Expand this Topic clickable element to expand a topic
Skip to content
Optica Publishing Group

Numerical analysis of the propagation properties of subwavelength semiconductor slit in the terahertz region

Open Access Open Access

Abstract

The propagation properties of terahertz (THz) waves passing through heavily doped semiconductor slit have been numerically investigated by using the transfer matrix method. The effects of geometrical parameters, carrier concentration, and dielectric materials filling in the slit have been considered. The contour for carrier concentration and slit width show that as slit width and carrier concentration decreases, the effective indices increase and the propagation lengths decrease. For the case of water filling in the slit, temperature has more effect on the imaginary part of propagation constant than the real part. Most of the energy stored in the slit is in the form of electric energy, which firstly decreases and then increases with the decreasing of slit width. It is expected that the semiconductor slit structure is very useful for the practical applications of THz waves in the fields of biological specimen analysis and medical diagnosis.

©2009 Optical Society of America

1. Introduction

Terahertz (THz) waves show great potential in many scientific research and application fields, such as free space communication, environmental sensing, medical imaging, and biology detection [13]. For instance, THz waves can be used to detect skin cancer and living cell membranes because many biological molecules demonstrate the distinctive optical response in the THz region. The development of THz technology is largely boosted by the rapid development of radiation sources [47] and detectors [8], while THz waveguide propagation methods and the design of relevant waveguide device are also of very importance. Firstly proposed by Wang and Mittleman [9], bare metal wire can be used to confine and propagate THz waves, showing the merits of simple structure and low attenuation [10,11]. Parallel-plate metal waveguide is another important kind of THz waveguide methods and demonstrates the advantages of low group velocity dispersion, low loss and no cutoff frequency [1214]. Besides those waveguide methods, many relevant waveguide components, such as filter, polarizer and attenuator, are also closely related to the applications of surface plasmon polaritons (SPPs) [15,16]. SPPs are quasi two dimensional electromagnetic excitations bounded to the metal-dielectric interface, exhibiting subwavelength confinement and offering the possibility of realizing subwavelength waveguide [17,18].

Gap (slit) surface polariton plasmons (GSPPs) waveguide is an important kind of subwavelength plasmonic devices and can be regarded as the metal-dielectric-metal (MDM) structure. Much research has been carried out to investigate the MDM structure in the visible [13,19,20], infrared, and THz region [2124]. MDM plasmonic waveguide demonstrates the merits of strong subwavelength localization of plasmon, weak dissipation, the possibility of single mode operation [25], showing potential in many practical applications fields. For instance, MDM plasmonic structure can be used to conduct single-molecule analysis requiring pico- to nanomolar concentrations of fluorophore [26], fabricate optical tweezers for transporting micrometer or nanometer dielectric particles (water molecular or DNA molecules) [27]. The solute (alcohol, sugar) concentration of the aqueous solution can be determined by measuring the changes of the dielectric properties of liquids compressed between the metal plate [28]. The thickness and refractive index of the nanometer water-layer [29] can also be acquired by comparing the dielectric constant from the empty guide with that from the waveguide containing the dielectric layer. With biological analyte compressed between the plates, MDM or modified MDM structure can also be utilized to develop SPPs biosensors by probing the interaction between analyte and THz waves [30].

Due to the fact that metals always show large dielectric constant in the THz region, the decay length of SPPs mode for metal-based MDM structure is very long, which weakens the SPPs mode and limits its practical applications. One of the possible solution methods is adopting active dielectric core materials [3134]. In the THz region, the imaginary part of many dielectric material, such as doped semiconductor (GaAs), biological analyte containing water or other polar molecules, are always large, which is different from the case in the visible and infrared spectral region. To Investigate the propagation properties of MDM structure with complex dielectric constant core materials is very interesting and important. For the plasma frequencies within far-infrared, heavily doped semiconductors also show metallic characters in the THz region, and their dielectric constant are similar to that of metal (Ag and Au) in visible/UV spectral range. Heavily doped semiconductors have many merits and show more flexibility in fabricating waveguide devices. Furthermore, the SPPs modes on heavily doped semiconductor is significantly more sensitive to the dielectric layer than surface modes supported on a metal substrate [3]. Indium Antimonide (InSb) is a kind of narrow gap semiconductor and has high intrinsic electron density at room temperature. Therefore, the propagation properties of GSPPs mode in the THz region based on heavily doped semiconductor slit have been shown and discussed. In addition, the effects of geometrical parameter, slit materials, and the core dielectric materials with complex refractive index on waveguide propagation properties have also been explored.

