Location via proxy:   [ UP ]  
[Report a bug]   [Manage cookies]                
Skip to content
Publicly Available Published by De Gruyter January 5, 2013

Towards nano-scale photonics with micro-scale photons: the opportunities and challenges of mid-infrared plasmonics

  • Stephanie Law , Viktor Podolskiy and Daniel Wasserman EMAIL logo
From the journal Nanophotonics

Abstract

Surface plasmon polaritons and their localized counterparts, surface plasmons, are widely used at visible and near-infrared (near-IR) frequencies to confine, enhance, and manipulate light on the subwavelength scale. At these frequencies, surface plasmons serve as enabling mechanisms for future on-chip communications architectures, high-performance sensors, and high-resolution imaging and lithography systems. Successful implementation of plasmonics-inspired solutions at longer wavelengths, in the mid-infrared (mid-IR) frequency range, would benefit a number of highly important technologies in health- and defense-related fields that include trace-gas detection, heat-signature sensing, mimicking, and cloaking, and source and detector development. However, the body of knowledge of visible/near-IR frequency plasmonics cannot be easily transferred to the mid-IR due to the fundamentally different material response of metals in these two frequency ranges. Therefore, mid-IR plasmonic architectures for subwavelength light manipulation require both new materials and new geometries. In this work we attempt to provide a comprehensive review of recent approaches to realize nano-scale plasmonic devices and structures operating at mid-IR wavelengths. We first discuss the motivation for the development of the field of mid-IR plasmonics and the fundamental differences between plasmonics in the mid-IR and at shorter wavelengths. We then discuss early plasmonics work in the mid-IR using traditional plasmonic metals, illuminating both the impressive results of this work, as well as the challenges arising from the very different behavior of metals in the mid-IR, when compared to shorter wavelengths. Finally, we discuss the potential of new classes of mid-IR plasmonic materials, capable of mimicking the behavior of traditional metals at shorter wavelengths, and allowing for true subwavelength, and ultimately, nano-scale confinement at long wavelengths.

Reviewed Publication:

Shalaev Vladimir


1 Surface plasmons

Optics and photonics have become increasingly interleaved with our everyday life and technology. Prime examples include optical communications’ role in the ongoing information technology revolution and the potential for optical sensing to address multiple health and security concerns our society faces every day. However, conventional photonic systems are limited, by diffraction, in their ability to focus and control light on a subwavelength scale. This diffraction limit prevents the confinement of light in volumes substantially smaller than the light’s wavelength in all three dimensions. The diffraction limit can be overcome, however, when photons are coupled to subwavelength-scale structures. The ability to control light on this scale is most effective when these subwavelength structures couple to light resonantly. Resonant interaction between photons and free-electron plasma in conductive nano- and micro-structures has emerged as a potentially revolutionary phenomenon to successfully solve a number of fundamental problems in sensing, photonics, and on-chip communications.

The interaction of light with conductive structures is dominated by the interaction of photons with the free-electron gas inside the conductive materials. To develop a quantitative description of such an interaction, we follow the well-developed Drude approach. The motion of the free-electron excited by a periodic optical field with angular frequency ω can be described by:

Assuming that the deflection of the electron from its equilibrium position can be written in the form x(t)=A exp (–iωt), the amplitude of the electron’s oscillation is given by

The polarization can now be calculated as P=enx=ε0χE with n being the density of the free-electron gas, ε0 being the permittivity of free space, and χ the polarizability. Finally, the relative permittivity of the plasmonic material, εm, is given by

with plasma frequency

γ the scattering rate for electrons in the metal, and the parameter εb the material’s background permittivity, taking into account the effect of bound electrons. Note that the plasma frequency is primarily determined by the concentration of free charge carriers inside the medium. In noble metals, the electron concentration is virtually fixed (n~1022cm-3). However, electron concentrations can be efficiently controlled in semiconductor structures, tuning the plasma frequency across the near-IR to the mid-IR and THz frequency ranges [1–7].

The permittivity of silver (Ag), a commonly used plasmonic material at near-IR and visible wavelengths, is shown in Figure 1, as a function of wavelength. Note that the permittivity is negative for wavelengths longer than the near-UV. In the visible and near-IR frequency ranges (Figure 1A), the permittivity of Ag is negative, and its absolute value is comparable to that of typical dielectrics. In contrast to this behavior, the |εm| of Ag at mid-IR frequencies is much larger than the permittivity of the surrounding media (Figure 1B). In this respect, Ag is similar to the vast majority of plasmonic metals (such as Au, Al, Ni, and Cu) investigated in the literature [8–10]. As will be discussed below, the negative value of a metal’s permittivity plays a vital role in defining the optical response of plasmonic structures. Thus, this drastic contrast between their short and long wavelength optical properties makes noble metal plasmonics at mid-IR frequencies significantly different from plasmonics at visible and near-IR frequencies.

Figure 1 Real (blue) and imaginary (red) values of permittivity for silver (Ag) in the (A) UV to near-IR wavelength range, and (B) imaginary and (C) real values of the permittivity of Ag from the near-IR well into the mid-IR. Open circles and squares: experimental data from ref. [10]. Solid lines: Drude fit to data of [10]. Note the large, negative ε of Ag at mid-IR wavelengths, an attribute shared with the majority of traditional plasmonic metals.
Figure 1

Real (blue) and imaginary (red) values of permittivity for silver (Ag) in the (A) UV to near-IR wavelength range, and (B) imaginary and (C) real values of the permittivity of Ag from the near-IR well into the mid-IR. Open circles and squares: experimental data from ref. [10]. Solid lines: Drude fit to data of [10]. Note the large, negative ε of Ag at mid-IR wavelengths, an attribute shared with the majority of traditional plasmonic metals.

For plasmonic materials, negative permittivity results from the collective oscillations of free carriers. However, negative permittivity can also result from lattice vibrations in polar dielectrics, giving rise to surface phonon polariton modes. The optics of these phononic materials often mimics the optics of plasmonic systems.

Of particular interest for mid-IR application are the phonon polaritons achievable on SiC, which has a negative permittivity resulting from optical phonons in the wavelength range between ~10.2–12.5 µm. Coupling to propagating as well as localized phononic modes has been demonstrated on SiC [11, 12], and SiC structures have been used to demonstrate directed thermal emission and superlensing [13, 14]. The advantages of the SiC system lie in the strong confinement of the phononic modes, their relatively weak damping, and the compatibility with Si [15]. Here, however, we focus primarily on plasmonic materials, as they offer broadband functionality across the mid-infrared, as opposed to the limited wavelength range associated with their phononic counterparts.

We illustrate the implications of the difference in the relative magnitudes of the metal and dielectric permittivity with three classical examples of coupling between photons and electronic plasma: localized surface plasmon (LSP) resonances in metallic nanostructures, propagating surface plasmon polariton (SPP) waves at metal-dielectric interfaces, and spoof SPP modes. We now discuss these resonances in more detail, and in particular, focus on the effect of the large negative permittivity of the plasmonic metals at the longer wavelengths of the mid-IR.

1.1 Localized surface plasmons

For simplicity, consider a spherical metallic nanoparticle (of diameter a) embedded in a dielectric medium of εd and excited by a homogeneous electromagnetic wave of amplitude Eo, as shown schematically in Figure 2A. Since the size of the particle is much smaller than the wavelength (a<<λ), the problem of electromagnetic field distribution inside and in the vicinity of the nanoparticle can be solved in the quasistatic approximation, where the time-varying electromagnetic field exciting the particle is considered to be constant in space, on a length scale larger than the nanoparticle. In this limit, it can be shown that the field inside the particle is homogeneous and is related to the incident field via [16].

Figure 2 Schematics of (A) localized surface plasmon (LSP) and (B) surface plasmon polariton (SPP) modes at metal dielectric interfaces, showing both the charge density oscillations in the metal and the electric fields in the surrounding dielectric.
Figure 2

Schematics of (A) localized surface plasmon (LSP) and (B) surface plasmon polariton (SPP) modes at metal dielectric interfaces, showing both the charge density oscillations in the metal and the electric fields in the surrounding dielectric.

As can be clearly seen from Eq. (4), the induced field inside the metallic sphere (and therefore the field in the immediate vicinity of this particle) is strongly enhanced when εm→-2εd. This field enhancement is primarily limited by absorption inside the metallic particle and also by the frequency detuning from the resonant condition. Enhancement of local fields can be used in a number of sensing applications as well as solar cell technology [17–19]. At the same time, the strong inhomogeneity of the electric field promotes a stronger enhancement of higher moments of the field and can be utilized for nonlinear optics applications [20, 21]. For the majority of noble metals, the resonant condition for SP excitation in spherical particles happens in the near-UV frequency range.

It is critical to note that, being a quasi-static excitation, the LSP resonance does not depend on the size of the particle, a. It can, however, be controlled by adjusting the shape of the particle and by varying the dielectric environment around the particle. These two mechanisms can be utilized to tune the SP resonance to some degree, with the former shown in Figure 3 [22]. In practice, in isolated nanostructures, such tuning is effective across the visible and near-IR frequency ranges [23–27]. A wider spectral tuning can be obtained in complex composite systems, such as fractals or percolation films [21].

Figure 3 Spectral location of the LSP resonance for subwavelength metallic nano-rod particles as a function of increasing aspect ratio; dots correspond to experimental data, line corresponds to quasi-static calculations. Deviation between quasi-static theory and experiments results from antenna-like resonances in longer nanoellipsoids. Copyright (2005), from ref. [22], with permission of the author.
Figure 3

Spectral location of the LSP resonance for subwavelength metallic nano-rod particles as a function of increasing aspect ratio; dots correspond to experimental data, line corresponds to quasi-static calculations. Deviation between quasi-static theory and experiments results from antenna-like resonances in longer nanoellipsoids. Copyright (2005), from ref. [22], with permission of the author.

At mid-IR frequencies, where |εm|→∞, the field inside plasmonic nanoparticles is vanishingly small, reflecting the fact that at these frequencies the properties of noble metals resemble those of perfect conductors. This effectively precludes subwavelength metallic particles from supporting such localized surface plasmon (LSP) excitations in the mid-IR. However, the visible-photonics of noble metals can be successfully mimicked in mid-IR “designer metals” fabricated, for example, by highly doped semiconductors, an effect discussed at greater length later in this article.

1.2 Surface plasmon polaritons

Consider now the interface between a metal with permittivity εm (occupying the half-space z<0) and a dielectric with permittivity εd (occupying the half-space z>0), shown in Figure 2B. Under some conditions, such an interface can support a guided mode propagating along the interface and exponentially decaying into both the metal and dielectric half-spaces. The properties of such a wave can be derived by considering the boundary conditions of Maxwell’s equations. Assuming that the mode propagates along the x direction, the field of such a wave can be derived based on the magnetic component [16]:

with

Continuity of the x component of the electric field

along with the dispersion relations inside both the metal and dielectric,
yields

Equation (6) represents dispersion of an interface-guided mode, most often referred to as a surface plasmon polariton (SPP), propagating along the metal-dielectric boundary. It can be seen that the solution to this dispersion equation corresponding to a propagating mode

is only possible when εm<0 and |εm|>εd.

Due to their exponential decay inside the dielectric, SPPs are often heralded as highly confined modes, potentially enabling sub-wavelength manipulation of light. Quantitatively, the light confinement can be characterized by the characteristic decay length of light into the dielectric,

where λo is the free-space wavelength and the ratio nSPP=kxc/ω represents the effective refractive index of the SPP mode. It is clearly seen in Figure 4A that the strong subwavelength confinement can be achieved only in the vicinity of the SPP resonance, where |εm|→εd so that nSPP>>1. For noble metals, these conditions are only satisfied at visible (or near-IR) frequencies.

Figure 4 (A) Calculated penetration of SPP mode into dielectric (normalized to free-space wavelength λo) for an SPP at a Ag/air interface. (B) Calculated propagation length (normalized to free-space wavelength λo) of an SPP at Ag/air interface. Note the large penetration depth, and propagation length, for SPPs at mid-IR wavelengths. Permittivity of Ag is taken from Figure 1.
Figure 4

(A) Calculated penetration of SPP mode into dielectric (normalized to free-space wavelength λo) for an SPP at a Ag/air interface. (B) Calculated propagation length (normalized to free-space wavelength λo) of an SPP at Ag/air interface. Note the large penetration depth, and propagation length, for SPPs at mid-IR wavelengths. Permittivity of Ag is taken from Figure 1.

At mid-IR wavelengths, conventional SPPs decay over macroscopic distances, moving away from the metal-dielectric interface, and effectively become surface-current guided modes propagating along the metal-dielectric interface with the speed of light in the dielectric. While such modes can propagate for very long distances (>>λo, Figure 4B), due to their weak penetration into the lossy metal, the large decay lengths into the dielectric (Figure 4A) lessens their utility for applications benefiting from strong confinement of the electric field.

1.3 Spoof plasmons

This drastic difference between SPPs in the visible and mid-IR frequency domains has fueled research into alternative means to confine propagating radiation to subwavelength length scales at mid-IR frequencies. In contrast to localized surface plasmon (LSP) resonances, where such efforts are effectively limited to choosing an alternative material platform (also a potential solution for enhanced mode confinement of SPPs), the behavior of SPPs can additionally be mimicked in the mid-IR frequency range using structured perfectly-conducting systems via spoof plasmons [28–30].

Originally introduced in [28], the spoof SPP is a collective excitation of the waveguide modes propagating in perforated films and free-space radiation. An example of such a structure is shown in Figure 5. When the surface of the structure is a perfect metal, each corrugation of the system can be considered to behave as a (subwavelength) waveguide with the core characterized by permittivity εh and permeability μh. Therefore, the field inside such a corrugation can be represented as a linear combination of the waveguide modes. In the limit when a <<λ0 the contribution from the fundamental mode becomes dominant.

Figure 5 Example of a THz spoof SPP structure. (A) Schematic of metal structure with subwavelength grooves of width a, depth h, and periodicity d, which can be approximated as an effective medium with εy, εz, and μx. (C) Real (red) and imaginary (blue) dispersion relation (k vs f) for a system with a/d=0.3, a=3 μm, and h=30 μm. Black dashed line shows dispersion for a perfect conductor. Copyright (2010), Springer, from ref. [31], with permission of the author.
Figure 5

Example of a THz spoof SPP structure. (A) Schematic of metal structure with subwavelength grooves of width a, depth h, and periodicity d, which can be approximated as an effective medium with εy, εz, and μx. (C) Real (red) and imaginary (blue) dispersion relation (k vs f) for a system with a/d=0.3, a=3 μm, and h=30 μm. Black dashed line shows dispersion for a perfect conductor. Copyright (2010), Springer, from ref. [31], with permission of the author.

