached nanoparticles drop off during the movement of the particles, forming a trial

    on the substrate. The scale bars are 500 nm in (a) and 1 μm in (b). ...........................26

  • ix

    Figure 3.1: Schematic of the Fourier transform infrared (FTIR) single cluster

    spectroscopy. Infrared light is generated by a globar source. After passing through a

    Michelson interferometer, the incident beam is focused by a 36x Cassegrain

    objective lens on the sample. The reflected light is collected by the same lens and

    directed to an MCT detector. A knife-edge aperture is applied in the conjugate image

    plane to define a small area containing exclusively the individual clusters of interest,

    represented by the rectangle bounded by the red dashed lines. ..................................30

    Figure 3.2: Infrared optical properties of different clusters obtained by single cluster

    FTIR spectroscopy. (a) Reflectance spectra of uncoated 2 μm silica clusters on a gold

    substrate. Measured clusters include a monomer (black), a dimer (red), a symmetric

    trimer (blue), and a symmetric heptamer (olive). (b) Reflectance spectra of 2 μm

    core-shell SiO2/Au clusters on a gold substrate. The shell thickness is about 10 nm.

    Measured clusters include a monomer (black), a dimer (red), a symmetric trimer

    (blue), a chain of three (cyan), an asymmetric quadrumer (purple), an asymmetric

    pentamer (orange), and a chain of five (green). Spectra are shifted vertically for

    clarity except for monomers. The numbers given in the figure legend denote the

    number of particles in the cluster. ...............................................................................32

    Figure 3.3: (a) Infrared optical properties of 2 μm clusters consisting of particles with a

    defective shell. The sizes of the defects on the shell are greater than 100 nm. The

    existence of large defects enable direct illumination of light on the silica core,

    evidenced by the reflection drop between 8 and 9 μm and the valley between 9 and

    10 μm (arising from the absorption of silica). Meanwhile, the line shape at shorter

    wavelengths also deviates from the measurement results in Figure 2.13(b), indicating

    undesired failure of resonant modes excitation. (b, c) SEM images of defective

    clusters measured in (a). .............................................................................................33

    Figure 3.4: Infrared optical properties of 1 μm core-shell particle clusters on a gold

    substrate. The shell thickness is about 30 nm. Measured configurations include a

    monomer (black), a dimer (red), a symmetric trimer (blue), an asymmetric

    quadrumer (purple), a chain of four (dark cyan), and an asymmetric pentamer

    (orange). Spectra are shifted vertically for clarity except for monomer. All the

    measurements were based on single cluster FTIR spectroscopy. ...............................34

    Figure 3.5: (a) Comparison of measured and simulated reflectance spectra of an uncoated

    silica monomer and an uncoated silica trimer on a gold substrate. The very good

    agreement between measurements and simulations confirms the validity of the

    simulation settings and the material properties. Slight shifts between simulations and

    measurements may come from the simplified normal incidence. Because of the

    absence of the gold shell and relatively low index of the particles, no resonance can

    be excited, as indicated by the field intensity maps at (b) 2.5 μm and (c) 3.3 μm at

    two cross sections cutting through a silica particle on the symmetry planes. ............36

    Figure 3.6: Optical properties of a 2 μm core-shell monomer on a gold substrate. The

    shell thickness is about 10 nm. (a) Comparison of measured (red) and simulated

    (black) reflectance. In the simulation, a y-polarized plane wave is incident from the

    top. Representative modes are denoted by the dashed lines. (b) Field intensity maps

    of the cavity plasmon modes and their correspondence to the spectral features. Only

    one quarter of the particle is shown because of symmetry. The electric (a11) and

    magnetic (b11) dipole modes show typical ring-like magnetic and electric field

  • x

    distributions, respectively, whereas the electric (a21) and magnetic (b21) quadrupoles

    are indicated by the two intensity lobes along the perimeter. (c) The electric field

    intensity distribution at 5.3 μm. The extremely weak field inside the particle and

    large field intensity outside the particle in the y-z plane (right panel) suggest the

    occurrence of the fundamental electric dipole resonance in response to the y-

    polarized incident field. (d) Surface charge density at 5.3 μm wavelength, which also

    manifests the excitation of the electric dipole resonance of the core-shell particle. ..38

    Figure 3.7: Comparison of simulated reflectance spectra for 2 μm (a) and 1 μm (b)

    monomers with different shell thicknesses. ................................................................40

    Figure 3.8: Optical properties of a 2 μm core-shell trimer on a gold substrate. The shell

    thickness is about 10 nm. (a) Comparison of measured (red) and simulated (black)

    reflectance. In the simulation, a y-polarized plane wave is incident from the top.

    Representative modes are marked by the dashed lines. (b) Field intensity maps of the

    cavity plasmon modes and their correspondence to the spectral features. The colored

    dashed lines in the central panels indicate the position of the cutting planes for

    observing different mode profiles. Only half of the trimer is shown because of

    symmetry. Compared with Figure 3.6, the electric dipole (a11) and magnetic

    quadrupole (b21) are still conspicuous and almost unshifted. The electric quadrupole

    (a21) becomes visible due to the interparticle coupling, while the magnetic dipole (b11)

    diminishes. (c) The electric field distribution at 7 μm wavelength. The intense field

    in the gap between two particles (left panel) suggests strong coupling between the

    electric dipole modes of the two particles. In contrast, the third particle (right panel)

    is off-resonance. (d) Surface charge density at 7 μm seen from the dimer side (left

    panel) and from the top (right panel). Both confirm the strong dipole resonances in

    the dimer. ....................................................................................................................42

    Figure 3.9: Comparison of simulated reflectance spectra for 2 μm (a) and 1 μm (b)

    trimers with different shell thicknesses. .....................................................................43

    Figure 3.10: (a) NIR reflectance spectrum of a 2 μm core-shell trimer on a gold substrate.

    Data is obtained by using a tungsten halogen lamp as the light source in combination

    with a quartz beam splitter and a MCT detector. (b) Field intensity profiles of the

    mode b31. Some of the cavity modes cannot be excited efficiently due to small shell

    thickness and symmetry requirement. ........................................................................45

    Figure 4.1: Flow chart of the fabrication of PS/Au core-shell particles. ..........................48

    Figure 4.2: Schematic (a) and experimental setup (b) for self-assembly of monolayers.

    Diluted PS particle suspension is slowly injected to the air-water interface to form a

    closely packed monolayer, which is picked up by a substrate after draining. ............50

    Figure 4.3: Optical images of the fabricated large-scale PS particle monolayers. Each

    chip is around 1.5 cm × 1.5 cm. The samples exhibit different vivid colors when

    observed from different perspectives, indicating the particles are in a closely packed

    pattern. The sizes of the PS particles are 1 μm (left), 2.1 μm (middle), and 2.8 μm

    (right), respectively. ....................................................................................................51

    Figure 4.4: Reflectance spectra of 500 nm (a) and 1 μm (b) closely packed PS particle

    monolayers. Colored solid curves are FTIR measurements, while the black dashed

    lines are simulation results..........................................................................................51

    Figure 4.5: Reflectance spectra of 2.1 μm closely packed PS particle monolayers after

    different O2 etching periods. The power is fixed at 60 watts. ....................................53

  • xi

    Figure 4.6: PS/Au core-shell particles at different fabrication stages. (a-b) Top view of a

    monolayer of 1 μm (a) and 2.1 μm (b) PS beads treated by 30s reactive ion etching.