2. Theoretic Model and Research Method

The MDM plasmonic structure has been schematically shown in Fig. 1 , which can be regarded as an infinitely long slit under the illumination of THz waves. The slit is filled with dielectric gain material in region 2 and bounded by semi-infinite regions of heavily doped InSb in region 1 and 3. The slit width is w, and the thickness of the InSb film is enough large. For this slit like structure, the transfer matrix method (TMM) has been used to acquire the propagation constant of GSPPs mode [35,36]. Compared with the analytic equation method [13,16,37], TMM has the merits of high efficiency and accuracy. Under the framework of TMM, the electric or magnetic field can be written as the summation of the incident waves and the reflected waves in each layer of planar multilayered structure. The magnitude of magnetic field of incident light at the first layer and that of the reflected waves at the last layer could be related via the following transfer matrix [35]:

γ2γ3γNε1ε2εN1[HN+HN]N,N1=(M11M12M21M22)[H1+H1]2,1, Eq.
where the matrix product of characteristic matrices could be expressed as:
(M11M12M21M22)=MN1MN2M1, Eq.
Mi is the characteristic matrices in the ith layer,
Mi=(εi1εiεi1εi)(γiγiγi1γi1)(exp(γidi)00exp(γidi)), Eq.
Hi+and Hi are the magnetic field of the propagation and reflected waves in the ith layer, di is the thickness of the ith layer, εi is the dielectric constant of the ith layer, the factor of γicould be expressed as [35]:
γj2=β2k02εi,i=1,2,3, Eq.
in which β is the z component complex propagation constant of fundamental mode in the gap. Because the electromagnetic field vanish at the top and bottom layer, the coefficients Hi+and Hi should be zero, leading to M11 = 0 .

 figure: Fig. 1

Fig. 1 Geometry of an infinitely long slit under illumination of THz waves. The slit width is w, which is filled with gain dielectric material in region 2 and bounded by semi-infinite regions of heavily doped InSb in region 1 and 3.

Download Full Size | PDF

The field component of transverse magnetic (TM) mode in the slit materials (|x|>w/2) can be written as [38]:

Exm(r,ω)=exp(α1,3|x|)×exp(i(βzωt)), Eq.
Hym(r,ω)=ωε0εmβexp(α1,3|x|)×exp(i(βzωt)), Eq.
Ezm(r,ω)=iα1,3β|x|xexp(α1,3|x|)×exp(i(βzωt)), Eq.
and in the slit (i.e. (|x|<w/2)),
Exd(r,ω)=2cosh(α2|x|)×exp(i(βzωt)), Eq.
Hyd(r,ω)=2ωε0ε2βcosh(α2|x|)×exp(i(βzωt)), Eq.
Ezd(r,ω)=2iα2βsinh(α2|x|)×exp(i(βzωt)), Eq.
where k is the wave vector, Eqs. (3)-(8) should be normalized with a common E0.

The electric and magnetic energy densities in the dielectric core are [38]:

uei(x,ω)=14ε0εi|εm(ω)εiexp(kmT/2)cosh(kiT/2)|2|A|2(|cosh(kix)|2+|kiksinh(kix)2|), Eq.
umi(x,ω)=14ε0|ωεm(ω)ckexp(kmT/2)cosh(kiT/2)|2|A|2(|cosh(kix)|2), Eq.
while the electric and magnetic energy densities in the heavily doped semiconductor are:

uem(x,ω)=14ε0Re((ωεm(ω))ω)|Aexp(km|x|)|2(1.0+|kmk|2), Eq.
umm(x,ω)=14ε0|Aωεm(ω)ckexp(km|x|)|2, Eq.

where km=β2εmk02, ki=β2εdk02, and the factor of A should be chosen to make the following Eq. (38):

Uei+Umi+Uem+Umm=1, Eq. (15) in which Uei and Umi are the total electric energy and magnetic energy in the dielectric core, Uem and Umm are the total electric energy and magnetic energy in the metal, which could be acquired by integrating energy density above Eqs. (11)-(14).