In the same limit, the collective response of the multiple corrugations can be approximated by effective medium theory, which yields the plasma-like response with effective plasma frequency [28]

The corrections to the effective medium theory due to the finite conductivity of noble metals, important for the mid-IR frequency range have been developed in Ref. [31]. As expected, these corrections result in absorption losses, somewhat similar to the propagation losses in visible-frequency SPPs. As noted in [31], however, the quality factors of spoof plasmons are typically smaller than those of their visible-frequency counterparts. While the periodicity of the spoof plasmon structures (and the associated propagating modes) can be significantly subwavelength, accurately mimicking short wavelength SPPs, it should be noted that the grooves themselves approximate quarter wave antennae, and as such, require that the spoof plasmon geometry be limited in size by the wavelength of the incident light, at least in one dimension. Nonetheless, spoof SPP structures offer significant design flexibility, as they can be designed across a wide range of wavelengths, in particular those in which metals have large, negative permittivity, such that the spoof mode mimics the highly confined and lossy modes of classical SPPs at short wavelengths.

1.4 Coupling to and from SPPs

Modulation, or control, of the plasmon confinement scale is useful in applications ranging from on-chip photonics to sensing to directional control of light [32–36]. However, coupling these modes to free-space radiation is essential for integration of plasmonic devices and structures with macro-scale optical components and devices.

Regardless of the nature of the SPP or spoof SPP, its effective modal index is always above that of the adjoining dielectric. This index mismatch presents an effective mismatch in the momentum spectrum of light, and thus prevents direct coupling between the guided mode and propagating plane waves in the dielectric. The coupling can be introduced [32, 33], and controlled [30, 32, 37] by either introducing a high-index layer, weakly coupled to the SPP (such as Kretchmann or Otto configurations) or by corrugating the surface of the metal.

Naturally, SPPs can propagate not only on the flat metal surfaces, but also on curved structures, such as cylinders or spheroids, etc. With some exceptions (propagating of a guided mode along the axis of an infinitely long cylinder), introduction of curvature or finite-size corrugations into the system results in the coupling between the SPPs and free-space modes, thus leading to radiation losses in SPP waves, but at the same time introducing the mechanism to couple diffraction-limited waves to confined SPP modes. Such coupling is typically strongly enhanced when the size of the SPP-supporting structure is close to an integer number of half-wavelengths of the SPP [38].

1.5 Surface plasmons and antennae

One of the most often used families of SPP-supporting structures is known as optical antennae. However, the vast majority of antenna work, until recently, has focused on the very long wavelengths used for radio communications or radar applications (λ~mmm range). At such long wavelengths, metals can effectively be treated as perfect conductors, and plasmonic effects are negligible. Thus the plasmonic component of traditional antenna modes is rarely considered. In the visible frequencies, the permittivity of the metal is still negative, but quite small in magnitude (εm~εd) and penetration of an antenna mode into the metal can be significant, compared to the wavelength of the incident light and the dimensions of the antenna itself. At these frequencies the deviation of nSPP from 1 can be significant, and thus the dimensions of these antennae approach those typically associated with structures supporting localized surface plasmon resonances. In this sense, antenna systems effectively bridge the divide between the quasi-static SP resonators and wavelength-scale SPP platforms [38, 39].

As shown in Figure 3, the resonant LSP frequency of sub-wavelength structures can be redshifted by increasing the aspect ratio of these structures. However, commensurate with the increasing resonant wavelength is a rapidly increasing magnitude of the negative metal permittivity. A number of works have focused on plasmonic antenna resonances in the mid-IR. However, the resonant wavelength of these structures is dependent on the structures’ dimensions and scales with increasing antenna size. In this sense, such structures do not exhibit the size-independent resonance associated with the sub-wavelength metallic nanoparticles described by Eq. (4). Though the localized modes of such structures cannot be perfectly modeled using a perfectly conducting metal, the higher aspect ratio of antennae and higher metal permittivity shifts the physics behind noble-metal based SPP-antennae [39, 40] toward their well-known RF counterparts.

2 The mid-infrared

The vast majority of plasmonic research has focused on the visible and near-IR wavelength range, where strong confinement of LSP and SPP modes makes plasmonics an attractive approach for a variety of nanophotonic applications, including sensing and optical interconnects, as well as for integration into solar cell technologies for clean energy applications. However, there are opportunities for plasmonics to make an impact at longer wavelengths, particularly in the mid-IR. In this section we give a brief overview of the mid-IR wavelength range, its importance, and the current state-of-the-art for mid-IR optical technologies.

2.1 Why the mid-infrared?

The accepted spectral limits of the mid-IR wavelength range vary between disciplines and applications. In this work, we will consider the mid-IR to consist of the wavelength range between 3–30 µm. Within this range, several sub-ranges are notable for a variety of technical applications. These are the mid-wave infrared (MWIR, 3–5 µm), the long-wave infrared (LWIR, 8–12 µm) and the very-long-wave infrared (VLWIR, 12–30 µm) [41]. Taken together, the entirety of the mid-IR is of vital technical importance for a range of applications, chief among them optical sensing and thermal signature detection and control. Additionally, mid-IR frequencies have been suggested for free-space optical communications [42].

The interest in mid-IR optical sensing is due to the large number of molecules with fundamental vibrational absorption resonances in this wavelength range. These molecules are of significance for numerous medical, environmental, and industrial sensing applications. Figure 6 shows the infrared absorbance for five representative molecules with strong and distinct absorption spectra in the mid-IR [43]. Each spectrum constitutes a unique spectral fingerprint for the molecule. For this reason, the mid-IR is often referred to as the ‘molecular fingerprint’ region of the optical spectrum.

Figure 6 Mid-IR absorbance for five molecules of interest for medical, environmental and industrial sensing applications. Data from the EPA spectral database [43].
Figure 6

Mid-IR absorbance for five molecules of interest for medical, environmental and industrial sensing applications. Data from the EPA spectral database [43].

In addition, the mid-IR is also home to the thermal radiation (emitted by any finite temperature object) of hot bodies across a wide range of temperatures. Planck’s Law describes the thermal radiation, as a function of wavelength and temperature, from a perfect blackbody, an object that absorbs all incident radiation. Because real world objects are not perfect blackbodies, this thermal emission is typically modified by an emissivity ε, where 0<ε(λ, T)<1, which modulates the object’s thermal emission as a function of wavelength and temperature, resulting in an adjusted blackbody spectral emittance (in W-m-3), given as:

where T is the object’s temperature, λ is the wavelength of the emitted light, c is the speed of light in a vacuum, h is Planck’s constant, and k is the Boltzmann constant. Figure 7B shows the calculated emission spectra from a perfect blackbody for a range of temperatures. From this Figure, and quantitatively from Wien’s displacement law, it can be seen that the peak of the blackbody emission varies inversely with temperature:

Figure 7 (A) Atmospheric transmission spectra [44] and (B) blackbody spectral emittance for a range of temperatures. Dashed line traces wavelength of peak blackbody emission as given by Eq. (10). Shaded regions correspond to MWIR (blue) and LWIR (red) wavelength ranges. (C) MWIR thermal image of Blackhawk helicopter (Thermal image compliments of FLIR Systems, Inc) and (D) LWIR thermal image of the author’s (DW) research team (taken with FLIR BX320 thermal imaging camera).
Figure 7

(A) Atmospheric transmission spectra [44] and (B) blackbody spectral emittance for a range of temperatures. Dashed line traces wavelength of peak blackbody emission as given by Eq. (10). Shaded regions correspond to MWIR (blue) and LWIR (red) wavelength ranges. (C) MWIR thermal image of Blackhawk helicopter (Thermal image compliments of FLIR Systems, Inc) and (D) LWIR thermal image of the author’s (DW) research team (taken with FLIR BX320 thermal imaging camera).

where b=2.897×10-3 K-m, is Wien’s displacement constant. The peak emission wavelength (λp) is in the mid-IR for temperatures ranging from approximately 200–1400 K. Moreover, the technological importance of the MWIR and LWIR bands can be seen to originate from the overlap of these bands with the 3–5 µm and 8–13 µm atmospheric transmission windows (Figure 7A) [44]. These atmospheric transmission bands act as natural bandpass filters for thermal detection or imaging applications. The MWIR is then the band associated with thermal emission from ‘hot’ objects (e.g., turbines and engines), while the LWIR is more usually associated with thermal emission from ‘room temperature’ biological objects (e.g., humans), as is evident in Figure 7C, D. Thus, for target acquisition and tracking, and infrared countermeasures, as well as night vision imaging and thermal auditing, mid-IR materials and optoelectronic devices in the MWIR and LWIR are of vital importance.

In addition to the technological importance of the mid-IR for sensing and security applications, the 3–30 µm range is of interest for a variety of fundamental scientific investigations. The longer wavelengths of the mid-IR, when compared to the visible (400–700 nm) and near-IR (700 nm–1.4 µm), allow for the investigation of materials with subwavelength features fabricated using standard lithographic techniques, or ultra-subwavelength structures fabricated using more complex and expensive patterning technologies. The additional available range in feature size, normalized to operation wavelength, afforded by the longer wavelengths of the mid-IR makes the mid-IR a valuable test-bed for the investigation of optical phenomena across the UV-IR frequency ranges.

Furthermore, operation in the mid-IR allows for integration of plasmonic and metamaterial structures with traditional semiconductor materials, transparent at these longer wavelengths. This allows for the incorporation of well-understood electronic material systems with plasmonic architectures using technologically mature fabrication techniques. Such plasmonic/semiconductor integration would open the door to hybrid optoelectronic-plasmonic devices, giving a unique level of control over both the dielectric and metal constituents of plasmonics structures and devices.

2.2 Mid-IR optoelectronics

Interest in the mid-IR has been historically driven by the aforementioned applications. However, mid-IR systems development has been hampered by the lack of an optical infrastructure commensurate with the technological importance of the mid-IR. Mid-IR counterparts for basic optical and optoelectronic components in the near-IR/VIS are either an order of magnitude more expensive or do not exist. Significant effort over the past decades has been made to ameliorate this discrepancy, with research efforts to develop mid-IR detectors, sources, and optoelectronics.

While the standard mid-IR detector is the HgCdTe (MCT) photodiode, interest in more efficient, higher operating temperature detectors has fueled research into quantum well and quantum dot infrared photodetectors, as well as a new class of Sb-based detectors, such as type-II superlattice and nBn detectors [45–53].

Perhaps no single recent development has driven interest in the mid-IR more than the initial demonstration, and subsequent rapid development, of the quantum cascade laser (QCL) [54–57], a semiconductor laser source which generates light from transitions between quantized states in the conduction band of complex multi-period quantum well (QW) heterostructures (Figure 8). QCLs provide a compact, high power, wavelength flexible, and now, commercially available [58], coherent source for numerous mid-IR applications, giving systems developers and researchers alike a powerful tool for investigating novel phenomena, materials, and structures, as well as developing new gas sensing and countermeasure systems in the mid-IR.

Figure 8 (A) Schematic of electron transport and optical transitions in a QCL, with inset showing the front facet of a fabricated QCL. (B) Light output and applied bias as a function of current (I) and temperature (T) for a QCL emitting at λ~4.6 µm, shown in (A), inset. Inset shows 80 K (red) and 300 K (purple) emission spectra for the QCL shown in (A), inset. Figures reproduced with permission from Peter Liu, Princeton University.
Figure 8

(A) Schematic of electron transport and optical transitions in a QCL, with inset showing the front facet of a fabricated QCL. (B) Light output and applied bias as a function of current (I) and temperature (T) for a QCL emitting at λ~4.6 µm, shown in (A), inset. Inset shows 80 K (red) and 300 K (purple) emission spectra for the QCL shown in (A), inset. Figures reproduced with permission from Peter Liu, Princeton University.

The long-time characterization work-horse of the mid-IR is the Fourier transform infrared (FTIR) spectrometer, an interferometry-based spectroscopy tool invaluable for characterization of the optical properties of bulk materials, thin films, and quantum heterostructures in the mid-IR. Spatially resolved (though diffraction-limited) spectral data can be obtained by coupling a mid-IR microscope (using all-reflective optics) to an FTIR [59].

However, the emergence of mid-IR optical components based upon subwavelength-geometry metamaterial and plasmonic structures requires a new class of imaging and spectroscopic tools. Key among these are near-field scanning systems which can provide spatially-resolved spectroscopic data on length scales much smaller than the wavelength of the probe light [12, 60–64].

2.3 Why mid-IR plasmonics?

Though the rapid maturation of mid-IR source technology has been impressive, the development of a commensurate optical infrastructure, necessary to support the envisioned QCL-based sensing and countermeasure systems, has been slower to develop. Mid-IR plasmonics may well prove to be an enabling technology for the development of this infrastructure. Plasmonics has played a role in the development of the QCL, from early QCL waveguides to the more recent QCL beam steering and polarization control by facet-patterning. Further development of optical and optoelectronic surface plasmon-based technologies may also aid in the future growth of the mid-IR optical infrastructure required to support QCL-based systems.

A prime motivation for plasmonics devices and structures at shorter wavelengths lies in the SPs confinement to subwavelength mode volumes, enhancing light-matter interactions for emitters and sensors, and opening the door to light-guiding via plasmonic optical interconnects. This enhanced interaction has applications in the mid-IR as well, for chemical and biological sensing as well as detector and emitter enhancement.

For sensing applications, the strong and distinct mid-IR vibrational absorption resonances of numerous molecular species make the mid-IR an important wavelength range. The ability to confine mid-IR light to nano-scale volumes with plasmonic structures has the potential to enhance the interaction between incident light and these molecules. In a similar fashion, mid-IR plasmonics structures are of interest for enhancing coupling to mid-IR detectors, improving the efficiency and responsivity of these devices. Alternatively, plasmonic structures can be integrated with QCLs, acting to shape or steer the emitted light, or to couple the emitted light to on-chip waveguiding structures.

Finally, mid-IR plasmonic structures have significant potential applications in controlling and modulating thermal emission from heated objects. Here there exists no direct analogue to shorter wavelength plasmonic structures, as the mid-IR is home to the peak blackbody emission for most all terrestrial finite temperature objects. Plasmonic materials allow for designer-engineered optical properties, with controllable absorption resonances. By designing selectively absorbing plasmonic materials, the emissivity of a thermal emitter can be controlled, allowing for the development of selective thermal emitters for the cloaking or mimicking of heated objects.

2.4 Summary

Applications for mid-IR technologies span disparate fields, from security and defense, to communication, to environmental, industrial, medical, and chemical engineering, with additional potential for fundamental scientific investigations. The rapid development of the QCL has spurred increasing interest and subsequently, research investment in the mid-IR. From these efforts, new tools, phenomena, and materials have the opportunity to form an enabling optical infrastructure. As we shall show in the following sections, mid-IR plasmonic and hybrid plasmonic/optoelectronic structures and devices have the potential to provide a core component of this optical infrastructure, enabling the development of new technologies with benefits to a range of scientific disciplines, and more importantly, to society as a whole.

3 Noble metal mid-IR plasmonics

Initial efforts to investigate plasmonic structures in the mid-IR attempted to scale the geometry of short-wavelength plasmonic structures to longer length scales. These structures generally showed very similar far-field properties as their shorter wavelength counterparts, and as such, were adapted to sensing, detector and emitter coupling, and waveguiding structures in the mid-IR, applications not dissimilar to those proposed for short-wavelength plasmonic structures. Recently, metal plasmonics research in the mid-IR has begun to tackle devices and phenomena unique to the wavelength range, such as QCL laser beam shaping and thermal emission control. In this section we discuss the development and progress of mid-IR plasmonics with traditional metals.