    (c-d) After gold deposition, the particles are flipped over by PDMS stamping to have

    their uncoated side facing up. (c) Top view of the transferred 1 μm monolayer. (d) A

    second cycle of gold deposition leads to fully coated core-shell particles. A

    perspective view at 45° shows the non-spherical shape caused by anisotropic etching.54

    Figure 4.7: Individual 2.1 μm (a) and 2.8 μm (b) PS/Au core-shell monomers on a gold

    substrate. The orientations are random after releasing. ..............................................55

    Figure 5.1: Schematic of the nanoantenna consisting of a gold nanoparticle trimer. The

    third particle (diameter D2 = 120 nm) and the dimer (diameter D1 = 100 nm) are

    slightly different in size, forming an isosceles triangle. An electric dipole (ED)

    emitter is positioned at the center of the dimer and is oriented along the y-axis. The

    dimer gap size is fixed to be g1 = 5 nm, while the larger particle can be displaced

    along the x-axis. The displacement h is counted relative to the equilibrium position

    (top), where the monomer-dimer gaps (g2) are also 5 nm. By moving the larger

    particle closer to the dimer (h0,

    bottom), the radiation direction is switched to −x or backward direction. .................60

    Figure 5.2: Results for the forward-radiation configuration. The larger particle is moved

    4 nm towards (+x direction, h = −4 nm) the dimer from the equilibrium position. (a)

    Spectra of the total radiated power (green) and of the ED (blue) and MD (red)

    amplitude, respectively. (b, c) Surface charge density (left) and electric field

    distribution in vector format (right) at (b) 845 nm and (c) 720 nm wavelengths,

    where the magnetic and electric resonances take place respectively. The minimum

    side lobe level of −22 dB for forward radiation is observed at 857 nm, where the ED

    and MD have approximately equal amplitude and are nearly in phase. (d) 3D and (e)

    2D radiation patterns at 857 nm. .................................................................................63

    Figure 5.3: Evolution of the amplitude (a) and phase (b) of the dipole modes. The

    numbers in the legend denote the displacement h of the larger particle. ....................64

    Figure 5.4: Results for the backward-radiation configuration. The larger particle is

    moved 4 nm away (−x direction) from the dimer counting from the equilibrium

    position. (a) Spectra of the radiated power (green) and of the ED (blue) and MD (red)

    amplitude, respectively. (b) Surface charge density (left) and electric field

    distribution (right) at 710 nm, where the ED and MD mode are both on resonance. (c)

    Same as (b) but at 780 nm, where the minimum side lobe level of −4.2 dB for

    backward radiation is achieved. (d) 3D and (e) 2D radiation patterns at 780 nm. .....66

    Figure 5.5: Minimum side lobe level that can be achieved with different trimer

    configurations (given by displacement h). Circled dots correspond to the results

    reported in Figures 5.2 and 5.4. Note that for different configurations, the best

    directionality occurs at different wavelengths. ...........................................................67

    Figure 5.6: Switching of forward-/backward radiation at the same wavelength with a

    reconfigurable nanoantenna (a) or at different wavelengths with a fixed structure (b).68

    Figure 5.7: Spectra of the Purcell factor Fp (red), emission enhancement f (blue), and

    quantum yield η (green) for (a) forward-radiation configuration when h = −4 nm, and

    (b) backward-radiation configuration when h = 4 nm, respectively. ..........................69

  • xii

    Figure 5.8: Radiation patterns reconstructed from the ED-MD interference model. (a)

    Forward radiation at 857 nm (left, cf. Figure 5.2(e)). (b) Backward radiation at 780

    nm (right, cf. Figure 5.4(e)). .......................................................................................71

    Figure 5.9: Comparison of normalized radiation patterns under different excitation

    conditions. In non-ideal cases, the emitter is either shifted or tilted. (a) Forward

    radiation at 857 nm wavelength with h = −4 nm. (b) Backward radiation at 780 nm

    wavelength with h = 4 nm. .........................................................................................72

    Figure 6.1: (a) Schematic of a metasurface consisting of subwavelength plasmonic

    resonators, which function as an array of local sources with relative phase and

    polarizability difference. Manipulation of wavefront, polarization and intensity of the

    reflected/transmitted light can be achieved by designing the optical response and

    arrangement of the resonators. (b) Generation of orbital angular momentum by a

    metasurface. The left panel shows the SEM image of the metasurface, which uses

    plasmonic nanoantennas to realize 2π phase distribution in the azimuthal angle

    direction. The right panel is the measured spiral patterns created by the interference

    of the vortex beam and a co-propagating Gaussian beam. (b) is adopted from Ref.

    [97]. .............................................................................................................................76

    Figure 6.2: Schematic of the reflective graphene metasurface consisting of

    subwavelength graphene ribbons on a dielectric/metal substrate. (a) By properly

    arranging the ribbon widths and choosing the material and thickness of the spacing

    layer, the resultant plasmonic resonance in each ribbon varies and the reflected IR

    light can be controlled in a prescribed manner. (b) The graphene metasurface forms

    an asymmetric Fabry-Perot resonator, which substantially enhances the interaction

    between light and graphene nanostructures on the top. ..............................................78

    Figure 6.3: (a) Reflectivity and (b) relative phase of the reflected light as the functions of

    ribbon width (w) and dielectric layer thickness (s). (c) Reflectivity and phase of the

    reflected light with varying ribbon widths, while the thickness of the dielectric layer

    is fixed to be s = 4.37 μm. The inset shows in cross-section the field distribution of

    the plasmonic resonance at a graphene ribbon when w = 1.035 μm. Hereafter unless

    specified, the frequency of the incident light is 12.32 THz, the periodicity is 3 μm,

    the refractive index of the dielectric layer is 1.4, and the Fermi energy of graphene is

    0.64 eV. .......................................................................................................................81

    Figure 6.4: Demonstration of anomalous reflection with (a) normal incidence, and (b)

    oblique incidence at an angle of 15.0°. Hereafter in this chapter, unless specified, the

    incident field is subtracted from the total field for clarity, and the amplitude of the

    reflected field is normalized to that of the incidence. Top: Diagrams of rays

    indicating the incident and reflected light. Bottom: Electric field distributions. ........83

    Figure 6.5: Schematic of achieving reflective focusing in (a) bulk optics and (b) by a

    metasurface. ................................................................................................................84

    Figure 6.6: Demonstration of a reflective focusing lens at three different frequencies: (a)

    12.32 THz, (b) 11.82 THz, and (c) 12.82 THz. ..........................................................85

    Figure 6.7: A dynamically tunable reflective focusing lens by varying the Fermi energy

    of graphene. (a) Ef = 0.56 eV and (b) Ef = 0.48 eV. ...................................................85

    Figure 6.8: Generation of Airy beams. (a) Ideal phase and amplitude profile at the

    emitting plane to generate an Airy beam. (b) Electric filed intensity map of an ideal

  • xiii

    Airy beam resulting from (a). (c) Simplified Airy function. (d) An Airy beam

    generated by a graphene metasurface following the profile in (c). ............................87

    Figure 6.9: Schematic of a graphene metamirror that strongly reflects right circularly

    polarized (RCP) light with preserved helicity but weakly reflects left circularly

    polarized (LCP) light, giving rise to large differential absorption or circular

    dichroism (CD). The inset illustrates the unit cell consisting of a pair of “LI”-shaped

    graphene patches. The periodicities in the x and y direction are 250 and 120 nm,

    respectively. Other geometric parameters are: Lx = 65 nm, Ly = 100 nm, w = 50 nm,

    and g = 50 nm. The CaF2 spacing layer thickness t = 1550 nm. ................................89