The permittivity of heavily doped InSb in the THz region could be expressed as [39]:

ε(ω)=(εωp2ω2+ωτ2)+iωτωp2ω(ω2+ωτ2), Eq.
where ε is the high frequency permittivity, ωp and ωτ are the plasma frequency and damping frequency, respectively. The relevant parameters for heavily doped InSb are taken from Ref. 39. It should be noted that the values of ωp and ωτ for doped InSb change with temperature and doping concentration. The dielectric constant of water in the THz region change with frequency and temperature, which could be well described by following double Debye model [40]:
ε(ω)=ε+εsε11+iωτD+ε1ε1+iωτ2, Eq.
in which
εs=87.91e0.00458T, Eq.
the fast relaxation time τ2 (in hundreds of femtoseconds) and the slow relaxation time τD (in picoseconds) are related to the reorientation time of a single water molecule with a small moment of inertia and the breakage of hydrogen bonds, respectively, the relevant parameters were taken from [40]. The dielectric constant of GaAs in the THz region can be written as [41]:
ε(ω)=ε+ine2τωm(1iωτ), Eq.
where ε is set as 12.96, τ is the Drude relaxation times, m is the effective mass, n is the carrier concentration. The skin depth in the heavily doped semiconductor can be expressed as [39]:
δsemic0ω(Re(εsemi)+1Re(εsemi)2)1/2. Eq.
The effective index neff and propagation length L can be expressed as [16]:

neff=Re(β)/k0, Eq.
L=[2Im(β)]1. Eq.

3. Results and discussion

Figure 2(a) shows the effective indices and propagation lengths of GSPPs mode with different core dielectric materials. The dielectric materials filling in the slit are air, low-density polyethylene, water, and GaAs, with the corresponding refractive index of 1.0, 1.51 [42], 2.375 + 0.502i [40], and 3.284 + 0.106i [41], respectively; the carrier concentration of InSb is 8.0 × 1015cm−3; the radiation frequency is 1.0 THz. It could be found that as the slit width decreases, the effective indices increase and the propagation lengths decrease, which may result from the fact that the fraction of total electromagnetic energy of GSPPs mode residing in the InSb increases when slit width become smaller. The effective indices of GSPPs mode increase with the increasing of the real part of permittivity for dielectric materials filling in the slit, which may result from the fact that the fraction of GSPPs mode pushed into InSb layer increases. It could also be found from Fig. 2(a) that the propagation length is closely relate to the imaginary part of permittivity for dielectric materials filling in the slit. For metal MDM structure, the large propagation length in the THz region is one of the drawbacks to limit its application. It can be learned from our earlier publication [13] that the propagation length of GSPPs mode of heavily doped InSb is much smaller than that of metal structure. The propagation length could also be largely reduced by filling the slit with different dielectric materials, which has been shown in Fig. 2(a). For example, the propagation length are 1.28 × 10⁴μm and 2.06 × 102 μm for air and GaAs filling in the slit (the slit width is 100 μm). The absorbing core dielectric materials lead to the propagation length of SPPs reducing significantly, which is according with the results in Ref. 3. The GSPPs mode can be better confined by filling dielectric materials (water or GaAs) in the slit, which is very interesting and important for the application of semiconductor made MDM structure. Figure 2(b) displays that the effects of imaginary part of permittivity of core dielectric materials on the dispersive properties, the dielectric material filling in the slit is GaAs. The dielectric constant of GaAs can be changed with carrier concentration, shown in Eq. (19). The real part of dielectric constant of GaAs keeps constant because it changes litter with the changing of carrier concentration, the imaginary part of dielectric constant of GaAs are adopted as 0.0, 0.02, 0.05, 0.10, 0.20, and 0.50, respectively. It also could be found from Fig. 2(b) that the real part of effective indices change litter, while the propagation lengths decrease noticeably with the increasing of imaginary part of dielectric materials filling in the slit. The reason is as follows. As the imaginary part of permittivity of dielectric materials filling in the slit increases, the imaginary part of effective index of GSPPs mode increases, leading to the propagation length increasing.