3.1 Extraordinary optical transmission

The field of plasmonics, though active for arguably a century beforehand, was recently spurred by the demonstration of extraordinary optical transmission (EOT) by Ebbesen et al. in 1998 [65]. In EOT, strong transmission peaks were observed through metal films perforated with periodic arrays of subwavelength apertures (over 100% transmission, when normalized to the open area of the grating). Because the EOT was observed at frequencies corresponding to the expected coupling frequencies to SPPs on the patterned films, it was initially suggested that the EOT phenomenon was wholly a result of the excitation and subsequent re-radiation of SPPs. Though future work would show this phenomenon to be more complex [66, 67], it is widely agreed that SPPs play a major role in the EOT effect in periodic systems.

Early demonstrations of the mid-IR EOT phenomenon utilized traditional metals (Ag, Au, Ni, Cu), and resulted in strikingly similar transmission spectra to their shorter wavelength counterparts [68–72]. Figure 9 shows the EOT through a metal grating on a GaAs substrate. Here, the fundamental peak (1, 0), shows a peak transmittance of 33%, well above the grating open area (~20%), and not corrected for losses external to the plasmonic structure.

Figure 9 Normal incidence transmission spectrum for a Au mid-IR EOT grating with periodicity Λ=2.8 µm and aperture diameter D=1.4 µm, fabricated on a GaAs substrate. Peak transmission through the structure is ~33%, larger than the percentage of exposed GaAs (~20%) at the surface.
Figure 9

Normal incidence transmission spectrum for a Au mid-IR EOT grating with periodicity Λ=2.8 µm and aperture diameter D=1.4 µm, fabricated on a GaAs substrate. Peak transmission through the structure is ~33%, larger than the percentage of exposed GaAs (~20%) at the surface.

Though the far field transmission and reflection of the mid-IR EOT grating closely resembles that of a shorter wavelength structure, the near-field properties will differ significantly. As described earlier, the penetration depth of a mid-IR SPP can be well over a wavelength, and it is this weak confinement which results in the low propagation losses and long propagation lengths of the excitation. These attributes allow the SPP mode to be mapped without the need for subwavelength resolution [73, 74]. Figure 10 shows results from work where propagation of SPPs on mid-IR EOT gratings was characterized using a knife-edge experimental technique, a measurement technique possible with the longer propagation lengths of these long-wavelength SPPs.

Figure 10 (A) Experimental set-up for knife-edge measurement of propagating SPPs on a Ti/Au EOT grating fabricated on a GaAs substrate. (B) Contour plot showing intensity of transmitted/scattered light as a function of the position of the slit in (A) as a function of laser frequency (x-axis). Overlaid black line shows broadband transmission of the EOT grating. Adapted from [74].
Figure 10

(A) Experimental set-up for knife-edge measurement of propagating SPPs on a Ti/Au EOT grating fabricated on a GaAs substrate. (B) Contour plot showing intensity of transmitted/scattered light as a function of the position of the slit in (A) as a function of laser frequency (x-axis). Overlaid black line shows broadband transmission of the EOT grating. Adapted from [74].

3.2 Mid-IR sensing

Before the interest in EOT gratings and associated plasmonic geometries, there already existed a significant body of work on IR absorption enhancement from metallic surfaces. Infrared reflection absorption spectroscopy (IRAS) [75] is a long-used tool for measuring molecular absorption resonances, not reliant on plasmon-enhancement of the absorption signal (Figure 11A). Replacing the smooth metal films of IRAS with islanded metal films allows for coupling to localized excitations on the film surface, and thus normally incident absorption measurements, a technique commonly referred to as surface enhanced IR absorption (SEIRA) spectroscopy [76, 77].

Figure 11 Experimental schematics for (A) IRAS (B) reflection-mode SEIRA, (C) Kretschmann configuration SEIRA, with inset showing expanded view of metal, and (D) SPEIRA measurements.
Figure 11

Experimental schematics for (A) IRAS (B) reflection-mode SEIRA, (C) Kretschmann configuration SEIRA, with inset showing expanded view of metal, and (D) SPEIRA measurements.

The physical mechanism for SEIRA is similar to that of surface enhanced Raman spectroscopy (SERS), with SEIRA measurements of molecular absorption showing an absorption enhancement of 10–1000, when compared to standard, direct transmission, IR absorption measurements [78]. This additional enhancement results from the vibrational absorption of molecules in the strengthened near-fields associated with localized resonances on the islanded metallic surfaces. Figure 11B shows a schematic of a reflection SEIRA experiment. SEIRA relies on the inherent multiscale roughness of thin evaporated or electroplated metallic films in order to provide a surface with a broad spectral range of resonances. However, little effort was made to tailor the metal surface geometry to enhance absorption at specific wavelengths.

The localized resonances giving the absorption enhancement in SEIRA are often referred to, in the literature, as ‘plasmonic’, though they are perhaps closer to classical antenna resonances. Earlier work investigating SEIRA on thin metal films in the prism (Kretschmann) configuration (Figure 11C) noted that absorption enhancement in this configuration could be attributed to both localized field enhancement on the textured metal surface as well as coupling to propagating surface plasmons [79], an effect known to give enhancement of Raman signals [80]. Other groups have investigated prism-coupling purely to SPP modes for mid-IR biological sensing applications with promising results [81].

Subsequent efforts looked to use EOT gratings for detection of thin layers of molecular coatings, using so-called surface plasmon enhanced infrared absorption (SPEIRA) spectroscopy measurements [82, 83]. These works demonstrated significant enhancement in absorption (×100), when compared to previous IRAS measurements [84], attributed to the increased path length of light when coupled into propagating SPPs, which in turn, interact with the absorbing monolayers. Interestingly, the observed absorption enhancement is spectrally removed from the primary transmission (SPP) resonances of the gratings, though the authors argue that absorption is spectrally coincident with a number of broader, weaker, resonances. Further work investigated thin films of hexadecane on EOT gratings, and in particular the interaction of the CH2 rocking absorption line with the primary EOT transmission peak. It was shown that absorption was enhanced by approximately a factor of 3 when compared to the same absorption peak for hexadecane on a ZnSe substrate [84].

A direct comparison of SERS and SEIRA (or SPEIRA) is difficult. While both rely on plasmonic enhancement (though in very different wavelength ranges), the enhancement is of fundamentally different phenomena (the nonlinear Raman shift for SERS and direct absorption for SEIRA), with different selection rules. While SERS can provide an enhancement of 1012, compared to SEIRA’s 10–1000 enhancement, the signal being enhanced in SEIRA is much stronger than in SERS, making the resulting effectiveness of the two techniques comparable. Furthermore, the short-wavelength LSP leveraged in SERS provides strong coupling to molecules in the LSP near-field (monolayers) on metals such as Au, Ag, or Cu. The long-wavelength ‘plasmonic’ or antenna resonance used for S(P)EIRA is better able to probe thicker films, on a wider range of metals, including transition metals (which all have large negative permittivity in the mid-IR). Ultimately, many research teams use a variety of surface-enhanced sensing techniques, including both SERS and SEIRA [85].

The use of metal films capable of supporting SPPs or localized resonances for sensing in the mid-IR has clearly shown promise. However, unlike plasmon-enhanced sensing mechanisms at shorter wavelengths, which have been widely adopted as standard laboratory techniques (SERS) [86] or even commercial systems [87], mid-IR plasmonics-based sensing has not been broadly adopted outside the research lab, despite the value of the mid-IR as the molecular fingerprint region of the electromagnetic spectrum. This may be due, in part, to the large mode volume and weak confinement of the mid-IR SPP, which limits the interaction strength of the SPP mode with the molecules of interest. For this reason, there has been significant interest in developing plasmonic structures in the mid-IR which are capable of confining light to ultra-small mode volumes.

The enhanced absorption observed from SPEIRA measurements may, in some part be evidence of such an effect. The enhanced propagation lengths of the mid-IR SPPs on the EOT gratings (and thus enhanced interaction lengths with the adsorbed molecules) should be balanced by the weak overlap of the mode with the molecules of interest (due to the large dielectric penetration depth of the mid-IR SPP), limiting the achievable absorption enhancement for SPEIRA measurements. Taken together with the observed enhancement of absorption away from wavelengths corresponding to SPP-coupling on the metal mesh, these results may point to a more complex enhancement process than can be described by a simple path-length enhancement. Because the transmission process requires the coupling of light or SPP modes through the subwavelength aperture, the field profile around the grating structure is not a pure SPP mode, and can be quite strong in the aperture itself. Transmission through random apertures shows significantly weaker, but also broader, peaks than for the SPP-mediated transmission through periodic arrays of apertures [88]. Thus, it may be the field strength in the apertures that is responsible for the SPIERA effect, as well as the larger-than-expected tuning seen from voltage tunable plasmonic devices using EOT top gates in a metal-oxide-semiconductor structure [89].

Field enhancement in the mid-IR can also be achieved by use of micro-antenna-like structures. Here, again, the line between antenna and localized plasmonic resonances is somewhat blurred. Such structures do not exhibit resonances at exactly the wavelengths expected for pure antenna resonances of micro-scale structures fabricated from perfect electrical conductors (PECs) [39]. Instead, slight redshifts are observed in the structures’ resonances, resulting from the finite (negative) value of the metal permittivity. As such, these resonances are often referred to as ‘plasmonic’. However, the spectral positions of these resonances are strongly dependent on the antenna geometry (both shape and size), requiring the antenna dimensions to be on the order of the resonant wavelength. In this sense, such structures are dissimilar to the localized surface plasmon resonances (LSPR) observed in the visible and IR wavelength ranges for strongly subwavelength metallic particles, where spectral position of the resonances are shape dependent, but less directly dependent on size (assuming particle size <<λ) [90, and references therein].

Nonetheless, the benefits associated with the enhanced field strength of these structures should translate directly to similar systems with purely LSPR-like resonances. Micro-antenna arrays have been shown to strongly enhance (~105) the spectral signature of proteins (shown in Figure 12) and single layers of octadecanethiol [91, 92]. In addition, metal-dielectric-metal structures in which the top metal layer is patterned into a periodic array of antennae have demonstrated distinct spectral signatures indicative of strong coupling between molecular resonances in the dielectric material and the geometric antenna resonances of the patterned top layer [93]. The above-described structures have shown potential for sensing applications, though they are limited to a minimum size of approximately λ/2.

Figure 12 (A) Scanning electron microscope image of a periodic array of nanoantenna. (B) Comparison of reflectivity for bare antenna array and antenna array coated with a thin film of silk fibroin. Copyright (2001) PNAS, from ref. [91], with permission of the author.
Figure 12

(A) Scanning electron microscope image of a periodic array of nanoantenna. (B) Comparison of reflectivity for bare antenna array and antenna array coated with a thin film of silk fibroin. Copyright (2001) PNAS, from ref. [91], with permission of the author.

3.3 Detectors, sources, and waveguides

The use of plasmonic structures for improving semiconductor-based mid-IR detectors is fundamentally quite similar to SP-enhanced sensing in the mid-IR. In both cases the SP mode is used to enhance local field intensities and thus, interaction with either molecules of interest, or the device active region. Early proposals sought to use plasmonic surfaces in two slightly different ways: first, to couple to incident light and funnel this light to a subwavelength detector, for improved signal to noise ratio [94], and second, to couple light into SP/semiconductor modes, and enhance light/SPP interaction with the detector active region, using patterned top contacts are used [95].

The majority of experimental demonstrations of plasmon-enhanced mid-IR light detection have utilized the latter approach for integration with quantum dot-, quantum dots in a well-, and quantum well infrared photodetectors (QDIP, DWELL, and QWIPs, respectively) [96–99]. Figure 13 shows the results for a QDIP structure patterned with a top-contact EOT grating. In addition, recent work has demonstrated a plasmon-enhanced DWELL focal plane array camera with strongly enhanced absorption at λ=6 µm [100]. In each of the above detectors, incident light is coupled into EOT gratings above the detector structure, which simultaneously serve as top contacts for the detector device. The responsivity of the detectors is spectrally modulated by the momentum matching conditions for coupling to the grating modes, and at peak response can be significant, with enhancement factors as high as 30 reported [97].

Figure 13 Micrograph of metal photonic crystal (MPC) QDIP top contact (with expanded view) and (B) schematic of MPC-QDIP layer structure. (C) Spectral response of MPC-QDIP at 2.4 V (red) and -3 V (blue) applied bias compared to reference sample at the same biases (green, black) without the EOT top contact. (D) Normalized spectral response of MPC-QDIP structures at 4.4 and -4V and EOT periodicities of 2.5 and 3.6 µm. Copyright (2009) OSA, from ref. [91], with permission of the author.
Figure 13

Micrograph of metal photonic crystal (MPC) QDIP top contact (with expanded view) and (B) schematic of MPC-QDIP layer structure. (C) Spectral response of MPC-QDIP at 2.4 V (red) and -3 V (blue) applied bias compared to reference sample at the same biases (green, black) without the EOT top contact. (D) Normalized spectral response of MPC-QDIP structures at 4.4 and -4V and EOT periodicities of 2.5 and 3.6 µm. Copyright (2009) OSA, from ref. [91], with permission of the author.

Interestingly, one of the first uses of plasmonic materials in the mid-IR was not for detector enhancement, but as waveguiding structures in quantum cascade lasers. Traditionally, semiconductor lasers use semiconductor (dielectric) waveguides which confine the laser mode to the device active region by means of the difference in index of refraction between the core (active region) and cladding (dielectric waveguides). Such waveguides work well for the InGaAs/AlInAs QCL material system grown on InP substrates, but for GaAs/AlGaAs QCLs, the refractive index difference between laser core and cladding is small, and very thick waveguide layers are required to efficiently overlap the laser mode with the device active region (quantitatively given by Γ, the waveguide overlap factor). To overcome these problems, surface plasmon waveguides were implemented. In this case, the SPP mode is confined to the interface between the semiconductor and a gold top contact, with the first demonstration of such a laser achieving a Γ of 70% [101]. Later work demonstrated single-mode [102] and far IR [103] emission from QCLs with SP-waveguides, with Γ’s>80%. At the long wavelengths used in these lasers, the penetration depth of the SPP into the metal is relatively small, leading to low losses and high transverse-optical-mode confinement despite the high optical absorption of the metal.

Plasmonic waveguides, while widely used for THz QCLs (where the long emission wavelengths make dielectric waveguides impractical), are used in the mid-IR only for long-wavelength devices. In fact, recent work has demonstrated that unintentional coupling to SP modes the sidewalls of wet-etched, long-wavelength (λ~14 µm) QCLs is a source of waveguide loss [104].