    Figure 6.10: Helicity-selective reflection of the graphene metamirror with the carrier

    mobility μ = 10000 cm2/V/s (a) and μ = 7500 cm

    2/V/s (b). The Fermi energy is fixed

    to be 0.8 eV. At the resonance frequency, RCP light is largely reflected with

    preserved helicity (green curve), while LCP light is strongly absorbed showing much

    lower reflection (blue curve). The cross-polarization conversion is weak and

    identical for both polarizations (overlapping red solid and black dashed lines). .......90

    Figure 7.1: Magnetic field assisted assembly of linear chains consisting of 2 μm

    polystyrene particles in a ferrofluid. (Images taken by Rasam Soheilian in the DAPS

    Lab) .............................................................................................................................95

    Figure 7.2: Schematic of the magnetic-field assisted self-assembly of programmable

    metamolecules. (a) In bulk fluid, core-shell microparticles are assembled into certain

    geometries by changing the external magnetic fields. (b) By using predefined

    magnetic templates on a substrate, core-shell particles can be settled by external

    magnetic fields with a global order. ...........................................................................95

    Figure 7.3: SEM images of fabricated nanorod pair array by electron beam lithography,

    aimed at testing the fabrication limit for potential templates preparation. .................96

    Figure 7.4: (a) Fully assembled cell for FTIR spectroscopy in aqueous environment.

    CaF2 slides and the spacer are sandwiched between the aluminum nest plates. Core-

    shell particle suspension is injected from one of the two ports. The IR transparent

    window (orange box) allows both observation and measurement. (b) Comparison of

    transmission through the cell filled by water (red) and by ferrofluid EMG 705 (blue),

    respectively. ................................................................................................................97

  • xiv

    Acknowledgements

    At this point of approaching the very end of my graduate study, a lot of memorable

    moments come out in my mind. During the past four and a half years, I have received so

    much invaluable help and support from my colleagues, friends, and family.

    First of all, I would like to thank my advisor, Prof. Yongmin Liu, for choosing me as

    his first student—this really saved me from a hard time deviating from my dream

    career—and for his patient guidance and unreserved support throughout all these years.

    The privilege of being the only audience of his lectures on the fundamentals of

    plasmonics and metamaterials could never happen again to the later group members, and

    either will I forget the wonderful discussions on the latest literatures. The most important

    thing I have learned from him is the ambition to keep chasing tough but sweeping

    research topics, even if they appear impossible at the very beginning. The lessons on time

    management and on establishing collaborations also influenced me a lot.

    I would also like to express my grateful thanks to all the committee members: Prof.

    Matteo Rinaldi, Prof. Randall Erb, and Prof. Nian. X. Sun for their helpful guidance,

    inspiring advice, and kind support. Before starting the PhD I had no prior knowledge on

    nanofabrication or chemistry. It was Matteo’s course that opened the door of

    nanofabrication to me and introduced me how everything works out. In the collaborative

    work on reconfigurable metamaterials, where I even needed to start from learning using a

    pipette, I received numerous suggestions from Randy on assembly, surface chemistry,

    magnetism and so forth. When there were any unplanned requests on metal deposition,

    Nian always kindly gave me permission to use the facilities in his lab, even they already

  • xv

    had a tight schedule. Without the support from any of you, it would be impossible for me

    to present most results in this dissertation.

    Practicing research in the multi-disciplinary field is highly collaborative. Here I

    would give some special thanks to the colleagues who had selflessly shared their time,

    thoughts, and expertise to promote my research. I acknowledge Yu Hui and Fangze Liu

    for their supervision on SEM and ELB, and thank Zheng Liu for spending so much time

    with me polishing these skills. I greatly appreciate Liao Chen, who helped me out with

    almost all the troubles in the chemical synthesis, and Shuai Chen, who assisted me to

    overcome the initial hesitation in performing chemistry. It was a great pleasure to work

    with Rasam Soheilian on the project of magnetic self-assembly. I have bothered him too

    much while he was busy. Hope this collaboration could become fruitful in the near future

    and the friendship lasts a lifetime. I am also very grateful to Qing Hu for her help on

    fabrication, to Xuefei Tan for her training on FTIR, and to Scott McNamara, David

    McKee, and Sivasubramanian Somu for their continuous support on using the facilities.

    I thank the rest of the Liu group, including but not limited to Zubin Li, Hui Jia, Zuojia

    Wang, Feng Cheng, Anna Nardella, Xiaokang Bai, Wei Ma, Lin Li, Zhong Huang, and

    Han Gao. And there is even a longer list of external colleagues and collaborators, such as

    Ji Hao, Xinjun Wang, Zhenyun Qian, Ryan Sungho Kang, Ismail Ebraheim, William

    Fowle, Wentao Liang, Abdelkrim Benabbas, etc.

    Last and most importantly, my deepest appreciation goes to my parents. Their love

    and understanding supported me not only these years but all my life so far. This work is

    dedicated to them.

  • 1

    Chapter 1

    Introduction

    Metamaterials are rationally engineered composite materials with exotic properties

    not easily obtained or altogether unavailable in nature [1]. Since the early 2000s,

    metamaterials have emerged as a new frontier of science involving physics, material

    science, optics, engineering, nanotechnology and so forth, mainly because they provide

    unprecedented opportunities to control light-matter interactions and light propagation in

    unusual ways. The exceptional properties of metamaterials arise from the elegant design

    of subwavelength resonators or “meta-atoms” as well as judicious arrangement of these

    building blocks (see Figure 1.1). This is distinctly different from natural materials, whose

    properties are primarily determined by the chemical constituents and bonds [2].

    Figure 1.1: Schematics of some representative metamaterials. (a) A lattice of metallic

    wires to achieve negative permittivity at low-frequency. (b) Metallic split ring resonators

    (SRRs) to produce artificial magnetism. (c) Metallic helices as building blocks for chiral

    metamaterials.

    Early proof-of-concept experiments of metamaterial were mainly done in the

    microwave regime [3]. The operating frequency was quickly pushed higher from

  • 2

    terahertz (THz) to infrared (IR) and then to visible [4]. Because most metamaterials are

    composed of metals, the plasmonic effect of metals plays an important role in optical

    metamaterials.

    Plasmonics focuses on the unique properties and applications of surface plasmon

    polaritons (SPPs), the quasiparticles arising from the strong interaction between light and

    free electrons in metals [5-7]. At the interface between a semi-infinite metal and a semi-

    infinite dielectric, SPPs behave as a surface wave propagating along the interface while

    exponentially decaying into both sides, as illustrated in Figure 1.2.

    Figure 1.2: (a) Illustration of SPPs and oscillation of surface charges at a dielectric/metal

    interface. (b) The typical electromagnetic field profile of SPPs, reflecting their bound,

    non-radiant nature.

    The interplay between metamaterials and plasmonics has led to tremendous efforts

    devoted to the invention and applications of plasmonic metamaterials in the optical

    region [8, 9]. However, a few areas under this research topic are not fully explored, and

    some restrictions limiting its further development are yet to be overcome. For instance,

    infrared light has found widespread applications such as biomedical sensing,

    spectroscopy, and surveillance [10]; while on the other hand, metals do not support

    pronounced plasmonic resonances at mid-infrared (MIR) and far-infrared (FIR)

    frequencies as in the visible and near-infrared (NIR) region. The lack of suitable

  • 3

    platforms that can provide strong light-matter interactions at infrared wavelengths, as a

    consequence, naturally leads to the quest of better alternatives for infrared plasmonics

    and metamaterials.