 figure: Fig. 2

Fig. 2 The effective indices and propagation lengths of GSPPs mode versus slit width for different dielectric materials, the radiation frequency is 1.0 THz; the carrier concentration of InSb is 8.0×1016 cm−3. (a) The dielectric materials filling in the slit are air, polyethylene, water, and GaAs, respectively. (b) The dielectric materials filling in the slit is GaAs, the real part of dielectric constant keeps 3.4, the imaginary part are 0.0, 0.02, 0.05, 0.10, 0.20 and 0.50, respectively.

Download Full Size | PDF

The dispersive properties of GSPPs mode are closely related to the dielectric constant of InSb, which changes with the carrier concentration. Figure 3 demonstrates that the effective indices and propagation lengths of GSPPs mode versus slit width for different carrier concentrations based on heavily doped InSb slit. The insets in Figs. 3(a) and 3(b) are the effective indices and propagation lengths contour for carrier concentration and slit width; the radiation frequency is 1.0 THz; air is filled in the slit. As carrier concentration decreases, the effective indices increase and the propagation lengths decrease. This phenomenon is related to skin depth, which describes the region where THz waves penetrate into InSb, resulting in the surface electromagnetic properties changing. It could be found from Eq. (20) that the skin depth depends on the dielectric properties of InSb, which changes with the carrier concentration of InSb. For instance, when the carrier concentration of InSb slit are 2.0 × 101⁶cm−3 and 8.0 × 101⁶cm−3 (the slit width is 100.0 μm), the dielectric constant are −254.64 + 70.30i and −1065.82 + 281.21i, with corresponding skin depth are 27.32 μm and 9.18 μm, respectively. The larger skin depth at lower concentration means that there are more THz waves penetration into InSb slit, leading to the effective indices decreasing and the propagation constants increasing.

 figure: Figures 3(a) and 3(b)

Figures 3(a) and 3(b) show the effective indices and propagation lengths of GSPPs mode versus slit width, respectively. The radiation frequency is 1.0 THz, the carrier concentration are 2.0×1016 cm−3, 4.0×1016 cm−3, 6.0×1016 cm−3, and 8.0×1016 cm−3, respectively, air is filled in the slit. The insets show the effective indices and propagation lengths contour for slit width and carrier concentration.

Download Full Size | PDF

Water is the major ingredient in biological materials and shows a large imaginary part of dielectric constant in the THz region. THz waves can be used to diagnose cancer or tumors by measuring water content in tumors or cancer, which contain a larger amount of water. Because the changes of propagation constant caused by THz waves with analyte is proportional to the refractive index change [30], subwavelength dimensions biological analyte can be investigated by using SPPs biosensors based on MDM structure. It could also be found from Eq. (17) that the dielectric constant of water is closely relative to temperature and frequency, which is different from the case in the visible and near infrared spectral region [13]. The effective indices and propagation length for the case of water filling in the slit at different frequencies have been shown in Fig. 4(a) ; the carrier concentration of InSb is 8.0 × 101⁶cm−3; the water temperature is 292.3 K. As frequency increases, the effective indices and the propagation lengths decrease. The reason may come from the fact that the permittivity of water decreases with the increasing of frequency. The dielectric constant of water are 15.56 + 8.84i, 7.57 + 6.16i, 6.12 + 4.14i, and 5.39 + 2.39i with the corresponding frequency of 0.1 THz, 0.3 THz, 0.5 THz, and 1.0 THz. As shown above, the effective indices are mainly depended on the real part of dielectric materials filling in the slit. Therefore, the larger dielectric constant of water at lower frequency leads to larger effective index. The effects of temperature on the dispersive properties have been shown in Fig. 4(b), which manifests that the propagation length decreases with the increasing of temperature; the radiation frequency is 1.0 THz; the temperature are 278.8 K, 292.3 K, 315.0 K, and 366.7 K, respectively; their dielectric constant are 5.40 + 1.86i, 5.39 + 2.39i, 5.30 + 3.15i, and 6.00 + 4.54i, respectively. As temperature increases, the real part of dielectric constant of water increases slowly, while the imaginary part of water increases seriously. This case is similar to the results given in Ref. 30, which displays that the solute (sucrose, alcohol) concentration has larger effect on the imaginary part of refractive index than that of real part. The possible reason maybe that as temperature increases, the motion of water molecular and the friction between them increases, the imaginary part of permittivity of water increasing, leading to the propagation length dropping. Therefore, temperature has more effects on the propagation length than the effective indices.