There has recently been significant interest in intentional coupling of QCL emission to plasmonic structures, either for beam shaping, polarization control, or on-chip waveguiding. In the latter case, researchers have demonstrated the ability to couple QCL emission to SPP modes, either by use of a grating element to convert a QCL dielectric waveguide mode to a propagating SP [105] (shown in Figure 14), or by direct coupling of QCL facet emission to a planar SPP waveguide [106]. In addition to demonstrating a method for electrical injection of SPP modes, this work demonstrates the mid-IR’s utility as a test-bed for technologies at shorter wavelengths. Here, the mid-IR devices were used as a launching point for the proposal of a telecom SPP generator with potential applications for on-chip interconnect technology [107].

Figure 14 Schematics of QCL waveguides (A) without and (B) with grating coupler for efficient coupling of generated SPP modes to a flat passive SPP waveguide. AFM images of the fabricated devices (C) without and (D) with grating coupler and NSOM images of the same devices (E) without and (F) with grating coupler. In the case where a grating coupler is patterned into the metal layer, a clear SPP standing wave pattern can be observed, indicative of the interference between SPPs launched in the “generation” regions of the device and coupled via the grating into the “passive section”. Copyright (2010) Aps, from ref. [105], with permission of the author.
Figure 14

Schematics of QCL waveguides (A) without and (B) with grating coupler for efficient coupling of generated SPP modes to a flat passive SPP waveguide. AFM images of the fabricated devices (C) without and (D) with grating coupler and NSOM images of the same devices (E) without and (F) with grating coupler. In the case where a grating coupler is patterned into the metal layer, a clear SPP standing wave pattern can be observed, indicative of the interference between SPPs launched in the “generation” regions of the device and coupled via the grating into the “passive section”. Copyright (2010) Aps, from ref. [105], with permission of the author.

3.4 Beam shaping and steering

In addition to use as waveguides in, or coupled to, emitter devices, plasmonic structures in the mid-IR offer the potential for beam shaping forming applications. Early work at visible/IR wavelengths [35] demonstrated narrow-divergence beam steering from a subwavelength slit in a metal film, flanked by linear arrays of periodic corrugations. For such structures, light incident on the subwavelength slit can be directly transmitted, or alternatively, can couple to propagating SPs at the corrugated back-side of the metal film. These surface waves are then scattered from the back surface by the corrugations, and interfere in the far-field of the device structure to form the resulting beams.

In the mid-IR, such structures hold particular promise. Traditional edge-emitting QCLs emit highly divergent beams, diffraction-limited to angles θ ~ arcsin(λ/T) (with T being the thickness of the QCL active region, usually on the order of a few µm). As such, collection and collimation of the emitted light requires additional, external optics, as well as careful alignment of these optics. However, when plasmonic beam steering/shaping structures are patterned on the front facet of a QCL, the resulting beam can be narrowed (in the QCL growth direction) to values as small as ~2.4º [108, 109]. The work showing these effects sprung from earlier demonstrations of antennae patterned on QCL facets for strong near-field confinement of light at the QCL facet [110]. Similar structures fabricated on freestanding substrates showed the ability to steer incident QCL light by control of the incident light’s wavelength, or alternatively, the permittivity a dielectric material at the corrugated metal/dielectric interface [111]. Further work demonstrated that similar plasmonic facet patterning (either with a single, angled corrugation array, or with two orthogonal facet gratings) allows for control of the polarization state of light emitted from the QCL, giving angled linear, elliptical or circularly polarized emission, dependent on grating geometry, as shown in Figure 15 [112].

Figure 15 (A) Scanning electron micrographs of the facet of a QCL patterned with a plasmonic polarizer designed to give polarized emission 45° with respect to the vertical. (B) Far-field emission from the device shown in (A). (C) Measured (red circles) and calculated (black line) emission intensity as a function of polarization. Copyright (2009), AIP, from ref. [112], with permission of the author.
Figure 15

(A) Scanning electron micrographs of the facet of a QCL patterned with a plasmonic polarizer designed to give polarized emission 45° with respect to the vertical. (B) Far-field emission from the device shown in (A). (C) Measured (red circles) and calculated (black line) emission intensity as a function of polarization. Copyright (2009), AIP, from ref. [112], with permission of the author.

For these beam steering and shaping structures, the long propagation length of SPPs on metal films at longer wavelengths allow for interaction of the weakly-bound SPP mode with a large number of corrugations, giving rise to the opportunity to effectively control the evolution of the beams formed above the plasmonic grating. Careful design of non-periodic corrugations on the backside of the beam steering structure can effectively focus the transmitted light above the subwavelength slit, as shown in Figure 16 [37]. In such a structure, the surface is designed so that the light scattered from each corrugation interferes constructively at a common point above the surface, whose position can be controlled by the designer. In Figure 16, one can see both simulations and experimental results for such devices, designed to focus the transmitted light 250 and 500 µm above the central slit, respectively. In such structures, the long propagation length of the mid-IR SPP is used to the designer’s advantage and allows for the control of beam evolution in the vicinity of the plasmonic geometry.

Figure 16 (A, D) Numerical simulation of transmitted light intensity on back side of an engineering grating structure flanking a subwavelength slit (B, E) Simulation data from (A, D) smoothed to approximate experimental setup’s limited resolution. (C, F) Experimental measurement of field intensity on backside of the grating structure, for devices designed to focus 250 µm (A, B, C) and 500 µm above the sample surface. Reproduced from [37].
Figure 16

(A, D) Numerical simulation of transmitted light intensity on back side of an engineering grating structure flanking a subwavelength slit (B, E) Simulation data from (A, D) smoothed to approximate experimental setup’s limited resolution. (C, F) Experimental measurement of field intensity on backside of the grating structure, for devices designed to focus 250 µm (A, B, C) and 500 µm above the sample surface. Reproduced from [37].

Recent work has also demonstrated mid-IR beam steering, of a sort, using localized resonances of V-shaped micro-antenna arrays on planar dielectric surfaces [113]. Here, the incident field interacts with the micro-antennas on the planar surface, introducing a phase shift to the transmitted light. If the antennas are designed to give equal scattering intensity and spatially-varying phase shifts, the angle of refraction for the antenna-coupled light can be controlled by the magnitude of the phase gradient along the surface. It has also been suggested that the phenomenon of beam steering by resonant gratings can be described in terms of diffraction theory [114]. The Au subwavelength antenna structures used here and in subsequent work [115] are essentially folded quarter-wave antenna, demonstrating the utility of noble-metal ‘plasmonic’ antenna structures with localized resonances for beam steering, shaping, and polarization control applications [116].

3.5 Thermal emission

Plasmonic devices such as the beam steering and shaping structures described above, the QCL and on-chip QCL-coupled waveguides, and the EOT gratings and islanded plasmonic surfaces used for sensing, must, at all wavelengths, contend with losses. These losses appear as a decaying field intensity with propagation of the SPP mode, or a spectral broadening of a localized resonance resulting from scattering or absorption, and in most cases adversely affect the desired performance of the plasmonic structure. However, the mid-IR holds a spectral position with the unique ability to link an object’s optical and thermal properties, providing the opportunity to use plasmonic structures for thermal emission/signature control applications. The thermal emission resulting from the finite temperature of any object takes the spectral form given by Planck’s law, modulated by the object’s surface emissivity (ε). Most materials have largely constant emissivity as a function of wavelength, which allows thermal imaging systems to identify the object’s temperature despite the limited spectral range of the system sensor. However, by integrating engineered absorbers on an object’s surface, one can alter the emissivity of the surface as a function of wavelength, so that when heated, the object selectively emits thermal radiation in preferred wavelength bands. These coatings can then be used either as narrow-band thermal light sources, or alternatively, can be used to change the thermal signature of an object when viewed with a thermal imaging system. As such, the losses which plague most plasmonic structures can, in the mid-IR, be harnessed to demonstrate a new class of selective thermal emitters.

As an example, a simple 1D grating structure patterned into commercially available steel (shown in Figure 17A) allows for the out-coupling of thermally excited SPPs to free space photons. Because the grating is 1D, SPPs can only out-couple in one polarization (TM-polarized), while emission in the orthogonal (TE) polarization remains blackbody-like, with ε<<1, and constant. When heated, such a structure shows a clear spectral selectivity in its thermal emission, when comparing TM to TE polarized emission, as shown in Figure 17B [117]. At resonance (λ~10 µm), SPP-enhanced, TM-polarized thermal emission is a factor of ~2.6 greater than TE-polarized emission, an effect which can be clearly observed in polarized thermal images of the sample surface (Figure 17C, D).

Figure 17 (A) Schematic and scanning electron micrograph of grating-patterned steel substrate. (B) Thermal emission of patterned steel for a variety of etch depths, compared to patterned glass. Here, the spectra shown represent the TM polarized emission (with background subtracted) normalized to the TE emission (background subtracted). (C) TE and (D) TM polarized thermal images of the selective emitter surface, taken with a FLIR BX320 thermal camera. Adapted from [117].
Figure 17

(A) Schematic and scanning electron micrograph of grating-patterned steel substrate. (B) Thermal emission of patterned steel for a variety of etch depths, compared to patterned glass. Here, the spectra shown represent the TM polarized emission (with background subtracted) normalized to the TE emission (background subtracted). (C) TE and (D) TM polarized thermal images of the selective emitter surface, taken with a FLIR BX320 thermal camera. Adapted from [117].

However, it is preferred that the absorption/emission resonance be angle-independent, which makes the use of SPP-based devices difficult, as momentum-matching is required to couple these modes to free space photons (whose momentum depends on emission angle). Thus, most examples of selective thermal emitters utilize some form of localized plasmonic, metamaterial, or photonic crystal resonance. One such example is arrays of nano-scale slots in a metal film, with depths on the order of λ/4, allowing the slots to support “organ-pipe” modes with distinct absorption and, when heated, thermal emission resonances [118]. Later work demonstrated that the spectral linewidth of the thermal emission from these structures can be significantly narrowed by tailoring the nano-slot array periodicity to efficiently outcouple the organ-pipe plasmon resonances to free-space. These engineered materials were then used as narrowband thermal light sources for sensing applications [119].

The ability to modify the emissivity of an object allows one, effectively, to adjust its apparent temperature when viewed by a thermal sensor, as evidenced in Figure 17D. While strongly selective in emission, the organ-pipe structures described above are polarization sensitive, and require numerous exacting and time-consuming fabrication steps, which are difficult to incorporate into low-cost, large area coating applications. Planar, patterned metallic structures can be used to achieve strong absorption at designed resonances. Plasmonic perfect absorbers (PAs) consisting of subwavelength metallic discs or lines separated from a ground plane by a thin dielectric layer, have been shown to give perfect absorption at telecom frequencies [120, 121]. At longer wavelengths, PAs and selective thermal emitters have been demonstrated with similar structures using metamaterial top layers [122, 123]. Alternatively, antenna-like structures can replace the lithographically challenging metamaterial resonators of the top metal layer, and perfect absorption, as well as strongly selective and angle-insensitive thermal emission can be achieved [124]. Figure 18 shows a schematic of such a structure, in addition to the experimental reflection and thermal emission spectra from the fabricated sample. Such meta-surfaces allow for strong enhancement of thermal emission (ε>0.9) at resonance, with weak emission at all other wavelengths (ε<0.2).

Figure 18 Experimental reflection and thermal emission from a mid-IR selective emitter structure, demonstrating strong spectral selectivity, with (ε>0.9) at resonance, and weak thermal emission at all other wavelengths (ε<0.2). Inset shows schematic of the device characterized.
Figure 18

Experimental reflection and thermal emission from a mid-IR selective emitter structure, demonstrating strong spectral selectivity, with (ε>0.9) at resonance, and weak thermal emission at all other wavelengths (ε<0.2). Inset shows schematic of the device characterized.

3.6 Summary

Despite the significant progress in mid-IR selective thermal emitters, angle-insensitive and polarization-independent plasmonic perfect absorbers have yet to be demonstrated in the mid-IR. This is simply a result of the antenna-like nature of localized resonances in the mid-IR when using traditional metals. Thus the length scales required for localized resonators at these wavelengths will always be on the order of the wavelength of light. Likewise, for the many SPP-based geometries reviewed in this section, specifically the beam steering, waveguiding, and sensing structures, it can be argued that none are capable of true subwavelength mode confinement (at least in all three dimensions), due to the nature of traditional metals in the mid-IR. Reaching the nano-scale with mid-IR plasmonics therefore requires engineering the optical properties of plasmonic metals. Happily, this may not be as difficult as it might seem, as we will demonstrate in the subsequent section of this review, where we discuss how semiconductors can be engineered to behave in the mid-IR similarly to traditional metals in the visible and near-IR wavelength ranges, opening the door to bringing micron-scale wavelengths of mid-IR light down to the nano-scale.

4 New metals for mid-IR plasmonics

In the mid-IR, the vast majority of traditional metals have large, negative real parts of permittivity [8], which results in optical behavior fundamentally different from the same metals at shorter wavelengths. For this reason, SPP modes in the mid-IR have weak penetration into the metal (low losses), but at the same time, large penetration depths into the dielectric (weak confinement). The end result is a mode that is ‘light-like’, with limited utility for applications requiring strong field confinement (sensing, on-chip communication). At the same time, because of this large, negative permittivity, traditional metals cannot support the strongly subwavelength LSP modes in the mid-IR that have generated such interest at shorter wavelengths. Localized modes on wavelength-scale metallic structures can be achieved, but these are closer to antenna-like than LSPR-like, and scale nearly linearly with antenna dimension.

Thus, while the geometries of plasmonic structures can be easily scaled from short to long wavelengths, and dielectric materials found whose long wavelength optical properties closely approximate those of dielectrics at shorter wavelengths, the optical properties of most short wavelength plasmonic metals are drastically different at longer wavelengths. In order for mid-IR plasmonic structures to replicate, and improve upon, the optical behavior of their visible/near-IR counterparts, new plasmonic materials must be engineered to mimic the behavior of traditional plasmonic metals at short wavelengths.

In this section we discuss the recent development of two new classes of mid-IR plasmonic materials: graphene and engineered, or designer, metals. The former allows not only for strong confinement of optical modes on an atomic monolayer-thick two-dimensional (2D) surface, but offers the intriguing possibility of voltage-tunable plasmonic metals. The latter gives researchers a level of design control not achievable in shorter wavelength plasmonic structures, with the potential to enable a new generation of low-loss mid-IR plasmonic structures.

4.1 Mid-IR graphene plasmonics

Graphene, a 2D monolayer honeycomb lattice of carbon atoms, was first experimentally demonstrated only recently [125–127], despite being an active field of theoretical investigation for decades. While much of the interest in graphene has focused on its unique electronic, mechanical, and thermal properties, there also has been growing interest in the optical properties of graphene. In particular, graphene has drawn attention as an alternative plasmonic material at long wavelengths due to its unusual band structure, high mobility, tunable carrier concentration, and the resulting predicted strong confinement and long propagation lengths of SPPs [128, 129]. A recent review of graphene plasmonics provides a comprehensive overview of the field [130], so here we limit our discussion to selected long-wavelength applications of graphene as an alternative plasmonic metal.