    Another restriction is the tunability. Up to date, most metamaterials and plasmonic

    devices are in solid phase and have fixed functionalities once they are fabricated. Since

    the optical properties of metamaterials and plasmonic nanostructures rely on both the

    constituents and the way in which they are shaped and arranged, introducing variables

    into the system, either changeable material properties or reconfigurable architectures, will

    open up new opportunities for dynamically tunable metamaterials and plasmonics [11].

    The recent progress towards this target is exciting. Figure 1.3 summarizes representative

    attempts that employ materials with tunable dielectric properties. To this end, phase-

    change materials are among the most promising candidates [12, 13]. It has been shown

    that the refractive-index-changing phase transition of chalcogenide glass (GST) can be

    induced optically by trains of femtosecond laser pulses [13]. Hence, despite the

    achievable refractive index is limited, it is possible to write arbitrary patterns on the GST

    film to perform functionality such as lensing or erase them, both with light (Figure 1.3(a-

    c)). Vanadium oxide (VO2) is also popular as a phase-change material [14, 15]. Figure

    1.3(d) and (e) shows a design of metamaterial absorber, which utilizes electrical currents

    to induce Joule heating and in turn the phase transition of the VO2 film [14]. Because the

    infrared reflection of the devices is sensitive to the index change, applying different

    currents allows turn-on/off and continuous contrast tuning of the infrared image. Another

    important class of materials that has tunable dielectric properties is the semiconductor. In

    particular, two-dimensional (2D) materials such as graphene can be dynamically tuned on

  • 4

    Figure 1.3: Reconfigurable metamaterials and metadevices by material property changes.

    (a) Functional devices can be written in or erased from a chalcogenide glass (GST)

    phase-change film by trains of femtosecond laser pulses at different conditions. (b) The

    optical image of two Fresnel zone patterns written by light. (c) The second Fresnel zone

    pattern on the right is erased by light. (d) A hybrid metamaterial consisting of a vanadium

    oxide (VO2) film sandwiched by gold meshes and a ground plane. Applying electrical

    currents will induce the insulator-to-metal transition of VO2 and change the reflectance.

    (e) Top left: Microscopic image of the metamaterial. Letters are bare VO2 areas without

    gold meshes. Rest panels in clockwise order: Infrared images at 2.67 μm wavelength with

    0 A, 2.03 A, and 2.20 A current, respectively. (f) An infrared light beam excites tunable

    plasmon resonances across the graphene nanoribbons biased at varying gate voltages. (g)

    Plasmon-enhanced light-matter interactions improve detection sensitivity of molecules

    adsorbed on graphene. (a-c) are adopted from Ref. [13], (d, e) are adopted from Ref. [14],

    (f) is adopted from Ref. [17], and (g) is adopted from Ref. [18], respectively.

  • 5

    their dielectric properties via electrostatic gating [16]. Figure 1.3(f) and (g) shows such

    an example, in which graphene is patterned into nanoribbons to provide a platform for

    plasmon-enhanced MIR biosensing [17-19].

    The other side of the solution to reconfigurable metamaterials is altering the structure.

    As the coupling between the constituent parts is critical in determining the global optical

    properties, introducing finite structural changes is sufficient to induce notable variations

    in the spectra and field distributions. In Figure 1.4(a)-(d), two prototypes of gap-tuning

    metamaterials are presented. The first one is fabricated on an elastomeric stamp; thus

    stretching the substrate will enlarge the gap size and shift the resonance [20]. In the

    second example, similar in-plane mechanical deformation can be induced by applying

    electrical voltages to the supporting strings [21]. The associated attractive forces are in

    the range of a few nanonewtons, providing a very fine control on the delicate structures.

    It will not be too ambitious to imagine that the ultimate reconfigurable metamaterials

    or metamolecules should possess the capability to actively transform between distinct

    structures under a regulated control. This reorganization of architecture will significantly

    extend the bounds of the material properties available so far. The path to the Holy Grail is

    rather rocky, and the attempts with self-assembly techniques appear to be most promising.

    Figure 1.4(e) and (f) illustrates two types of reconfigurable chiral metamolecules enabled

    by the DNA stapling [22], where upon the addition of specifically designed DNA strands,

    the relative position of the paired nanorods varies under control [23, 24]. The resulting

    states of the metamolecules correspond to different chirality, which can be readily

    distinguished in the spectra in Figure 1.4(g). Although the structural changes are still

    finite, topologically the metamolecules at each state are different.

  • 6

    Figure 1.4: Reconfigurable metamaterials and metamolecules by finite structural changes.

    (a) Split ring resonator (SRR)-bar on a stretchable film. Mechanical deformation of the

    elastomeric substrate changes the gap size and thus the coupling, which causes noticeable

    resonance shift in (b). (c) A metamaterial consisting of “meander near the wire” gold

    nanostructures on silicon nitride strings. (d) Applying voltages to the neighboring strings

    induces an attractive electrostatic force in the plane, which reduces the gap size, enabling

    modulation of transmission and reflection. (e, f) Two gold nanorods are hosted on a DNA

    template to form a chiral metamolecule exhibiting tunable chiroptical properties as in (g).

    (e) The relative angle between the nanorods can be actively controlled by adding certain

    DNA strands to pair the locks on the sides of the template. (f) Alternatively, the nanorods

    are assembled perpendicular to each other, and the DNA locks function as binding sites

    that allow the rod on the top (yellow) to “walk” stepwise between the two ends of the

    template. (a, b) are adopted from Ref. [20], (c, d) are adopted from Ref. [21], (e) is

    adopted from Ref. [23], and (f, g) are adopted from Ref. [24].

  • 7

    In this work, I explore the possible solutions for achieving infrared reconfigurable

    plasmonics and metamaterials. Simulations and experiments are both conducted to study

    the different aspects of this topic, from fundamentals to applications, from near-infrared

    to far-infrared, and from static material property tuning to finite structure altering and

    eventually towards the ultimate architecture programming. The contents of this work are

    organized in the following order:

    The first two consecutive chapters, namely Chapters 2 and 3, explore in depth the

    optical properties of SiO2/Au core-shell particles and their self-assembled clusters for the

    application of mid-infrared resonators. In Chapter 2, based on wet chemistry, high quality

    core-shell particles are synthesized and the conditions leading to different surface

    morphologies are discussed. Self-assembly is then carried out to obtain clusters of varied

    configurations.

    The optical characterization of the core-shell particles and associated numerical study

    are presented in Chapter 3. Infrared spectroscopy is performed to measure the optical

    properties of the particles at the single cluster level. It is found that depending on the

    number and configuration of the constituent particles, each cluster exhibits a unique

    spectrum line shape at the infrared, especially mid-infrared wavelengths. Full-wave

    simulations are conducted to reproduce the measured spectra. In addition to the excellent

    agreement, clear correspondence is established between the spectral features and various

    resonance modes. The systematic study in this chapter undoubtedly shows that plasmonic

    core-shell particle can be tailored to achieve tunable resonances.

    Chapter 4 discusses an alternative method to make core-shell particles. Based on

    microfabrication techniques, polystyrene/Au core-shell particles are produced with a high

  • 8

    yield. Self-assembled monolayers (SAMs) of polystyrene (PS) particles are realized first

    as a template. Plasma etching is applied to reshape the particles. Mediated by a transfer

    process, the dielectric cores are coated by two cycles of deposition, resulting in another

    type of designer infrared resonators.

    In Chapter 5, an ultra-compact plasmonic nanoantenna with switchable directionality

    is proposed and systematically studied by simulations and analyses. The antenna

    comprises an asymmetric gold nanoparticle trimer and is fed by a nano emitter in the gap.