 figure: Fig. 4

Fig. 4 The GSPPs mode of effective indices and propagation lengths versus slit width, the dielectric material filling in the slit is water with different radiation frequencies and temperatures, the doping concentration of InSb is 8.0×1016 cm−3. (a) The water temperature is 292.3 K, the radiation frequencies are 0.1 THz, 0.3 THz, 0.5 THz, and 1.0 THz, respectively. (b) The water temperature are 278.8 K, 292.3 K, 315.0 K, and 366.7 K, respectively.

Download Full Size | PDF

Figure 5(a) shows the field distribution of the electric component along x direction, the length along x axis is normalized by w, the slit materials is heavily doped InSb. The carrier concentration is 8.0×1016 cm−3; the slit width is 20 μm; the radiation frequency is 1.0 THz; the dielectric materials filling in the slit are air, polyethylene, water, and GaAs, respectively. It can be found that the GSPPs mode shows the maximum at the metal-dielectric interface for different slit widths. Furthermore, as the refractive indices of dielectric materials filling in the slit increase, the modes decrease more quickly, which means that mode can be better confined with the increasing refractive index of the core dielectric materials. The penetration depth can be defined by the distance where the absolute value of Ex field decreased by a factor e with respect to the value at the interface between InSb and dielectric core. As the permittivity of core dielectric materials increases, the propagation constant increases, the penetration depth decreases, leading to the fact that THz waves penetrate InSb more quickly. For example, the penetration depth are 1.44 μm, 0.96 μm, 0.60 μm, and 0.44 μm for the case of air, polyethylene, water and GaAs filling in the slit, respectively. The normalized mode distribution at different slit width have been shown in Fig. 5(b), the slit width are 1.0 μm, 10.0 μm, 20.0 μm, 50.0 μm, 100.0 μm, and 200.0 μm, air is filled in the slit. Additionally, Fig. 5(b) displays that the mode can be well confined in the wide slit. The reason may be as follows. As slit width decreases, there are more THz waves penetrating into InSb, leading to the effective indices increasing, shown in Fig. 3. This phenomenon can be well explained by the ratio of penetration depth to slit width, which are 1.447, 0.145, 0.072, 0.029, 0.014, and 0.007 with corresponding for slit width of 1.0 μm, 10.0 μm, 20.0 μm, 50.0 μm, 100.0 μm, and 200.0 μm, respectively.

 figure: Fig. 5

Fig. 5 The field distribution of electric component of GSPPs mode along the x-direction. The slit material is heavily doped InSb with carrier concentration of 8.0×1016 cm−3, the radiation frequency is 1.0 THz. (a) The dielectric materials filling in the slit are air, polyethylene, water, and GaAs, respectively, the slit width is 20 μm. (b) The slit width are 1.0 μm, 10.0 μm, 20.0 μm, 50.0 μm, 100.0 μm, and 200.0 μm, respectively, air is filled in the slit.

Download Full Size | PDF

 figure: Fig. 6

Fig. 6 The electric and magnetic energy densities in the metal and slit versus frequency for different slit width are shown in Fig. 6(a)-6(d), respectively. Air is filled in the slit; the carrier concentration of InSb is 8.0×1016 cm−3; the slit width are 1 μm, 2 μm, 10 μm, 20 μm, 50 μm, and 100 μm, respectively.