A combination of low dimensionality and unique band structure results in a graphene plasma frequency which scales as n1/4, as opposed to the n1/2 for typical 2D or bulk (3D) structures (Eq. 3). Extremely large confinement, characterized by the ratio of the free-space wavelength λ0 to plasmon wavelength λp can be expected on graphene. The 2D carrier concentration of graphene can be varied over a wide range, using either electrostatic doping via a gate electrode, or by doping using molecular charge transfer [131], with carrier concentrations as high as 5×1013cm-2 reported [132]. This ability to design, or electrostatically control, the 2D carrier concentration, and thus the plasma frequency, makes graphene an exciting new material for IR plasmonic applications, as well as for fundamental investigations of 2D plasmons. IR spectroscopy has proved to be a valuable tool for probing the unique bandstructure of graphene, as well as demonstrating voltage tuning of graphene’s Fermi energy [133, 134].

Geometry plays equally important role in tuning graphene’s optical response. SPPs propagating in graphene nanoribbons have been reported at terahertz frequencies [135]. By either changing the ribbon width or applying a gate voltage to the ribbons, the frequency of the SPP excitation can be tuned. Surface plasmons in graphene sheets deposited on an SiO2/Si substrate have been imaged with scattering-type SNOM at λ0=11.2 μm [136], revealing extremely high confinement factors (λo/λp>50), tunable by application of a gate voltage. Similar SNOM studies on tapered graphene nanostructures have imaged propagating plasmonic modes and demonstrated strong confinement (λop>40) [137]. Moreover, propagation of the plasmonic mode can be effectively switched on and off by application of a gate voltage, opening intriguing possibilities for hybrid plasmonic/nano-electronic devices.

Coupling graphene and dielectric systems provides additional design elements in engineering the optical response. Localized plasmonic modes in nano-pillars with alternating graphene/insulator layers have shown plasmonic resonances in the far-infrared [138]. The spectral position of the resonance can be tuned by changing the number of layers in the stack. Though the field of graphene plasmonics is in its infancy, these initial results indicate that graphene could be a promising new material for infrared plasmonics.

4.2 Highly doped designer plasmonic metals

One of the primary benefits of graphene plasmonics is the design flexibility that comes with a material whose plasma frequency can be controlled, either by design (doping) or dynamically (electrostatic gating). Recently, significant design flexibility has also been achieved in a technologically mature class of materials: doped semiconductors.

Doped semiconductors offer an intriguing option for the development of ‘designer’ mid-IR plasmonic materials. The optical response of a doped semiconductor can be modeled as a Drude material, in the same manner as the classical metals described by Eq. (3),

where γ is the scattering rate for electrons in the semiconductor and εS is the undoped semiconductor’s relative permittivity. Here, the plasma frequency ωp differs from that of Eq. (3) in two key aspects. First, the carrier density n (and charge polarity) can be controlled either during growth or during fabrication, by choice of dopant (n-type or p-type) and doping density. Second, the electron (hole) mass used in Eq. (3) must be replaced by a frequency dependent effective mass m*(ω) which reflects the effects of the semiconductor’s bandstructure. Here, in contrast to the noble metal expression for ωp, we normalize the ωp of Eq. 11 with respect to the background permittivity, such that Re{ε(ω)}=0 for ω=ωp. A combination of high doping and small effective masses can lead to mid-IR plasma frequencies for III/V and group IV semiconductors.

The concept of designer plasmonic metals is not unique to semiconductors. Recent work has demonstrated transparent conducting oxides (TCOs), such as indium tin oxide (ITO), and nitrides with plasma frequencies in the near-IR to visible region, including the technologically important telecom wavelength range [1]. TCOs can be grown using a variety of techniques, including sputtering and pulsed laser deposition (PLD). In PLD, the dopant concentration is controlled by ablating two separate targets, for example In2O3 and SnO2 for ITO, with the appropriate number of pulses on each target. TCOs show significantly smaller losses in these wavelength ranges than the traditional plasmonic materials of gold and silver, as evidenced by their smaller imaginary dielectric constant shown in Figure 19B.

Figure 19 Real (A) and imaginary (B) part of the dielectric constant for nitride and TCO films as compared to gold and silver. Copyright (2011), OSA, from ref. [1], with permission of the author.
Figure 19

Real (A) and imaginary (B) part of the dielectric constant for nitride and TCO films as compared to gold and silver. Copyright (2011), OSA, from ref. [1], with permission of the author.

Typical doped TCOs demonstrate a plasma wavelength in the 1.5 μm range, which would result in extremely good confinement of plasmonic modes in the near IR, as opposed to the rather poor confinement seen in gold and silver. The plasma frequency of the TCOs can be tuned across a range of microns (a significant tuning range, given the short wavelengths of operation) by changing the doping of the material during growth [1].

Similarly impressive results have been demonstrated with nitride-based materials. Transition metal nitrides can be deposited by sputtering a pure metal target with a nitrogen-argon gas mixture. The film properties can be controlled by tuning the nitrogen:argon ratio. Though nitrides have higher losses than gold and silver, they are also easier to process using standard semiconductor techniques and can be grown epitaxially on substrates such as sapphire. This leads to lower losses from surface roughness in nanostructures than are seen in gold or silver, making TiN a good alternative to gold for structures such as metal-insulator-metal waveguides and sensors utilizing localized surface plasmon resonances [2]. In addition, extremely thin films of TiN grow continuously, unlike the percolating islands seen in gold films with thicknesses >10 nm. Smooth, non-islanded, films of these thicknesses are necessary for transformation optics applications, such as hyperbolic metamaterials, recently demonstrated with TCOs [139].

Traditional semiconductors provide a powerful potential material system for mid-IR designer metals with plasma wavelengths in the 3–30 µm range. Semiconductors have a number of clear advantages when compared to both traditional metals and the newer class of conducting oxide and nitride materials systems. First, epitaxial growth of semiconductors allows for single-crystal quality materials, atomic-scale precision in layer thicknesses, and accurate control of doping concentrations. Other semiconductor doping mechanisms, such ion-implantation or thermal diffusion of dopants do not offer similar precision in the vertical direction, but do offer the opportunity for lateral patterning of dopants. Moreover, traditional semiconductors are well understood material systems with an extensive existing processing and fabrication infrastructure. As a rule, these materials exhibit higher mobility (and thus smaller scattering rate) than their noble metal or transparent oxide counterparts [140]. Finally, the use of doped semiconductor designer metals allows for the potential integration of plasmonic structures with traditional semiconductor electronic and optoelectronic devices.

Doped semiconductors such as silicon and germanium, were among the first to be proposed as candidates for mid-IR plasmonic devices [141, 142]. Silicon has the obvious materials advantage of being perhaps the most well-understood of the two, with an extensive fabrication infrastructure available. Mid-IR plasmonic behavior has been realized both in ion-implanted silicon [143] as well as commercially-doped silicon wafers [144]. As shown in Figure 20, which gives ε of the doped Si for each of these two works, the plasma wavelength of the doped films can be tuned from approximately 4 μm to 11 μm using both n- and p-type material.

Figure 20 (A) Real (solid) and imaginary (dashed) parts of the permittivity for n-type (blue) and p-type (red) silicon. (B) Real and imaginary parts of the permittivity for silicon wafers doped n-type at 1×1020 cm-3 (red) and 6×1019 cm-3 (black). The use of doped Si allows for control of the material ωp across a wide wavelength range. Copyright (2011), AIP, from refs. [143, 144], with permission of the authors.
Figure 20

(A) Real (solid) and imaginary (dashed) parts of the permittivity for n-type (blue) and p-type (red) silicon. (B) Real and imaginary parts of the permittivity for silicon wafers doped n-type at 1×1020 cm-3 (red) and 6×1019 cm-3 (black). The use of doped Si allows for control of the material ωp across a wide wavelength range. Copyright (2011), AIP, from refs. [143, 144], with permission of the authors.

Designer metals in the mid-IR have also been shown in III/V semiconductor materials. III/V materials have distinct advantages when compared to Si. First, they possess both the potential for extremely high doping concentrations, as well as small electron effective masses, leading to remarkable wavelength flexibility across the entire mid-IR wavelength range. Furthermore, the III-Vs have the advantage of bottom up growth by well-established epitaxial techniques such as molecular beam epitaxy (MBE) or metalorganic chemical vapor deposition (MOCVD). Use of MBE or MOCVD for growth of doped semiconductors allows for single-crystal growth, precision control of layer thicknesses, and potential integration of semiconductor metals with the complex heterostructures and nanostructures used for cutting edge optoelectronic devices such as quantum cascade lasers or quantum well, quantum dot, or type-II superlattice infrared photodetectors.

Doped InAs in particular, is an excellent candidate for a mid-IR plasmonic material [145]. It has an extremely small effective mass and can be doped up to 1×1020 cm-3[146, 147]. Experimentally, the plasma wavelength can be tuned from 5.5 μm to 17 μm (or longer), as shown in Figure 21A for InAs films grown with various doping and growth parameters. Interestingly, the narrow band-gap of InAs, which would normally give strong absorption at shorter wavelengths in the mid-IR for these materials, shifts to frequencies in the near-IR due to conduction band state filling (Burstein-Moss effect), making these designer metals transparent out to telecom wavelengths [148, 149].

Figure 21 (A) Real and (B) imaginary parts of the permittivity of heavily doped InAs films with various carrier concentrations, layer structures, and growth parameters.
Figure 21

(A) Real and (B) imaginary parts of the permittivity of heavily doped InAs films with various carrier concentrations, layer structures, and growth parameters.

These films can also show low losses, as indicated in Figure 21B. This figure plots the imaginary part of the dielectric constant as a function of frequency normalized to the plasma frequency, and allows materials with very different plasma frequencies to be compared. The best highly-doped InAs films have losses at their plasma frequency comparable to the TCOs. A comparison to gold and silver, at least near the plasma frequency, is difficult due to the large interband absorption seen in these metals near the plasma frequency, which distorts the imaginary components of the metal permittivity away from a simple Drude model.

The development of mid-IR metals capable of mimicking the optical properties of classical metals at shorter wavelengths opens the door to a new class of mid-IR plasmonic structures, where truly subwavelength optical mode confinement can be achieved. As an example, subwavelength structures from traditional plasmonic metals can only support the LSP resonances discussed earlier when |εm|~εd, limiting the LSP resonance frequency to shorter wavelengths. However, when such structures are fabricated with ‘metals’ designed to have |εm|~εd in the mid-IR, LSP resonances can be supported at longer wavelengths. Subwavelength puck structures fabricated from highly doped InAs films show a just such a resonance (Figure 22), in agreement with both numerical simulations, as well as an analytical model of the puck structure as a subwavelength metallic nanoparticle [3]. As would be expected, and as has been demonstrated for metallic nanoparticles at shorter wavelengths, the spectral location of the resonance does not depend on the size of the puck structure, indicating that the observed resonance is, in fact, a true LSP resonance.

Figure 22 Experimental transmission, simulated absorption, and modeled absorption for 1.7 μm-diameter pucks of heavily-doped InAs, showing a LSP resonance at 9 µm. Figure adapted from [3].
Figure 22

Experimental transmission, simulated absorption, and modeled absorption for 1.7 μm-diameter pucks of heavily-doped InAs, showing a LSP resonance at 9 µm. Figure adapted from [3].

4.3 Designer metal ENZ- and meta-materials

The ability to grow plasmonic materials with designed optical properties opens the door to a wide range of materials and structures demonstrating unique optical phenomena. Key among these are epsilon near zero (ENZ) materials and hyperbolic metamaterials. The unique properties of ideal ENZ materials, from guiding fields around cloaked space to squeezing light through deep subwavelength openings, have their origin in the ability of this media to drastically increase the local wavelength of light [150–152]. The increase of the local wavelength causes the electric field to be homogeneous over areas that can be much larger than the vacuum wavelength of incident light. As a consequence, electromagnetic energy can relatively easily bend around the geometric obstacles formed by “conventional” materials in the ENZ matrix. In particular, such energy “tunneling” leads to the dramatic increase of light transmission through subwavelength openings filled by the ENZ material.

As a demonstration of this effect, sub-wavelength apertures were formed in Au over a highly doped InAsSb layer λENZ~8 µm [4]. At ENZ, for TM polarized light (which is the only polarization which can couple to the waveguide formed by the subwavelength Au slit), a strong enhancement of transmitted light is observed, as shown in Figure 23B. In fact, when compared to transmission through undoped GaAs (Figure 23A), transmission is actually enhanced, indicating that the funneling effect is large enough to overcome the significant losses associated with the large carrier concentrations in the InAsSb.

Figure 23 Transmission of TM (red) and TE (blue) light through (A) a 1.0 μm slit fabricated in a Au layer on an undoped GaAs substrate and (B) a 0.9 μm metal slit fabricated in a Au layer on an epitaxially grown heavily-doped InAsSb ENZ layer. Note the strong transmission peak at λENZ~8 µm for the TM polarized light in (B). Schematic representations of light transmission through a 1 µm gold slit on GaAs (top) and InAsSb (bottom) are shown between (A) and (B).
Figure 23

Transmission of TM (red) and TE (blue) light through (A) a 1.0 μm slit fabricated in a Au layer on an undoped GaAs substrate and (B) a 0.9 μm metal slit fabricated in a Au layer on an epitaxially grown heavily-doped InAsSb ENZ layer. Note the strong transmission peak at λENZ~8 µm for the TM polarized light in (B). Schematic representations of light transmission through a 1 µm gold slit on GaAs (top) and InAsSb (bottom) are shown between (A) and (B).

Highly doped semiconductor layers have also been utilized as plasmonic layers in optically thick hyperbolic metamaterial structures consisting of alternating layers of undoped and highly doped AlInAs/GaInAs. As mentioned earlier, the use of doped semiconductors as plasmonic materials allows for the integration of such material with traditional semiconductors. If grown by MBE (as is the case for the metamaterial work), precise control of layer thicknesses allows for exact control of the metal layers, while control of dopant incorporation allows for control of the metals’ optical properties.

The interleaved layers of doped In0.53Ga0.47As and undoped In0.52Al0.48As, lattice matched to an InP substrate, demonstrated negative refraction for TM-polarized light [5]. For these samples, the onset of negative refraction occurs at the plasma frequency (8.8 μm) and extends to near 11.8 μm. Negative refraction was demonstrated by shining light on the metamaterial while blocking part of the transmitted light with a blade. When the sample exhibits normal refraction, the light is bent away from the blade and more transmitted light is seen. When the sample exhibits negative refraction, the transmitted light is bent toward the blade, resulting in a reduction in the transmitted light intensity, shown in Figure 24. Observation of negative refraction validates the hyperbolic dispersion in these metamaterials, which could potentially be utilized for subwavelength control of radiation and for emission engineering [5, 153–158].

Figure 24 Ratio of transmission of a partially blocked beam to an unblocked beam for TM-polarized light through a hyperbolic metamaterial consisting of alternating layers of undoped (AlInAs) and heavily doped (InGaAs) semiconductor. The shift in the relative transmission intensity results from the negative refraction of light through the hyperbolic metamaterial layer. Figure from [5].
Figure 24

Ratio of transmission of a partially blocked beam to an unblocked beam for TM-polarized light through a hyperbolic metamaterial consisting of alternating layers of undoped (AlInAs) and heavily doped (InGaAs) semiconductor. The shift in the relative transmission intensity results from the negative refraction of light through the hyperbolic metamaterial layer. Figure from [5].