    When the so-called “Kerker conditions” are satisfied, the antenna shows superior

    radiation directionality in the near-infrared region. Moreover, by displacing one of the

    three particles for less than 10 nm, the radiated light beam can be switched to the

    opposite direction. Analyses based on a dipole interference mdoel are performed to

    understand the evolution of the radiation patterns. Techniques for potential experimental

    demonstrations are also discussed.

    In Chapter 6, several designs of tunable flat lenses based on graphene metasurfaces

    are proposed to steer far-infrared light. Practical functionalities, such as anomalous

    reflection, reflective focusing, generation of non-diffracting beams and strong chiroptical

    effects are demonstrated by full-wave simulations. Because the dielectric properties of

    graphene can be tuned via chemical doping or electrostatic gating dynamically, the

    possibility of tuning the focal length of a reflective metalens is also numerically proven

    by changing the Fermi energy. These results expand the applications of graphene-based

    devices towards complete beam shaping in the infrared region.

  • 9

    Finally, following a brief summary, an outlook and some preliminary results on the

    ongoing study of programmable plasmonic metamolecules via magnetic self-assembly

    and their sensing applications are presented in Chapter 7.

  • 10

    Chapter 2

    Plasmonic Core-Shell Particles by Chemical

    Synthesis

    2.1 Introduction

    The use of subwavelength structures to achieve exotic control of electromagnetic

    fields has stimulated the blossom of metamaterials, plasmonics, and nanophotonics over

    the past decades [9]. The cornerstone of these fields lies on the fact that micro- or nano-

    scale building blocks can resonantly interact with light waves in a predefined manner by

    programmatically designing their structures and arrangements. Metallic structures play a

    crucial role in building resonators operating in the visible and NIR region, predominantly

    because of their capability to support SPPs. The collective oscillation of free electrons in

    metals can couple efficiently with light, which ensures strong light-matter interactions

    necessary for further construction and engineering of metamaterials. However, using

    metallic elements as building blocks for metamaterials is accompanied by several

    limitations, among which the most influential one is the loss. Ohmic losses in metals exist

    across the entire visible spectrum and remain considerable in the infrared region, whereas

    the excitation efficiency of surface plasmons decreases rapidly as the wavelengths

    deviating from the visible. As a consequence, metals are not ideal platforms for NIR

    plasmonic devices and are seldom used to implement resonators at MIR wavelengths.

  • 11

    With the urgent quest of infrared optical applications, suitable building blocks for

    infrared optics are in a pressing need [25, 26].

    Recently, high-refractive-index semiconductors such as silicon, germanium, and lead

    telluride are proposed as alternative platforms for infrared metamaterials and photonic

    devices [27-31]. As shown in Figure 2.1(a), due to the large refractive index and low-loss

    properties, individual subwavelength particles can interact with infrared light strongly via

    a series of Mie resonances. In addition, the opportunity of tuning the dielectric constants

    by doping [30] or by thermal effects [31] makes these materials very attractive compared

    with conventional dielectrics. However, the fabrication of regular and uniform micro- or

    nano-particles is challenging for high index semiconductors. Laser ablation can create

    small amounts of particles from various substrates but lacks control of the particle size

    and roundedness (see insets of Figure 2.1(a) and (b)) [31-33]. Chemically synthesized

    amorphous silicon particles show improved size uniformity, but they are still far from

    being identical, hindering the acquisition of strictly symmetric clusters [34]. Furthermore,

    compared with their metallic counterparts, dielectric particles generate less intense local

    fields and possess weaker coupling when they are tightly packed [35]. The former issue

    limits the light-matter interaction strength, which is critical in numerous applications such

    as sensing and spectroscopy; the latter restricts the spectral tunability of the cluster,

    because introducing additional particles into an existing cluster will not cause significant

    changes to its optical properties [33]. In Figure 2.1(b), although the number, size, and

    configuration of the particles vary a lot, the reflection curves only show limited changes

    in the line shape. Therefore, the task of seeking suitable building blocks for infrared

  • 12

    plasmonics and metamaterials with notable spectral tunability is still unsolved by using

    dielectrics alone.

    Figure 2.1: Optical resonators at different wavelengths. (a) A high-index dielectric (lead

    telluride (PbTe)) particle can support a series of Mie resonance modes in the mid infrared

    (MIR) region. (b) High-index dielectric (silicon) particle clusters exhibit limited spectral

    tunability at visible frequencies. (c) Plasmonic core-shell nanoparticle clusters show

    complex spectral features in the near infrared (NIR) region. Figures are adopted from

    Refs. [31], [35], and [39], respectively.

    Core-shell particles consisting of a dielectric core and a metallic shell offer a

    promising solution. Such core-shell structures can support sophisticated, hybridized

    plasmonic modes as well as cavity modes [36-38]. When they are closely packed into

    clusters, strong interparticle coupling arises, resulting in complex near-field distribution

    and far-field scattering patterns [39, 40]. It has been demonstrated that magnetic

    resonance and Fano resonance can be excited in core-shell nanoparticle clusters at large

    incidence angles (Figure 2.1(c)). However, the sizes of nanoparticles determine that these

    resonances occur only in the NIR region, and cavity modes can hardly be excited or

  • 13

    utilized. In this chapter, the chemical synthesis of micron-sized plasmonic core-shell

    SiO2/Au particles (SiO2/Au denotes silica core with gold shell) is discussed in detail. The

    resulting particles are uniformly coated and almost identical, enabling the preparation of

    highly symmetric core-shell clusters by subsequent self-assembly process. The infrared

    optical characterization of individual clusters and the theoretical study on their optical

    properties will be discussed in the next chapter.

    2.2 Synthesis of Plasmonic Core-Shell Particles

    The chemical synthesis of SiO2/Au core-shell particles involves a series of steps. The

    general flow is illustrated in Figure 2.2. For all the experiments, the chemicals are

    purchased from Sigma-Aldrich without further modification. Although silica nano- and

    micro-spheres can also be synthesized in the lab following the standard procedure

    developed by Stöber et al. [41], commercially available products are used for better size

    and shape control. Because the main focus of this work is to explore the optical properties

    of plasmonic core-shell particles as infrared resonators, the associated silica core particles

    are chosen to be 1 and 2 μm in size. Smaller particles resonate in the visible region rather

    than in the infrared, whereas larger particles may not fulfill the requirement for potential

    applications for infrared metamaterials. Other necessary chemicals include: chloroauric

    acid (gold(III) chloride or HAuCl4·3H2O, 99%), tetrakis(hydroxymethyl)phosphonium

    chloride (THPC, 80% aqueous solution), sodium hydroxide (NaOH), potassium

    carbonate (K2CO3), sodium chloride (NaCl), 3-aminopropylthiethoxysilane (APTES,

    98%), formaldehyde (H2CO, 37%), and 200 proof ethanol. This protocol is adapted from

    an earlier work by Brinson et al. [42, 43]. Mainly caused by the large sizes of the

    particles used in this study, the amount of reactants in many key steps needs to be

  • 14

    modified, which nevertheless does not follow the scaling law simply. Meanwhile, how

    the way of introducing gold ions and reducing agent will affect the shell completeness

    and surface morphology is also investigated, which appears less sensitive for small

    particles.