Download Full Size | PDF

The ratio of electric energy, magnetic energy stored in heavily doped InSb and slit versus frequency for different slit width are shown in Figs. 6 (a)-6(d), respectively. The carrier concentration of InSb is 8.0 × 1016 cm−3; air is filled in the slit; the slit width are 1.0 μm, 2.0 μm, 10.0 μm, 20.0 μm, 50.0 μm, and 100.0 μm, respectively. As frequency increases, the electric energy in heavily doped InSb increases (Uem), the electric energy in slit (Uei), the magnetic energy in the slit (Umi) and heavily doped InSb (Umm) decreases. It can be found that most of the energy stored in the slit is in the form of electric energy, which is similar to the results in the visible spectral region shown in Ref. 38. But the ratio of Uei in the THz region is larger than that in the visible spectral region, which may result from the former has larger dielectric constant. Furthermore, Uem firstly increases and then decreases with the decreasing of slit width, while the energy stored in the slit vice versa. This agrees with the experimental results in Ref. 23 that the percent of field energy in the metal increases with the decreasing the thickness of dielectric layer. The possible reason maybe as follows. As slit width decreases, more mode have been squeezed from the slit into heavily doped InSb, the interaction region where THz waves with InSb become larger, the effective indices increase and the propagation lengths decrease, resulting in the fact that the fractional electric field energy stored in conductor increasing. At certain value of slit width (which may be related to skin depth), the electric field of the GSPPs mode quickly approaches that of the electrostatic (capacitor) mode. Therefore, the electric energy stored in the slit (i.e. the magnetic energy stored in InSb reducing) increases with the decreasing of slit width. It should also be noted that the ratio of Uei is always larger than Uem in the THz region, while in the visible spectral region, the ratio of Uem increases and can even exceeds than that of Uei with the decreasing of slit width. In addition, MDM structure is very similar to the structure of THz quantum cascade lasers (QCLs), which is one of the most important kinds of semiconductor radiation sources. The results and conclusion can also be used to explain the relevant phenomena of QCLs.

4. Conclusions

The waveguide properties of THz waves through heavily doped InSb slit have been investigated by using the TMM method. The effects of geometrical parameters, carrier concentration and dielectric properties of core materials on the dispersion properties of GSPPs mode have been given and discussed. The results show that as the permittivity of dielectric material filling in the slit increases, the effective indices of GSPPs mode increase, the propagation lengths decrease. The effective indices and propagation lengths of GSPPs mode are closely related to the real and imaginary part of the core dielectric materials, respectively. The contour for carrier concentration and slit width shows that as slit width and carrier concentration decreases, the effective indices increase and the propagation lengths decrease. For the case of water filling in the slit, the effective indices decrease and propagation lengths increase with the decreasing of frequency, and the imaginary part of propagation constant is significantly affected by the temperature than the real part. As the slit width and the refractive index of dielectric materials filling in the slit increases, the mode field can be better confined in the slit. The ratio of electric energy stored in the slit is larger than that for the case of visible spectral, which may result from heavily doped InSb showing larger dielectric constant. As the slit width decreases, the electric energy in the slit Uei firstly decreases and then increases. It is expected the semiconductor MDM structure is very useful for the practical applications of THz waves, such as in the fields of semiconductor biosensors, medical diagnosis, and security detection.

Acknowledgments

This work is supported by the Doctoral Funding of Henan University of Technology (2007BS044).

References and links

1. B. Ferguson and X. C. Zhang, “Materials for terahertz science and technology,” Nat. Mater. 1(1), 26–33 (2002).

2. M. Lee and M. C. Wanke, “Design of n-type silicon-based quantum cascade lasers for terahertz light emission,” Science 316, 64–65 (2007). [PubMed]  

3. T. H. Isaac, W. L. Barnes, and E. Hendry, “Determining the terahertz optical properties of subwavelength films using semiconductor surface plasmons,” Appl. Phys. Lett. 93(24), 241115 (2008).

4. R. Köhler, A. Tredicucci, F. Beltram, H. E. Beere, E. H. Linfield, A. G. Davies, D. A. Ritchi, R. C. Iotti, and F. Rossi, “Terahertz semiconductor-heterostructure laser,” Nature 417, 154–157 (2002).

5. J. C. Cao, A. Z. Li, X. L. Lei, and S. L. Feng, “Current self-oscillation and driving-frequency dependence of negative-effective-mass diodes,” Appl. Phys. Lett. 79(21), 3524–3526 (2001).