4.4 Summary

The development of mid-IR designer metals has the potential to dramatically affect the field of mid-IR plasmonics. These materials have the ability to effectively mimic the behavior of traditional short-wavelength plasmonic materials at mid-IR wavelengths. By designing metals with |εm|~εdat mid-IR frequencies, plasmonic structures with either tightly bound SPP modes or LSP resonances can be demonstrated, allowing for greater mode confinement, and in many cases, greater functionality for mid-IR plasmonic devices. However, it is not simply the ability to mimic the behavior of shorter wavelength structures that makes these doped semiconductor materials so powerful. The ability to tune the optical properties of the metal component of a plasmonic structure realizes a design parameter simply not available with traditional plasmonic metals at shorter wavelengths. Moreover, compatibility of these materials with both traditional and new, cutting edge, semiconductor electronic and optoelectronic devices gives, arguably, a much broader range of potential devices and applications to the mid-IR plasmonics researcher than his/her short-wavelength counterpart.

5 Conclusion

The vast majority of plasmonic research has been performed in the visible and near IR regions of the optical spectrum, where both localized and propagating plasmonic modes show the strong field enhancement and mode confinement desired for a variety of sensing, interconnect, and imaging applications. The ability to strongly confine or guide light with subwavelength dimensioned structures also has significant appeal for mid-infrared wavelengths, where such structures could be integrated into sensing, beam steering, thermal signature control, and thermal imaging systems.

Initial work on mid-IR plasmonics focused on structures fabricated with traditional metals. Such structures replicate the far field optical properties of their short wavelength counterparts, suggesting the potential for geometric scaling of plasmonic structures with traditional material systems. Impressive results, such as significant molecular absorption enhancement, quantum cascade laser beam steering and shaping, mid-IR detector enhancement and filtering, and selective thermal emission have been demonstrated from mid-IR plasmonic structures using traditional plasmonic metals.

However, while the scaling of short wavelength plasmonic geometries to longer wavelengths replicates the far field optical properties of the scaled devices, it does not result in a similar subwavelength confinement of the mid-IR optical modes. True scaling requires both a geometric scaling and a new class of metals with plasma frequencies in the mid-IR. The recent demonstrations of graphene and highly doped semiconductors as plasmonic materials have the potential to dramatically alter the way plasmonic devices and structures are designed in the mid-IR wavelength range.

The wide range of vital applications, as well as a new class of sources, detectors, and optoelectronic devices, makes the mid-infrared an exciting, vibrant research frontier. Mid-IR plasmonic structures and devices, be they fabricated with traditional metals, or the new class of ‘designer’ metal materials, may well be an enabling technology, with the potential to serve as a key building blocks in next-generation sensing, imaging, communication and photonic applications. Ultimately, the emerging class of new mid-IR plasmonic metals may allow for more than just accurate mimicking of shorter wavelength plasmonic effects. In fact, these new materials can provide new functionality and open new design parameters not achievable with traditional plasmonic metals.

Traditionally, once a metal is chosen for a particular plasmonic architecture, the metal design space is closed, and little can be done to alter the metals’ fundamental optical properties. Yet the new class of low-loss engineered plasmonic metals allows for an entirely new approach to the design and development of plasmonic structures, where the plasmonic material itself can be engineered to give an optimized response for a given plasmonic architecture. For graphene, the potential for ultra-strong confinement and dynamically tunable plasma frequencies may open new doors for plasmonics in sensing and nano-electronic applications. At the same time, the highly crystalline, wavelength flexible, and low-loss nature of doped semiconductors offer intriguing possibilities. Because these metals are based on traditional semiconductor material systems, highly precise growth, patterning, and fabrication techniques are already well-established, and integration with existing electronic and optoelectronic devices can be achieved. In this sense, the mid-IR offers opportunities for new plasmonic devices, architectures, and fundamental investigations not achievable at shorter wavelengths, and may well serve as the testing grounds for the next generation of plasmonic technologies.


Corresponding author: Daniel Wasserman, University of Illinois Urbana Champaign, Department of Electrical and Computer Engineering, Urbana, IL 68101, USA, e-mail:

The authors would like to acknowledge funding support from the AFOSR YIP (DW # #FA9550-10-1-0226) and NSF MWN (DW #DMR-1210398 and VP #DMR-1209761), NSF ECCS-0925542, and NSF ECCS-1102183.

References

[1] Naik GV, Kim J, Boltasseva A. Oxides and nitrides as alternative plasmonic materials in the optical range. Opt Mater Exp 2011;1:1090–9.10.1364/OME.1.001090Search in Google Scholar

[2] Naik GV, Schroeder JL, Ni X, Kildishev AV, Sands TD, Boltasseva A. Titanium nitride as a plasmonic material for visible and near-infrared wavelengths. Opt Mater Exp 2012;2:478–89.10.1364/OME.2.000478Search in Google Scholar

[3] Law S, Adams DC, Taylor AM, Wasserman D. Mid-infrared designer metals. Opt Exp 2012;20:12155–65.10.1364/OE.20.012155Search in Google Scholar PubMed

[4] Adams DC, Inampudi S, Ribaudo T, Slocum D, Vangala S, Kuhta NA, Goodhue WD, Podolskiy VA, Wasserman D. Funneling light through a subwavelength aperture with epsilon-near-zero materials. Phys Rev Lett 2011;107:133901.10.1103/PhysRevLett.107.133901Search in Google Scholar PubMed

[5] Hoffman AJ, Alekseyev L, Howard SS, Franz KJ, Wasserman D, Podolskiy VA, Narimanov EE, Sivco DL, Gmachl C. Negative refraction in semiconductor metamaterials. Nat Mater 2007;6:946–50.10.1038/nmat2033Search in Google Scholar PubMed

[6] Howells SC, Schlie LA. Transient terahertz reflection spectroscopy of undoped InSb from 0.1 to 1.1 THz. Appl Phys Lett 1996;69:550–2.10.1063/1.117783Search in Google Scholar

[7] Huggard PG, Cluff JA, Moore GP, Shaw CJ, Andrews SR, Keiding SR, Linfield EH, Ritchie DA. Drude conductivity of highly doped GaAs at terahertz frequencies. J Appl Phys 2000;87:2382–5.10.1063/1.372238Search in Google Scholar

[8] Ordal MA, Long LL, Bell RJ, Bell SE, Bell RR, Alexander, Jr. RW, Ward CA. Optical properties of the metals Al, Co, Cu, Au, Fe, Pb, Ni, Pd, Pt, Ag, Ti, and W in the infrared and far infrared. Appl Opt 1983;22:1099–119.10.1364/AO.22.001099Search in Google Scholar

[9] Lynch DW, Hunter WR. Handbook of Optical Constants of Solids. E. D. Palik, ed. (Academic, New York, 1998).Search in Google Scholar

[10] Johnson PB, Christy RW. Optical constants of the noble metals. Phys Rev B 1972;6:4370–9.10.1103/PhysRevB.6.4370Search in Google Scholar

[11] Neuner III B, Korobkin D, Fietz C, Carole D, Ferro G, Shvets G. Critically coupled surface phonon-polariton excitation in silicon carbide. Opt Lett 2009;34:2667–9.10.1364/OL.34.002667Search in Google Scholar PubMed

[12] Hillenbrand R, Taubner T, Keilmann F. Phonon-enhanced light-matter interaction at the nanometre scale. Nature 2002;418:159–62.10.1038/nature00899Search in Google Scholar PubMed

[13] Greffet JJ, Carminati R, Joulain K, Mulet JP, Mainguy S, Chen Y. Coherent emission of light by thermal sources. Nature 2002;416:61–4.10.1038/416061aSearch in Google Scholar

[14] Taubner T, Korobkin D, Urzhumov Y, Shvets G, Hillenbrand R. Near-field microscopy through a SiC superlens. Science 2006;313:1595.10.1126/science.1131025Search in Google Scholar

[15] Ruppin RJ. Electromagnetic Surface Modes. Wiley, Chichester, 1982;345–98.Search in Google Scholar

[16] Landau LD, Lifshitz EM. Electrodynamics of Continuous Media. 1960, Pergamon, Oxford, UK.Search in Google Scholar

[17] Hutter E, Fendler J. Exploitation of localized surface plasmon resonance. Adv Mat 2004;16:1685.10.1002/adma.200400271Search in Google Scholar

[18] Maier SA, Brongersma ML, Kik PG, Meltzer S, Requicha AAG, Atwater HA. Plasmonics: a route to nanoscale optical devices. Adv Mat 2001;13:1501.10.1002/1521-4095(200110)13:19<1501::AID-ADMA1501>3.0.CO;2-ZSearch in Google Scholar

[19] Rand BR, Peumans P, Forrest SR. Long range absorption enhancement in organic tandem thin-film solar cells containing silver nanoclusters. J Appl Phys 2004;96:7519.10.1063/1.1812589Search in Google Scholar

[20] Moskovits M. Surface enhanced spectroscopy. Rev Mod Phys 1985;57:783.10.1103/RevModPhys.57.783Search in Google Scholar

[21] Shalaev VM. Nonlinear Optics of Random Media: Fractal Composites and Metal-Dielectric Films. Springer Tracts in Modern Physics, 2000, 158, Springer, Berlin Heidelberg.10.1007/BFb0109599Search in Google Scholar

[22] Perez-Juste J, Pastoriza-Santos I, Liz-Marzan LM, Mulvaney P. Gold nanorods: synthesis, characterization and applications. Coord Chem Rev 2005;249:1870–901.10.1016/j.ccr.2005.01.030Search in Google Scholar

[23] Bardhan R, Mukherjee S, Mirin NA, Levit S, Norlander PR, Halas NJ. Nanosphere in a nanoshell: a simple nanomatryshka. J Phys Chem C 2010;114:7378.10.1021/jp9095387Search in Google Scholar

[24] Gobin AM, Lee MH, Halas NJ, James WD, Drezek RA, West JL. Near-Infrared resonant nanoshells for combined optical imaging and photothermal cancer therapy. Nano-Lett 2007;7:1929–34.10.1021/nl070610ySearch in Google Scholar

[25] Sherry LJ, Jin R, Mirkin CA, Schatz GC, Van Duyne RP. Localized surface plasmon resonance spectroscopy of single silver triangular nanoprisms. NanoLett 2009;6:2060–5.10.1021/nl061286uSearch in Google Scholar PubMed

[26] Payne EK, Shuford K, Park S, Schatz GS, Mirkin CA. Multipole plasmon resonances in gold nanorods. J Phys Chem B 2006;110:2150.10.1021/jp056606xSearch in Google Scholar PubMed PubMed Central

[27] Stockman MI. Nanoplasmonics: the physics behind the applications. Phys Today 2011;64:39.10.1063/1.3554315Search in Google Scholar

[28] Pendry JB, Martín-Moreno L, García-Vidal FJ. Mimicking surface plasmons with structured surfaces. Science 2004;305:847–8.10.1126/science.1098999Search in Google Scholar PubMed

[29] Williams CR, Andrews SR, Maier SA, Fernández-Dominguez AI, Martín-Moreno L, García-Vidal FJ. Highly confined guiding of terahertz surface plasmon polaritons on structured metal surfaces. Nat Photon 2008;2:175–9.10.1038/nphoton.2007.301Search in Google Scholar

[30] Yu N, Wang QJ, Kats MA, Fan JA, Khanna SP, Li L, Davies AG, Linfield EH, Capasso F. Designer spoof surface plasmon structures collimate terahertz laser beams. Nat Mat 2010;9:730.10.1038/nmat2822Search in Google Scholar PubMed

[31] Rusina A, Durach M, Stockman MI. Theory of spoof plasmons in real metals. Appl Phys A 2010;100:375–8.10.1007/s00339-010-5866-ySearch in Google Scholar

[32] Raether H. Surface Plasmons on Smooth and Rough Surfaces and on Gratings. (Springer Tracts in Modern Physics), 1998, 111.Search in Google Scholar

[33] Barnes WL, Dereux A, Ebbesen TW. Surface plasmon subwavelength optics. Nature 2003;424:824.10.1038/nature01937Search in Google Scholar PubMed

[34] Ebbesen TW, Genet C, Bozhevolnyi SI. Surface-plasmon circuitry. Phys Today 2008;61:44–50.10.1063/1.2930735Search in Google Scholar

[35] Lezec HJ, Degiron A, Devaux E, Linke RA, Martín-Moreno L, Garcia-Vídal FJ, Ebesen TW. Beaming light from a subwavelength aperture. Science 2002;297:820–2.10.1126/science.1071895Search in Google Scholar PubMed

[36] Martín-Moreno L, García-Vidal FJ, Lezec HJ, Degiron A, Ebbesen TW. Theory of highly directional emission from a single subwavelength aperture surrounded by surface corrugations. Phys Rev Lett 2003;90:167401-1–4.10.1103/PhysRevLett.90.167401Search in Google Scholar PubMed

[37] Thongrattanasiri S, Adams DC, Wasserman D, Podolskiy VA. Multiscale beam evolution and shaping in corrugated plasmonic systems. Opt Express 2011;19:9269–81.10.1364/OE.19.009269Search in Google Scholar PubMed

[38] Podolskiy VA, Sarychev AK, Shalaev VM. Plasmon modes in metal nanowires. J Nonlinear Opt Phys Mater 2002;11:65.10.1142/S0218863502000833Search in Google Scholar

[39] Novotny L. Effective wavelength scaling for optical antennas. Phys Rev Lett 2007;98:266802-1–4.10.1103/PhysRevLett.98.266802Search in Google Scholar PubMed

[40] Olmon R, Rang M, Krenz P, Lail B, Saraf L, Boreman G, Raschke M. Determination of electric field, magnetic field and electric current distribution of infared optical antennas: a near field optical vector network analyzer. Phys Rev Lett 2010;105:167403.10.1103/PhysRevLett.105.167403Search in Google Scholar PubMed

[41] Miller DL. Principles of infrared technology: a practical guide to the state of the art. Springer 1994.Search in Google Scholar

[42] Martini R, Bethea C, Capasso F, Gmachl C, Paiella R, Whittaker EA, Hwang HY, Sivco DL, Baillargeon JN, Cho AY. Free-space optical transmission of multimedia satellite data streams using mid-infrared quantum cascade lasers. Electron Lett 2002;38:181–3.10.1049/el:20020122Search in Google Scholar

[43] http://www.epa.gov/ttn/emc/ftir/welcome.html. Accessed 7/5/12.Search in Google Scholar

[44] These data, produced using the program IRTRANS4, were obtained from the UKIRT worldwide web pages (http://www.jach.hawaii.edu/UKIRT/astronomy/utils/atmos-index.html), accessed 7/6/12.Search in Google Scholar

[45] West LC, Eglash SJ. First observation of an extremely large dipole infrared transition within the conduction band of a GaAs quantum well. Appl Phys Lett 1985;46:1156–8.10.1063/1.95742Search in Google Scholar