    Figure 2.2: Flow chart of the synthesis of SiO2/Au core-shell particles. (a) Synthesis of

    gold nanoparticle seeds (i) and modification of SiO2 particles (ii). (b) Preparation of

    precursor particles, where gold nanoparticles are attached to the surface of the modified

    silica beads. (c) Continuous reduction of gold ions in the planting solution causes the

    nanoparticles to increase in size, coalesce, and eventually form complete shells covering

    the silica cores.

    During a single synthesis cycle, aging takes quite a large portion of time. Therefore,

    despite several sections and associated preparations are parallel in the flow chart, careful

    considerations in arranging the sequence of experiments can be very helpful to shorten

    the synthesis period. In the following, the experimental details will be introduced in an

    optimized order by assuming all the chemicals are in the as-arrived state.

    First, 1 g of gold(III) chloride powder is dissolved in deionized water (referred to DI

    water or simply water herein) to produce a 1% (by weight) chloroauric acid solution. The

  • 15

    dissolving and subsequent storage should both be in a glass container, and the final

    solution is stored in the dark for at least 48 hours before use.

    Second, the silica particles are functionalized with APTES. After diluting the as-

    purchased suspension to 1 wt%, the treatment starts by washing the particles thoroughly

    three times via centrifugation for 30 minutes, complete removal of the supernatant, and

    redispersion in water under strong sonication. These operations are intended to detach

    undesired stabilization groups on the particle surface. Following the last washing cycle,

    the particles are redispersed in ethanol and washed for additional three times. After that,

    the particles are resuspended in ethanol, placed in a polypropylene bottle and stirred

    rapidly. The volume of ethanol is equal to the diluted 1 wt% aqueous suspension before

    washing (4 mL in my experiments). The modification is carried out in a glovebox filled

    with nitrogen. Fresh APTES is added to the suspension with continuous stirring for 10

    minutes. The amount of APTES can be calculated by:

    30[μL] / 4[mL] 126[nm]/ ,APTESV V D (2.1)

    where V is the volume of suspension and D is the diameter of the particles. The amount

    of APTES from Equation (2.1) is about five times the amount needed to form a silane

    monolayer (~0.7 nm thick) covalently attached to the silica surface. The actual thickness

    of the silane film varies from 1 to 3 nm, corresponding to several monolayers. However,

    it is also noticed that in my experiments, successful synthesis of high quality core-shells

    is achieved even with wild excess of APTES. The impact of the silane film thickness on

    the properties of core-shell particles may need further study. The suspension is sealed and

    stirred slowly overnight. Subsequent operations can be done in the open lab environment.

    The suspension is boiled in a borosilicate container for 2 hours. Total evaporation is

  • 16

    prevented by continuous addition of ethanol. Care should be taken to prevent bumping.

    The particles are then washed three times in ethanol by centrifugation, followed by

    redispersion in a proper volume of ethanol (2 mL in my experiments). The final

    suspension is stored at 4 °C and has a long shelf life of one year or longer. After

    functionalization, the surface of silica particles is coated by amino groups, which allows

    attachment of gold nanoparticles to be prepared in the next step.

    Gold nanoparticles are synthesized following the procedures developed by Duff et al.

    [44]. 1.2 mL of THPC is diluted to 100 mL, and 1 M NaOH solution is prepared by

    dissolving a NaOH pellet in a certain amount of water. Under rapid stirring to a strong

    vortex, 1.2 mL of the 1 M NaOH solution is added into 180 mL of water in a glass beaker,

    Figure 2.3: Typical ultraviolet-visible (UV-Vis) extinction spectra of the THPC gold

    nanoparticle solution. Different curves are collected from different batches of samples.

    All of them exhibit a vague bump close to 500 nm wavelength, indicating the sizes of the

    synthesized nanoparticles are 1-3 nm. Inset shows the image of an aged solution, which is

    dark brown in color.

    followed by addition of 4 mL of diluted THPC. After 5 minutes of continuous stirring,

    6.795 mL of 1% HAuCl4 solution is poured into the beaker in a quick motion. Within a

    second, the solution turns from colorless to medium brown, indicating preliminary

  • 17

    success of the reaction. After 10 minutes of stirring, the resulting solution is aged at 4 °C

    in the dark for 14 days, during which its color changes to dark brown gradually. Because

    the optical extinction is very sensitive to the nanoparticle size, any deviation of color

    from above may signify potential failure, which will cause quick settlement of the

    precursor particles in the later step. Figure 2.3 compares the ultraviolet-visible (UV-Vis)

    extinction spectra from three batches of samples. As can be seen, although the curves

    offset slightly from each other, they follow the same trend consistently and exhibit a

    weak plasmon feature: a broad and vague bump close to 500 nm wavelength. According

    to transmission electron microscopy (TEM) inspection, the mean size of the gold

    nanoparticles is between 1 and 3 nm. Larger particles giving stronger plasmonic

    responses will cause a red-shifted and more distinct extinction peak, rendering the

    solution purpler in color. The THPC gold solution has a shelf life of several months and

    can be used reliably at least four months after synthesis according to my experiments.

    Gold planting solution is prepared next by adding 300 μL of 1 wt% HAuCl4 solution

    into 20 mL of 1.8 mM K2CO3 aqueous solution. The latter is prepared by dissolving 50

    mg of K2CO3 powders in 200 mL of water. The final planting solution has a pale yellow

    color and should be aged at 4 °C in the dark for 2-3 days before use, while its shelf life

    has not been tested.

    Once the THPC gold solution is ready for use (sufficiently aged), the preparation of

    precursor particles can be performed. This is the last step before the final core-shell

    formation and is the trickiest one. The precursor particles consist of a functionalized

    silica core and uniformly distributed gold nanoparticles attached on its surface. The

    attachment is naturally driven by the electrostatic force, since the amino groups on silica

  • 18

    are positively charged whereas the gold nanoparticles are negatively charged. On the

    other hand, the Coulombic repulsion between nanoparticles will limit the achievable

    surface coverage. Radloff compared two methods to increase the ionic strength of the

    solution and thereby decrease the repulsive force for better coverage [45]. I adopt the

    more convenient one in my experiments, which requires only addition of an inert salt

    (NaCl) solution. To do this, 1 M NaCl aqueous solution is prepared by dissolving 0.584 g

    NaCl powder in 10 mL of water. Then, 40 mL of aged THPC gold solution is sonicated

    for 1 minute, followed by addition of 4 mL of the NaCl solution. 300 μL of the

    functionalized silica particle ethanol suspension is added immediately after that. The final

    solution is stored in the dark after additional sonication for 2-3 minutes. Two findings

    different from or not mentioned in previous literatures are found in my experiments. First,

    although highly concentrated NaCl can effectively increase the ionic strength of the

    solution, a threshold seems to exist for a certain size of the silica beads. According to my

    trials, 2 M NaCl already has a high risk to cause aggregation and settlement of the 2 μm

    particles in a few hours. Therefore, 1 M concentration is used for all my experiments.

    Second, the reliable period for the solution to reach equilibrium is about 72 hours, much

    longer than the reported values (12-20 hours). This is probably because of the larger

    particle sizes. The overall surface coverage of the silica particles is estimated to be

    around 25-30% according to literature. However, direct evidence from TEM is not

    practical due to the large particle size. The resulting precursor particles are washed 2-3

    times in water and then redispersed in 5 mL of water. The solution is stored at 4 °C and

    has a shelf life of about 3 months. Beyond this period aggregation may occur.

  • 19

    Figure 2.4: Scanning Electron Microscope (SEM) images of a 1 μm (a) and a 2 μm (b)

    core-shell particle (monomer) on a gold-coated substrate. The shell thicknesses are

    estimated to be around 30 and 10 nm, respectively.