6. J. T. Lü and J. C. Cao, “Coulomb scattering in the Monte Carlo simulation of terahertz quantum-cascade lasers,” Appl. Phys. Lett. 89(21), 211115 (2006).

7. H. Li, J. C. Cao, J. T. Lü, and Y. J. Han, “Monte Carlo simulation of extraction barrier width effects on terahertz quantum cascade lasers,” Appl. Phys. Lett. 92(22), 221105 (2008).

8. J. C. Cao, “Interband impact ionization and nonlinear absorption of terahertz radiation in semiconductor heterostructures,” Phys. Rev. Lett. 91(23), 237401 (2003). [PubMed]  

9. K. L. Wang and D. M. Mittleman, “Metal wires for terahertz wave guiding,” Nature 432(7015), 376–379 (2004). [PubMed]  

10. J. A. Deibel, K. Wang, M. D. Escarra, and D. Mittleman, “Enhanced coupling of terahertz radiation to cylindrical wire waveguides,” Opt. Express 14(1), 279–290 (2006). [PubMed]  

11. X. Y. He, “Investigation of terahertz Sommerfeld propagation along conical metal wire,” J. Opt. Soc. Am. B 26(9), A23–A28 (2009).

12. J. Q. Zhang and D. Grischkowsky, “Adiabatic compression of parallel-plate metal waveguides for sensitivity enhancement of waveguide THz time-domain spectroscopy,” Appl. Phys. Lett. 86(6), 061109 (2005).

13. X. Y. He, “Comparison of the waveguide properties of gap surface plasmon in the terahertz region and visible spectra,” J. Opt. A, Pure Appl. Opt. 11(4), 045708 (2009).

14. R. Mendis and M. Daniel, “Mittleman, “An investigation of the lowest-order transverse-electric (TE1) mode of the parallel-plate waveguide for THz pulse propagation,” J. Opt. Soc. Am. B 26(9), A6–A13 (2009).

15. W. L. Barnes, A. Dereux, and T. W. Ebbesen, “Surface plasmon subwavelength optics,” Nature 424(6950), 824–830 (2003). [PubMed]  

16. S. I. Bozhevolnyi and J. Jung, “Scaling for gap plasmon based waveguides,” Opt. Express 16(4), 2676–2684 (2008). [PubMed]  

17. H. Raether, “Surface plasmons on smooth and rough surfaces and on gratings,” (Springer, Berlin, 1988).

18. P. Neutens, P. V. Dorpe, I. D. Vlaminck, L. Lagae, and G. Borghs, “Electrical detection of confined gap plasmons in metal-insulator-metal waveguides,” Nat. Photonics 3(5), 283–286 (2009).

19. J. Lindberg, K. Lindfors, T. Setala, M. Kaivola, and A. T. Friberg, “Spectral analysis of resonant transmission of light through a single sub-wavelength slit,” Opt. Express 12(4), 623–632 (2004). [PubMed]  

20. J. A. Dionne, H. J. Lezec, and H. A. Atwater, “Highly confined photon transport in subwavelength metallic slot waveguides,” Nano Lett. 6(9), 1928–1932 (2006). [PubMed]  

21. T. H. Isaac, J. Gomez., J. R. Rivas, W. L. Sambles, Barnes, and E. Hendry, “Surface plasmon mediated transmission of subwavelength slits at THz frequencies,” Phys. Rev. B 77(11), 113411 (2008).

22. Y. Todorov, A. M. Andrews, I. Sagnes, R. Colombelli, P. Klang, G. Strasser, and C. Sirtori, “Strong light-matter coupling in subwavelength metal-dielectric microcavities at terahertz frequencies,” Phys. Rev. Lett. 102(18), 186402 (2009). [PubMed]  

23. R. M. Gelfand, L. Bruderer, and H. Mohseni, “Nanocavity plasmonic device for ultrabroadband single molecule sensing,” Opt. Lett. 34(7), 1087–1089 (2009). [PubMed]  

24. M. A. Seo, H. R. Park, S. M. Koo, D. J. Park, J. H. Kang, O. K. Suwal, S. S. Choi, P. C. M. Planken, G. S. Park, N. K. Park, Q. H. Park, and D. S. Kim, “Terahertz field enhancement by a metallic nano slit operating beyond the skin-depth limit,” Nat. Photonics 3(3), 152–156 (2009).