[46] Levine BF, Choi KK, Bethea CG, Walker J, Malik RJ. New 10 µm infrared detector using intersubband absorption in resonant tunneling GaAlAs superlattices. Appl Phys Lett 1987;50:1092–5.10.1063/1.97928Search in Google Scholar

[47] Berryman KW, Lyon SA, Segev M. Mid-infrared photoconductivity in self-organized InAs quantum dots. Appl Phys Lett 1997;70:1861–3.10.1063/1.118714Search in Google Scholar

[48] Pan D, Towe E, Kennerly S. Normal incidence intersubband (In,Ga)As/GaAs quantum dot infrared photodetectors. Appl Phys Lett 1998;73:1937–9.10.1063/1.122328Search in Google Scholar

[49] Sai-Halasz GA, Esaki L. InAs-GaSb superlattice energy structure and its semiconductor-semimetal transition. Phys Rev B 1978;18:2812–8.10.1103/PhysRevB.18.2812Search in Google Scholar

[50] Smith DL, Mailhoit C. Proposal for strained type II superlattice infrared detectors. J Appl Phys 1987;62:2545–8.10.1063/1.339468Search in Google Scholar

[51] Wei Y, Gin A, Razeghi M, Brown GJ. Advanced InAs/GaSb superlattice photovoltaic detectors for very long wavelength infrared applications. Appl Phys Lett 2002;80:3263–5.10.1063/1.1476395Search in Google Scholar

[52] Maimon S, Wicks GW. nBn detector, an infrared detector with reduced dark current and higher operating temperature. Appl Phys Lett 2006;89:151109–11.10.1063/1.2360235Search in Google Scholar

[53] Rodriguez JB, Plis E, Bishop G, Sharma YD, Kim H, Dawson LR, Krishna S. nBn structure based on InAs/GaSb type-II strained layer superlattices. Appl Phys Lett 2007;91:043514–6.10.1063/1.2760153Search in Google Scholar

[54] Faist J, Capasso F, Sivco DL, Sirtori C, Hutchinson AL, Cho AY. Quantum cascade laser. Science 1994;264:553.10.1126/science.264.5158.553Search in Google Scholar PubMed

[55] Beck M, Hofstetter D, Aellen T, Faist J, Oesterle U, Ilegems M, Gini E, Melchior H. Continuous wave operation of a mid-infrared semiconductor laser at room temperature. Science 2002;295:301.10.1126/science.1066408Search in Google Scholar PubMed

[56] Liu PQ, Hoffman AJ, Escarra MD, Franz KJ, Khurgin JB, Dikmelik Y, Wang XJ, Fan JY, Gmachl CF. Highly power-efficient quantum cascade lasers. Nat Photon 2010;4:95.10.1038/nphoton.2009.262Search in Google Scholar

[57] Bai Y, Slivken S, Kuboya S, Darvish SR, Razeghi M. Quantum cascade lasers that emit more light than heat. Nat Photon 2010;4:99.10.1038/nphoton.2009.263Search in Google Scholar

[58] Commercially available QCLs can be obtained from Daylight Solutions (http://www.daylightsolutions.com/) or Pranalytica (http://www.pranalytica.com/) (both accessed 7/7/12), among other vendors.Search in Google Scholar

[59] Levin IW, Bhargava R. Fourier transform infrared vibrational spectroscopic imaging: integrating microscopy and molecular recognition. Annu Rev Phys Chem 2005;56:429–74, and references therein.10.1146/annurev.physchem.56.092503.141205Search in Google Scholar PubMed

[60] Lahrech A, Bachelot R, Gleyzes P, Boccara AC. Infrared-reflection-mode near-field microscopy using an apertureless probe with a resolution λ/600. Opt Lett 1996;21:1315–7.10.1364/OL.21.001315Search in Google Scholar

[61] Knoll B, Keilmann F. Near-field probing of vibrational absorption for chemical microscopy. Nature 1999;399:134–7.10.1038/20154Search in Google Scholar

[62] Houel J, Homeyer E, Sauvage S, Boucaud P, Dazzi A, Prazeres R, Ortega JM. Midinfrared absorption measured at a λ/400 resolution with an atomic force microscope. Opt Express 2009;17:10887–94.10.1364/OE.17.010887Search in Google Scholar PubMed

[63] Raschke MB, Molina L, Elsaesser T, Kim DH, Knoll W, Hinrichs K. Apertureless near-field vibrational imaging of block-copolymer nanostructures with ultrahigh spatial resolution. ChemPhysChem 2005;6:2197–203.10.1002/cphc.200500218Search in Google Scholar PubMed

[64] Kjoller K, Felts JR, Cook D, Prater CB, King WP. High-sensitivity nanometer-scale infrared spectroscopy using a contact mode microcantilever with an internal resonator paddle. Nanotechnology 2010;21:185705.10.1088/0957-4484/21/18/185705Search in Google Scholar PubMed

[65] Ebbesen TW, Lezec HJ, Ghaemi HF, Thio T, Wolff PA. Extraordinary optical transmission through sub-wavelength hole arrays. Nature 1998;391:667–9.10.1038/35570Search in Google Scholar

[66] Martin-Moreno L, Garcia-Vidal FJ, Lezec HJ, Pellerin KM, Thio T, Pendry JB, Ebbesen TW. Theory of extraordinary optical transmission through subwavelength hole arrays. Phys Rev Lett 2001;86:1114–7.10.1103/PhysRevLett.86.1114Search in Google Scholar PubMed

[67] Liu H, Lalanne P. Microscopic theory of the extraordinary optical transmission. Nature 2008;452:728–31.10.1038/nature06762Search in Google Scholar PubMed

[68] Ye Y-H, Zhang J-Y. Middle-infrared transmission enhancement through periodically perforated metal films. Appl Phys Lett 2004;84:2977–9.10.1063/1.1711166Search in Google Scholar

[69] Williams SM, Stafford AD, Rogers TM, Bishop SR, Coe JV. Extraordinary infrared transmission of Cu-coated arrays with subwavelength apertures: hole size and the transition from surface plasmon to waveguide transmission. Appl Phys Lett 2004;85:1472–5.10.1063/1.1786664Search in Google Scholar

[70] Williams SW, Coe JV. Dispersion study of the infrared transmission resonances of freestanding Ni microarrays. Plasmonics 2006;1:87–93.10.1007/s11468-005-9001-4Search in Google Scholar

[71] Wasserman D, Shaner EA, Cederberg JG. Midinfrared doping-tunable extraordinary optical transmission from subwavelength gratings. Appl Phys Lett 2007;90:191102.10.1063/1.2737138Search in Google Scholar

[72] Bao YJ, Peng RW, Shu DJ, Wang M, Lu X, Shao J, Lu W, Ming NB. Role of interference between localized and propagating surface waves on the extraordinary optical transmission through a subwavelength-aperture array. Phys Rev Lett 2008;101:87401.10.1103/PhysRevLett.101.087401Search in Google Scholar

[73] Ribaudo T, Shaner EA, Howard SS, Gmachl C, Wang XJ, Choa F-S, Wasserman D. Active control and spatial mapping of mid-infrared propagating surface plasmons. Opt Express 2009;17:7019–24.10.1364/OE.17.007019Search in Google Scholar

[74] Ribaudo T, Adams DC, Passmore B, Shaner EA, Wasserman D. Spectral and spatial investigation of midinfrared surface waves on a plasmonic grating. Appl Phys Lett 2009;94:201109.10.1063/1.3140569Search in Google Scholar

[75] Hoffman FM. Infrared reflection-absorption spectroscopy of adsorbed molecules. Surf Sci Reports 1983;3:107–92.10.1016/0167-5729(83)90001-8Search in Google Scholar

[76] Osawa M. Dynamic processes in electrochemical reactions studied by surface-enhanced infrared absorption spectroscopy (SEIRAS). Bull Chem Soc Jpn 1997;70:2861–80, and references therein.10.1246/bcsj.70.2861Search in Google Scholar

[77] Aroca RF, Ross, DJ. Surface-enhanced infrared spectroscopy. Appl Spectros 2004;58:324A–38A.10.1366/0003702042475420Search in Google Scholar PubMed

[78] Osawa M. Surface-enhanced infrared absorption. Topics Appl Phys 2001;81:163–87.10.1007/3-540-44552-8_9Search in Google Scholar

[79] Hatta A, Suzuki Y, Suëtaka W. Infrared absorption enhancement of monolayer species on thin evaporated Ag films by use of a Kretschmann configuration: evidence for two types of enhanced surface electric fields. Appl Phys A 1984;35:135–40.10.1007/BF00616965Search in Google Scholar

[80] Sanda PN, Warlaumont, Demuth JE, Tsang JC, Christmann K, Bradley JA. Surface-enhanced raman scattering from pyridine on Ag(111). Phys Rev Lett 1980;45:1519–23.10.1103/PhysRevLett.45.1519Search in Google Scholar

[81] Golosovsky M, Lirtsman V, Yashunsky V, Davidov D, Aroeti B. Midinfrared surface-plasmon resonance: a novel biophysical tool for studying living cells. J Appl Phys 2009;105:102036.10.1063/1.3116143Search in Google Scholar

[82] Williams SM, Stafford AD, Rodriguez KR, Rogers TM, Coe JV. Accessing surface plasmons with Ni microarrays for enhanced IR absorption by monolayers. J Phys Chem B 2003;107:11871–9.10.1021/jp034934wSearch in Google Scholar

[83] Rodriguez KR, Shah S, Williams SM, Teeters-Kennedy S, Coe JV. Enhanced infrared absorption specra of self-assembled alkanethiol monolayers using the extraordinary infrared transmission of metallic arrays of subwavelength apertures. J Chem Phys 2004;121:8671–5.10.1063/1.1814052Search in Google Scholar PubMed

[84] Rodriguez KR, Tian H, Heer JM, Teeters-Kennedy S, Coe JV. Interaction of an infrared surface plasmon with an excited molecular vibration. J Chem Phys 2007;128:151101-1–5.10.1063/1.2730781Search in Google Scholar PubMed

[85] Osawa M, Matsuda N, Yoshii K, Uchida I. Charge transfer resonance raman process in surface-enhanced raman scattering from p-aminothiophenol adsorbed on silver: Herzberg-Teller contribution. J Phys Chem 1994;98: 12702–7.10.1021/j100099a038Search in Google Scholar

[86] Jeanmaire DL, Van Duyne RP. Surface raman electrochemistry. Part 1. Heterocyclic, aromatic and aliphatic amines adsorbed on the anodised silver electrode. J Electroanal Chem 1977;84:1–20.10.1016/S0022-0728(77)80224-6Search in Google Scholar

[87] http://www.biacore.com/lifesciences/products/systems_overview/index.html. Accessed 7/26/2012.Search in Google Scholar

[88] Yanik AA, Adato R, Erramilli S, Altug H. Hybridized nanocavities as single-polarized plasmonic antennas. Opt Express 2009;17:20900–10.10.1364/OE.17.020900Search in Google Scholar PubMed

[89] Anglin K, Ribaudo T, Adams DC, Qian X, Goodhue WD, Dooley S, Shaner EA, Wasserman D. Voltage-controlled active mid-infrared plasmonic devices. J Appl Phys 2011;109:123103.10.1063/1.3600230Search in Google Scholar

[90] Willets KA, Van Duyne RP. Localized surface plasmon resonance spectroscopy and sensing. Annu Rev Phys Chem 2007;58:267–97.10.1146/annurev.physchem.58.032806.104607Search in Google Scholar PubMed

[91] Adato R, Yanik AA, Amsden JJ, Kaplan DL, Omenetto FG, Hong MK, Erramilli S, Altug H. Ultra-sensitive vibrational spectroscopy of protein monolayers with plasmonic nanoantenna arrays. Proc Natl Acad Sci USA 2009;106: 19227–32.10.1073/pnas.0907459106Search in Google Scholar PubMed PubMed Central

[92] Neubrech F, Pucci A, Cornelius TW, Karim S. Resonant plasmonic and vibrational coupling in a tailored nanoantenna for infrared detection. Phys Rev Lett 2008;101:157403.10.1103/PhysRevLett.101.157403Search in Google Scholar PubMed

[93] Mason JA, Allen G, Podolskiy VA, Wasserman D. Strong coupling of molecular and mid-infrared perfect absorber resonances. IEEE Photon Technol Lett 2012;24:31–3.10.1109/LPT.2011.2171942Search in Google Scholar

[94] Yu Z, Veronis G, Fan S, Bringersma ML. Design of midinfrared photodetectors enhanced by surface plasmons on grating structures. Appl Phys Lett 2006;89:151116.10.1063/1.2360896Search in Google Scholar

[95] Hu X, Li M, Ye Z, Leung WY, Ho KM, Lin SY. Design of midinfrared photodetectors enhanced by resonant cavitities with subwavelength metallic gratings. Appl Phys Lett 2008;93:241108.10.1063/1.3052893Search in Google Scholar

[96] Chang CY, Chang HY, Chen CY, Tsai MW, Chang YT, Lee SC, Tang SF. Wavelength selective quantum dot infrared photodetector with periodic metal hole arrays. Appl Phys Lett 2007;91:163107.10.1063/1.2800378Search in Google Scholar

[97] Lee SC, Krishna S, Brueck SRJ. Quantum dot infrared photodetector enhanced by surface plasma wave excitation. Opt Express 2009;17:23160–8.10.1364/OE.17.023160Search in Google Scholar PubMed

[98] Rosenberg J, Shenoi R, Vandervelde TE, Krishna S, Painter O. A multispectral and polarization-selective surface plasmon resonant midinfrared detector. Appl Phys Lett 2009;95:161101.10.1063/1.3244204Search in Google Scholar

[99] Wu W, Bonakdar A, Mohseni H. Plasmonic enhanced quantum well infrared photodetector with high detecticity. Appl Phys Lett 2010;96:161107.10.1063/1.3419885Search in Google Scholar

[100] Lee SJ, Ku Z, Barve A, Montoya J, Jang WY, Brueck SRJ, Sundaram M, Reisinger A, Krishna S, Noh SK. A monolithically integrated plasmonic infrared quantum dot camera. Nat Commun 2011;2:286. doi: 10.1038/ncomms1283.10.1038/ncomms1283Search in Google Scholar PubMed

[101] Sirtori C, Gmachl C, Capasso F, Faist J, Sivco DL, Hutchinson AL, Cho AY. Long-wavelength (λ ≈8-11.5µm) semiconductor lasers with waveguides based on surface plasmons. Opt Lett 1998;23:1366–8.10.1364/OL.23.001366Search in Google Scholar

[102] Tredicucci A, Gmachl C, Capasso F, Hutchinson AL, Sivco DL, Cho AY. Single-mode surface-plasmon laser. Appl Phys Lett 2000;76:2164–6.10.1063/1.126183Search in Google Scholar

[103] Collombelli R, Capasso F, Gmachl C, Hutchinson AL, Sivco DL, Tredicucci A, Wanke MW, Sergent AM, Cho AY. Far-infrared surface-plasmon quantum-cascade lasers at 21.5 µm and 24 µm wavelengths. Appl Phys Lett 2001;78:2620–2.10.1063/1.1367304Search in Google Scholar