    The final step of synthesis is the core-shell formation. With the help of a reducing

    agent, the attached THPC gold nanoparticles function as “seeds” or nucleation sites, upon

    which the gold ions in the planting solution are continuously reduced. Depending on the

    ratio of gold ion to precursor particle, core-shells with different surface morphologies can

    be readily synthesized under control. The first example is plasmonic core-shell particles

    with complete coating, as shown in Figure 2.4. Typically, 200 μL of precursor particle

    suspension is placed in a glass vial, followed by addition of 1 mL of planting solution

    under rapid stirring. 2.2 μL of formaldehyde is then added into the mixture in a quick

    motion. With continuous stirring, the color of the solution turns from light brown

    gradually to light purple or pink within 10-15 minutes. The ultimate color does not

    correspond to the plasmonic response of the micron-sized core-shell particles but is

    determined by that of the excess THPC gold nanoparticles, which are unavoidable in the

    precursor particle solution. Upon successful synthesis, the silica particles are uniformly

    coated by a gold shell, while the surface is quite smooth, showing only occasional cracks

    such as pinhole defects and attached gold nanoparticles. From high magnification

  • 20

    scanning electron microscope (SEM) images (e.g. Figure 2.5(a)), it is clearly seen that

    these imperfections are not larger than 40 nm, much smaller than the particles and the

    infrared wavelengths. Therefore, their influence on the optical properties will be

    negligible.

    Figure 2.5: Zoom-in SEM images for shell inspection. (a) The surface of a 2 μm particle

    coated by a complete and smooth gold shell. Occasional cracks such as pinhole defects

    and excess gold nanoparticles are smaller than 40 nm and have negligible impact on the

    infrared optical properties of the core-shell particles. (b) Perspective view (~70°) of a 2

    μm core-shell particle at a rectangular cut on its surface. The cut is made by focused ion

    beam (FIB). Measurements at different locations on the cutting edges suggest a shell

    thickness of ~10 nm.

  • 21

    The determination of the exact shell thickness, however, becomes difficult. Typically

    for nanoparticles that allow high energy electrons to penetrate through, transmission

    electron microscopy (TEM) is a standard technique for shell inspection. In the present

    study, the large particle size impedes the use of TEM while the tiny pinhole defects can

    only provide a rough estimate of the shell thickness. In Figure 2.5(b), by making a

    through cut on the surface by focused ion beam (FIB) milling, the shell thickness is

    measured at different locations of the cutting edge. Although the poor conductivity of the

    silica core hinders a sharp imaging contrast at the gold-silica interface, the measurements

    give direct evidence of the shell thickness, which is about 10 nm.

    The way how gold ions are introduced into the reaction is critical in determining the

    surface morphology. As the second example, it is found that if the planting solution is

    added by a small amount at one time and for multiple times, the attached THPC gold

    nanoparticles will only increase in size but cannot coalesce into a complete shell. Figure

    2.6(a) and (c) shows the SEM images of a 1 μm and a 2 μm particle, respectively.

    Compared with the protocol used to obtain core-shells in Figure 2.4, the volumes of the

    precursor particle suspension and of the planting solution are reversed. The reaction is

    conducted for 5 cycles, and particles are washed carefully after each cycle to remove

    excess formaldehyde. From the SEM images, it can be seen that in both cases, the growth

    of gold nanoparticles are obvious. However, despite the high surface coverage, the grains

    are isolated and do not form a continuous monolayer. When the same process is repeated

    for 12 cycles, interestingly, the grain size increases significantly, as shown in Figure

    2.6(c), but the shell is still not compete, exhibiting widespread discontinuities and large

    roughness. Up to 20 cycles of reduction have been tested. Nevertheless, the particles

  • 22

    aggregate easily during later cycles of reaction, frustrating further attempts on shell

    formation. In comparison, in Figure 2.6(d), when a large amount of planting solution is

    added to react with the particles shown in Figure 2.6(c), a complete shell is obtained,

    although the surface smoothness degrades slightly. Therefore, it can be concluded that the

    amount of gold ions introduced to a single reaction is critical to the shell formation.

    Figure 2.6: (a,c) SEM images of a 1 μm (a) and a 2 μm (c) particle after 5 cycles of gold

    reduction with a small amount of planting solution. The gold nanoparticles increase in

    size (~10 nm) but are isolated. (b) Zoom-in image of a 1 μm particle after 12 cycles of

    reduction. The grain size increases to 20 nm and larger, whereas the nanoparticles still do

    not coalesce into a complete shell. (d) Zoom-in images of a 2 μm particle after 5 cycles of

    reduction with a small amount of planting solution and an additional cycle with a large

    amount (5 times in volume) of planting solution. A complete shell is formed whereas the

    smoothness degrades slightly.

    The concentration of gold ions also plays an important role. In a separate trial, a

    condensed planting solution is used. To make a fair comparison to the result in Figure

    2.6(a), the concentration is increased by 5 times while the volume of planting solution is

  • 23

    unchanged (200 μL), and the reduction is conducted for only one cycle. The image of

    resulting particles is shown in Figure 2.7, where the morphology further differs from

    previous results. The increase of gold ion concentration causes abrupt growth of the seed

    nanoparticles but does not promote the coalescence. The grains are clearly isolated,

    forming satellite nanostructures on the silica surface.

    Figure 2.7: SEM image of a 1 μm particle after a single cycle of reduction with a

    condensed planting solution. The attached seed nanoparticles increase significantly in

    size but are clearly isolated.

    In this section, the synthesis of SiO2/Au core-shell particles is discussed in detail.

    Three different morphologies can be obtained by changing the ratio of the reactants and

    by varying the concentration of the planting solution. Silica particles coated by a

    complete shell (Figure 2.4), a defective shell (Figure 2.6(d)), a discontinuous shell

    (Figure 2.6(a-c)), and satellite nanostructures (Figure 2.7) have different optical

    properties and applications, which will be addressed in later chapters.

    2.3 Assembly of Clusters

    The self-assembly process is then conducted in a relatively straightforward manner

    [39, 46]. A droplet of diluted core-shell particle suspension is pipetted on a gold-coated

  • 24

    substrate and dried slowly at room temperature. During total evaporation, the particles are

    assembled by the capillary force randomly into small clusters of different planar

    geometries. Figure 2.8 illustrates the schematic. Although any insoluble substrate can be

    used, gold-coated high reflective mirrors are favored in the subsequent infrared optical

    characterization. Polished silicon substrate, TEM grids, infrared transparent windows

    made of Calcium Fluoride (CaF2) are also possible options.

    Figure 2.8: Self-assembly of core-shell particles. A droplet of diluted particle suspension

    is pipetted on a gold-coated substrate. During total evaporation, the core-shell particles

    are assembled by the capillary force into different planar structures.

    Figure 2.9: Self-assembled 1 μm core-shell clusters. Representative geometries include

    but are not limited to: (a) dimers, (b-c) trimers, and (d) chains. The shell thicknesses are

    estimated to be about 30 nm. Scale bars are 200 nm.

  • 25

    Figure 2.10: Self-assembled 2 μm core-shell clusters. Representative configurations

    include but are not limited to: (a) dimers, (b) trimers, (c) quadrumers, (d) chains, and (e)

    pentamers. The shell thicknesses are estimated to be about 10 nm. Scale bars are 500 nm.