[104] Huang X, Chiu Y, Charles WO, Gmachl C. Ridge-width dependence of the threshold of long wavelength (λ≈14µm) Quantum Cascade lasers with sloped and vertical sidewalls. Opt Express 2012;20:2539–47.10.1364/OE.20.002539Search in Google Scholar PubMed

[105] Babuty A, Bousseksou A, Tetienne J-P, Moldovan-Doyen I, Sirtori C, Beaudoin G, Sagnes I, De Wilde Y, Colombelli R. Semiconductor surface plasmon sources. Phys Rev Lett 2010;104:226806.10.1103/PhysRevLett.104.226806Search in Google Scholar PubMed

[106] Tetienne J-P, Bousseksou A, Costantini D, Colombelli R, Babuty A, Moldavan-Doyen I, De Wilde W, Sirtori C, Beaudoin G, Largeau L, Mauguin O, Sagnes I. Injection of midinfrared surface plasmon polaritons with an integrated device. Appl Phys Lett 2010;97:211110.10.1063/1.3519985Search in Google Scholar

[107] Tetienne J-P, Bousseksou A, Costantini D, De Wilde Y, Colombelli R. Design of an integrated coupler for the electrical generation of surface plasmon polaritons. Opt Express 2011;19:18155–63.10.1364/OE.19.018155Search in Google Scholar PubMed

[108] Yu N, Wang Q, Capasso F. Beam engineering of quantum cascade lasers. Laser Photonics Rev 2012;6:24–46.10.1002/lpor.201100019Search in Google Scholar

[109] Yu N, Fan J, Wang QJ, Pflügl C, Diehl L, Edamura T, Yamanishi M, Kan H, Capasso F. Small-divergence semiconductor lasers by plasmonic collimation. Nat Photon 2008;2:564–70.10.1038/nphoton.2008.152Search in Google Scholar

[110] Yu N, Cubukcu E, Diehl L, Belkin MA, Crozier KB, Capasso F. Plasmonic quantum cascade laser antenna. Appl Phys Lett 2007;91:173113.10.1063/1.2801551Search in Google Scholar

[111] Adams DC, Thongrattanasiri S, Ribaudo T, Podolskiy VA, Wasserman D. Plasmonic mid-infrared beam steering. Appl Phys Lett 2010;96:201112.10.1063/1.3431665Search in Google Scholar

[112] Yu N, Wang QJ, Pflügl C, Diehl L, Capasso F, Edamura T, Furuta S, Yamanishi M, Kan H. Semiconductor lasers with integrated plasmonic polarizers. Appl Phys Lett 2009;94:151101.10.1063/1.3093476Search in Google Scholar

[113] Yu N, Genevet P, Kats MA, Aieta F, Tetienne J-P, Capasso F, Gaburro Z. Light propagation with phase discontinuities: generalized laws of reflection and refraction. Science 2011;334:333.10.1126/science.1210713Search in Google Scholar PubMed

[114] Larouche S, Smith DR. Reconciliation of generalized refraction with diffraction theory. Opt Lett 2012;37:2391.10.1364/OL.37.002391Search in Google Scholar PubMed

[115] Ni X, Emani NK, Kildishev AV, Boltasseva A, Shalaev VM. Broadband light bending with plasmonic nanoantennas. Science 2012;335:427.10.1126/science.1214686Search in Google Scholar PubMed

[116] Kats MA, Genevet P, Aoust G, Yu N, Blanchard R, Aieta F, Gaburro Z, Capasso F. Giant birefringence in optical antenna arrays with widely tailorable optical anisotropy. Proc Natl Acad Sci USA 2012;109:12364.10.1073/pnas.1210686109Search in Google Scholar

[117] Mason JA, Adams DC, Johnson Z, Smith S, Davis AW, Wasserman D. Selective thermal emission from patterned steel. Opt Express 2010;18:25192–8.10.1364/OE.18.025192Search in Google Scholar PubMed

[118] Ikeda K, Miyazaki HT, Kasaya T, Yamamoto K, Inoue Y, Fujimura K, Kanakugi T, Okada M, Hatade K, Kitagawa S. Controlled thermal emission of polarized infrared waves from arrayed plasmon nanocavities. Appl Phys Lett 2008;92:021117.10.1063/1.2834903Search in Google Scholar

[119] Miyazaki HT, Ikeda K, Kasaya Y, Yamamoto K, Inoue Y, Fujimura K, Kanakugi T, Okada M, Hatade K, Kitagawa S. Thermal emission of two-color polarized infrared waves from integrated plasmon cavitites. Appl Phys Lett 2008;92:141114.10.1063/1.2906375Search in Google Scholar

[120] Liu N, Mesch M, Weiss T, Hentschel M, Giessen H. Infrared perfect absorber and its application as plasmonic sensor. Nano Lett 2010;10:2342.10.1021/nl9041033Search in Google Scholar PubMed

[121] Wu C, Neuner III B, Shvets G, John J, Milder A, Zollars B, Savoy S. Large-area wide-angle spectrally selective plasmonic absorber. Phys Rev B 2011;84:075102.10.1103/PhysRevB.84.075102Search in Google Scholar

[122] Jian ZH, Yun S, Toor F, Werner DH, Mayer TS. Conformal dual-band near-perfectly absorbing mid-infrared metamaterial coating. ACS Nano 2011;5:4641.10.1021/nn2004603Search in Google Scholar PubMed

[123] Liu X, Tyler T, Starr T, Starr AF, Jokerst NM, Padilla WJ. Taming the blackbody. Phys Rev Lett 2011;107:045901.10.1103/PhysRevLett.107.045901Search in Google Scholar PubMed

[124] Mason JA, Smith S, Wasserman D. Strong absorption and selective thermal emission from midinfrared metamaterials. Appl Phys Lett 2011;98:241105.10.1063/1.3600779Search in Google Scholar

[125] Novoselov KS, Geim AK, Morozov SV, Jiang D, Zhang Y, Dubonos SV, Grigorieva IV, Firsov AA. Electric field effect in atomically thin carbon films. Science 2004;306:666–9.10.1126/science.1102896Search in Google Scholar PubMed

[126] Novoselov KS, Jiang D, Schedin F, Booth TJ, Khotkevich VV, Morozov SV, Geim AK. Two-dimensional atomic crystals. Proc Natl Acad Sci USA 2005;102:10451–310.1073/pnas.0502848102Search in Google Scholar PubMed PubMed Central

[127] Novoselov KS, Geim AK, Morozov SV, Jiang D, Katsnelson MI, Grigorieva IV, Dubonos SV, Firsov AA. Two-dimensional gas of massless Dirac fermions in graphene. Nature 2005;438: 197–200.10.1038/nature04233Search in Google Scholar PubMed

[128] Koppens FHL, Chang DE, Javier Garcia de Abajo F. Graphene plasmonics: a platform for strong light-matter interactions. NanoLett 2011;11:3370.10.1021/nl201771hSearch in Google Scholar PubMed

[129] Jablan M, Buljan H, Soljacic M. Plasmonics in graphene at infrared frequencies. Phys Rev B 2009;80:245435.10.1103/PhysRevB.80.245435Search in Google Scholar

[130] Grigorenko AN, Polini M, Novoselov KS. Graphene plasmonics. Nat Photon 2012;6:749–58.10.1038/nphoton.2012.262Search in Google Scholar

[131] Panchakarla LS, Subrahmanyam KS, Saha SK, Govindaraj A, Krishnamurthy HR, Waghmare UV, Rao CNR. Synthesis, structure, and properties of boron- and nitrogen-doped graphene. Adv Mater 2009;21:4726–30.10.1002/adma.200901285Search in Google Scholar

[132] Das A, Pisana S, Chakraborty B, Piscane S, Saha SK, Waghmare UV, Novoselov KS, Krishnamurthy HR, Geim AK, Ferrari AC, Sood AK. Monitoring dopants by Raman scattering in an electrochemically top-gated graphene transistor. Nat Nanotech 2008;3:210–5.10.1038/nnano.2008.67Search in Google Scholar PubMed

[133] Wang F, Zhang Y, Tian Chuanshan, Girit C, Zettl A, Crommie M, Shen YR. Gate-variable optical transitions in graphene. Science 2008;320:206–9.10.1126/science.1152793Search in Google Scholar PubMed

[134] Li ZQ, Henrikson EA, Jiang Z, Hao Z, Martin MC, Kim P, Stormer HL, Basov DN. Dirac charge dynamics in graphene by infrared spectroscopy. Nat Phys 2008;4:532–5.10.1038/nphys989Search in Google Scholar

[135] Ju L, Geng B, Horng J, Girit C, Martin M, Hao Z, Bechtel HA, Liang X, Zettl A, Shen YR, Wang F. Graphene plasmonics for tunable terahertz metamaterials. Nat Nanotechnol 2011;6:630.10.1038/nnano.2011.146Search in Google Scholar PubMed

[136] Fei Z, Rodin AS, Andreev GO, Bao W, McLeod AS, Wagner M, Zhang LM, Zhao Z, Thiemens M, Dominguez G, Fogler MM, Castro Neto AH, Lau CN, Keilmann F, Basov DN. Gate-tuning of graphene plasmons revealed by infrared nano-imaging. Nature 2012;487:82.10.1038/nature11253Search in Google Scholar PubMed

[137] Chen J, Badioli M, Alonso-Gonzalez P, Thongrattanasiri S, Huth F, Osmond J, Spaseovic M, Centeno A, Pesquera A, Godignon P, Zurutuza Elorza A, Camara N, Garcia de Abajo FJ, Hillenbrand R, Koppens FHL. Optical nano-imaging of gate-tunable graphene plasmons. Nature 2012;487:77.10.1038/nature11254Search in Google Scholar PubMed

[138] Yan H, Li X, Chandra B, Tulevski G, Wu Y, Freitag M, Zhu W, Avouris P, Xia F. Tunable infrared plasmonic devices using graphene/insulator stacks. Nat Nanotechnol 2012;7:330.10.1038/nnano.2012.59Search in Google Scholar PubMed

[139] Naik GV, Liu J, Kildishev AV, Shalaev VM, Boltasseva A. Demonstration of Al:ZnO as a plasmonic component for near-infrared metamaterials. Proc Natl Acad Sci USA 2012;109:8834–8.10.1073/pnas.1121517109Search in Google Scholar PubMed PubMed Central

[140] Khurgin JB, Boltasseva A. Reflecting upon the losses in plasmonics and metamaterials. MRS Bull 2012;37:768–79.10.1557/mrs.2012.173Search in Google Scholar

[141] Marquier F, Joulain K, Mulet JP, Carminati R, Greffet JJ. Engineering infrared emission properties of silicon in the near field and the far field. Opt Commun 2004;237:379–88.10.1016/j.optcom.2004.04.024Search in Google Scholar

[142] Soref R, Hendrickson J, Cleary JW. Mid- to long-wavelength infrared plasmonic photonics using heavily doped n-Ge/Ge and n-GeSn/GeSn heterostructures. Opt Express 2012;20: 3814–24.10.1364/OE.20.003814Search in Google Scholar PubMed

[143] Ginn JC, Jarecki RL, Shaner EA, Davids PS. Infrared plasmons on heavily-doped silicon. J Appl Phys 2011;110: 043110.10.1063/1.3626050Search in Google Scholar

[144] Shahzad M, Medhi G, Peale RE, Buchwald WR, Cleary JW, Soref R, Boreman GD, Edwards O. Infrared surface plasmons on heavily doped silicon. J Appl Phys 2011;110:123105.10.1063/1.3672738Search in Google Scholar

[145] Li D, Ning CZ. All-semiconductor active plasmonic system in mid-infrared wavelengths. Opt Express 2011;19:14594–603.10.1364/OE.19.014594Search in Google Scholar PubMed

[146] Tokumitsu E. Correlation between Fermi level stabilization positions and maximum free carrier concentrations in III-V compound semiconductors. Jpn J Appl Phys 1990;29: L698–701.10.1143/JJAP.29.L698Search in Google Scholar

[147] Zhang S. The microscopic origin of the doping limits in semiconductors and wide-gap materials and recent developments in overcoming these limits: a review. J Phys Condens Matter 2002;14:R881–903.10.1088/0953-8984/14/34/201Search in Google Scholar

[148] Moss TS. The interpretation of the properties of indium arsenide. Proc Phys Soc B 1954;67:775–82.10.1088/0370-1301/67/10/306Search in Google Scholar

[149] Burstein E. Anomalous optical absorption limit in InSb. Phys Rev 1954;93:632–3.10.1103/PhysRev.93.632Search in Google Scholar

[150] Ziolkowski RW. Propagation in and scattering from a matched metamaterial having a zero index of refraction. Phys Rev E 2004;70:046608.10.1103/PhysRevE.70.046608Search in Google Scholar PubMed

[151] Alu A, Engheta N. Light squeezing through arbitrarily shaped plasmonic channels and sharp bends. Phys Rev B 2008;78:035440.10.1103/PhysRevB.78.035440Search in Google Scholar

[152] Silveirinha M, Engheta N. Tunneling of electromagnetic energy through subwavelength channels and bends using ε- near-zero materials. Phys Rev Lett 2006;97:157403.10.1103/PhysRevLett.97.157403Search in Google Scholar PubMed

[153] Podolskiy VA, Narimanov EE. Strongly anisotropic waveguide as a nonmagnetic left-handed system. Phys Rev B 2005;71:201101(R).10.1103/PhysRevB.71.201101Search in Google Scholar

[154] Jacob Z, Alekseyev LV, Narimanov EE. Optical hyperlens: far-field imaging beyond the diffraction limit. Opt Exp 2006;14:8247–56.10.1364/OE.14.008247Search in Google Scholar

[155] Liu Z, Lee H, Xiong Y, Sun C, Zhang X. Far-field optical hyperlens magnifying sub-diffraction-limited objects. Science 2007;315:1686.10.1126/science.1137368Search in Google Scholar PubMed

[156] Salandrino A, Engheta N. Far-field subdiffraction optical microscopy using metamaterial crystals: theory and simulations. Phys Rev B 2006;74:075103.10.1103/PhysRevB.74.075103Search in Google Scholar

[157] Cortes CL, Newman W, Molesky S, Jacob Z. Quantum nanophotonics using hyperbolic metamaterials. J Opt 2012;14:063001.10.1088/2040-8978/14/6/063001Search in Google Scholar

[158] Krishnamoorthy HNS, Jacob Z, Narimanov EE, Kretzschmar I, Menon VM. Topological transitions in metamaterials. Science 2012;336:205.10.1126/science.1219171Search in Google Scholar PubMed

Received: 2012-09-11
Accepted: 2012-11-30
Published Online: 2013-01-05
Published in Print: 2013-04-01

©2013 by Science Wise Publishing & De Gruyter Berlin Boston

Downloaded on 3.8.2024 from https://www.degruyter.com/document/doi/10.1515/nanoph-2012-0027/html
Scroll to top button