    Figures 2.9 and 2.10 show the SEM images of representative clusters consisting of 1

    and 2 μm core-shell particles, respectively. In addition to single particles (monomers),

    other regular configurations include dimers, trimers, quadrumers, chains of different

    lengths, and by chance pentamers, heptamers or larger ensembles. The interparticle

    spacing within a cluster is tiny compared with the particle sizes, not larger than 30 nm

    estimated from the SEM images. On the other hand, the assembled clusters are relatively

    isolated from each other. The typical distance between neighboring clusters ranges from

    10 to some hundreds micrometers. As a result, the scattering from one cluster will not be

    affected much by nearby particles, which allows us to characterize the optical properties

    of individual clusters.

    During the assembly process there are two phenomena worthy of attention. First, the

    probability of forming simple and symmetric structures, such as dimers, trimers, and

  • 26

    quadrumers, are much higher than that of forming large and complex structures.

    Meanwhile, it is found that large clusters tend to appear near the edge of the droplet,

    accompanied by more excess nanoparticles, as shown in Figure 2.10(d) and (e). This is a

    consequence of the kinetics of evaporation. To make the clusters cleaner, either thorough

    washing after synthesis or moving the droplet during evaporation could be helpful.

    Figure 2.11: Two clusters consisting of particles with incomplete coating. Some of the

    attached nanoparticles drop off during the movement of the particles, forming a trial on

    the substrate. The scale bars are 500 nm in (a) and 1 μm in (b).

    The second phenomenon is about the bonding strength between the silica core and the

    gold shell. In Figure 2.6(a), (b) and (d), several defects sized over 100 nm can be seen.

    However, it is not very clear if they are caused by incomplete functionalization—this

    cannot be fixed in later synthesis steps—or by centrifugation. The reason is distinguished

  • 27

    by a revised assembly process, during which the droplet is gently blown to move a short

    distance on the substrate. Figure 2.11 shows the resulting clusters formed by 2 μm

    particles with incomplete coating, which are synthesized by using small amounts of

    plating solution as discussed above. From the images, unexpectedly, one can see dropped

    nanoparticles from the silica surface. Moreover, the distribution of these nanoparticles is

    not random but follows clear trials, indicating that the exfoliation occurs during the

    movement of the silica cores. In an extreme case in Figure 2.11(b), numerous

    nanoparticles drop off from a particle in a heptamer, likely during rolling, leaving a belt-

    like uncoated area on the silica surface. Similar defects are only observed for particles

    with incomplete coating or from multi-step reduction, whereas for complete shells, no

    visible damage is found after centrifugation, sonication, and assembly. Defective shells

    can diminish spectral features of the particles and clusters and thus have negative impacts

    on their optical properties. These results provide useful information to avoid defects in

    the synthesis and assembly processes.

  • 28

    Chapter 3

    Self-Assembled Plasmonic Core-Shell Particles

    as Infrared Resonators

    The successful synthesis of high quality core-shell particles and assembly of

    symmetric clusters provide a set of targets that is ideal for optical characterization.

    Following the fabrication processes above, in this chapter, the infrared optical properties

    of individual core-shell particles and self-assembled clusters are systematically studied by

    both experiments and simulations. It is shown that core-shell particle clusters exhibit

    unique optical properties in the infrared, especially MIR region. Depending on the

    number and configuration of the constituent particles, the reflection spectra from

    individual ensembles show distinct line shapes. In other words, the optical properties of

    the clusters are sensitive to structural and geometric changes, indicating a promising

    toolbox for infrared plasmonics, metamaterials and associated applications.

    3.1 Bright-Field Reflection Measurements

    Fourier transform infrared (FTIR) spectroscopy is employed to characterize the

    optical properties of individual clusters with different numbers of particles and varying

    configurations. The measurements are conducted using the Bruker spectrometer Vertex

    70 combined with a microscope Hyperion 1000. Figure 3.1 illustrates the schematic of

    the setup for reflection measurement. Infrared light is generated by a globar source (for

    MIR) or a tungsten halogen lamp (for NIR). After passing through a Michelson

  • 29

    interferometer, the light beam is focused on the sample surface by a 36x Cassegrain

    objective (NA = 0.5), which shapes the incident beam to be donut-like in cross section,

    corresponding to incident angles ranging from ~18 to 30°. In the present study, the

    reflected light is collected by the same objective and directed to a Mercury-Cadmium-

    Telluride (MCT) detector. In the case of transmission measurement, two identical

    Cassegrain objectives are needed. The incident light is launched by one lens below the

    sample stage, and the transmitted light is collected by the other on the top. In order to

    measure the signal from a single cluster, a knife-edge aperture in the conjugate image

    plane is carefully adjusted to define a rectangular area containing exclusively the cluster

    of interest. Depending on the signal-to-noise ratio (SNR) as well as the number and

    configuration of the particles in the clusters, the size, shape, and orientation of the

    aperture varies, roughly from 100 to 400 μm2. Focusing should be carefully tuned to

    maximize the signal intensity. For each measurement at MIR (NIR) wavelengths, the

    sample is scanned for at least 2500 times with a 4 cm-1

    (8 cm-1

    ) spectral resolution and a

    proper amplification. This scheme can effectively suppress noises and improve SNR. A

    reference measurement in a nearby sample-free area is carried out immediately after each

    sample measurement, while all the settings on focusing, knife-edge aperture size, number

    of scans and spectral resolution remain unchanged. Reflectance is then calculated by

    normalizing the sample spectrum to the reference spectrum. All the measurements are

    based on unpolarized illumination.

    Uncoated silica particles and particles with incomplete coating (see Figures 2.6(c) and

    2.11) are first measured. Because bare silica particles will aggregate into a monolayer

    easily on the substrate, extremely low concentration is used to obtain isolated monomers.

  • 30

    Figure 3.1: Schematic of the Fourier transform infrared (FTIR) single cluster

    spectroscopy. Infrared light is generated by a globar source. After passing through a

    Michelson interferometer, the incident beam is focused by a 36x Cassegrain objective

    lens on the sample. The reflected light is collected by the same lens and directed to an

    MCT detector. A knife-edge aperture is applied in the conjugate image plane to define a

    small area containing exclusively the individual clusters of interest, represented by the

    rectangle bounded by the red dashed lines.

    Surprisingly, it is found that the infrared optical properties of an uncoated silica particle

    and of a particle with incomplete coating are nearly identical. Although the gold

    nanoparticles cover about half of the silica surface, their sizes are around 10 nm, more

    than 150 times smaller than the infrared wavelengths (1.5 μm and longer). Therefore, the

    existence of the non-coalesced gold nanoparticles on silica beads will only change the

    optical properties in the visible region but will not affect the infrared optical response.

    This allows one to investigate the infrared optical properties of uncoated silica particle

    clusters by measuring their counterparts with incomplete coating, which are easy to

    assemble. In the following, these two types of particles will not be distinguished, and

    both of them will be referred to uncoated silica particles.

  • 31

    Figure 3.2(a) compares the reflectance spectra of four different 2 μm silica bead

    clusters on a gold substrate. With the number of constituent particles increasing from 1

    up to 7, the measured clusters include a monomer, a dimer, a symmetric trimer, and a

    symmetric heptamer. As one can see, in addition to the absorption from carbon dioxide

    (CO2) in air at 4.3 μm and from silica at 8-10 μm, a broad reflection dip occurs at 1.5-4

    μm, which is caused mainly by light scattering. However, instead of exhibiting any

    noticeable shifts, all the spectral features (peaks and dips) line up very well in the vertical

    direction and simply become more evident with the increasing number of particles. This

    is because the refractive index of silica is relatively low, less than 1.5 below 8 μm and not

    significantly greater than the surrounding medium (air). When the particle size is

    comparable to the wavelengths, no