25. K. C. Vernon, D. K. Gramontnev, and D. F. P. Pile, “Channel plasmon-polariton modes in V grooves filled with dielectric,” J. Appl. Phys. 103(3), 034304 (2008).

26. M. J. Levene, J. Korlach, S. W. Turner, M. Foquet, H. G. Craighead, and W. W. Webb, “Zero-mode waveguides for single-molecule analysis at high concentrations,” Science 299(5607), 682–686 (2003). [PubMed]  

27. A. H. J. Yang, S. D. Moore, B. S. Schmidt, M. Klug, M. Lipson, and D. Erickson, “Optical manipulation of nanoparticles and biomolecules in sub-wavelength slot waveguides,” Nature 457(7225), 71–75 (2009). [PubMed]  

28. P. U. Jepsen, U. Møller, and H. Merbold, “Investigation of aqueous alcohol and sugar solutions with reflection terahertz time-domain spectroscopy,” Opt. Express 15(22), 14717–14737 (2007). [PubMed]  

29. J. Q. Zhang and D. Grischkowsky, “Waveguide terahertz time-domain spectroscopy of nanometer water layers,” Opt. Lett. 29(14), 1617–1619 (2004). [PubMed]  

30. Y. B. Chen, “Development of mid-infrared surface plasmon resonance-based sensors with highly-doped silicon for biomedical and chemical applications,” Opt. Express 17(5), 3130–3140 (2009). [PubMed]  

31. S. A. Maier, “Gain-assisted propagation of electromagnetic energy in subwavelength surface plasmon polariton gap waveguides,” Opt. Commun. 258(2), 295–299 (2006).

32. M. P. Nezhad, K. Tetz, and Y. Fainman, “Gain assisted propagation of surface plasmon polaritons on planar metallic waveguides,” Opt. Express 12(17), 4072–4079 (2004). [PubMed]  

33. R. Mendis, “Nature of subpicosecond terahertz pulse propagation in practical dielectric-filled parallel-plate waveguides,” Opt. Lett. 31(17), 2643–2645 (2006). [PubMed]  

34. R. Mendis, “THz transmission characteristics of dielectric-filled parallel-plate waveguides,” J. Appl. Phys. 101(8), 083115 (2007).

35. S. W. Gao, J. C. Cao, and S. L. Feng, “Waveguide design of long wavelength semiconductor laser based on surface plasmons,” Physica B 337(1-4), 230–236 (2003).

36. J. T. Lü and J. C. Cao, “Confined optical phonon modes and electron-phonon interactions in wurtzite GaN/ZnO quantum wells,” Phys. Rev. B 71(15), 155304 (2005).

37. F. Yang, J. R. Sambles, and G. W. Bradberry, “Long-range surface modes supported by thin films,” Phys. Rev. B 44(11), 5855–5872 (1991).

38. Y. Kurokawa and H. T. Miyazaki, “Metal-insulator-metal plasmon nanocatities: Analysis of optical properties,” Phys. Rev. B 75(3), 035411 (2007).

39. J. A. Sánchez-Gil and J. G. Rivas, “Thermal switching of the scattering coefficients of terahertz surface plasmon polaritons impinging on a finite array of subwavelength grooves on semiconductor surfaces,” Phys. Rev. B 73(20), 205410 (2006).

40. C. Rønne, P. O. Åstrand, and S. R. Keiding, “THz spectroscopy of liquid H2O and D2O,” Phys. Rev. Lett. 82(14), 2888–2891 (1999).

41. S. Kohen, B. S. Williams, and Q. Hu, “Electromagnetic modeling of terahertz quantum cascade laser waveguides and resonators,” J. Appl. Phys. 97(5), 053106 (2005).

42. A. K. Azad, Y. Zhao, and W. Zhang, “Transmission properties of terahertz pulses through an ultrathin subwavelength silicon hole array,” Appl. Phys. Lett. 86(14), 141102 (2005).

Cited By

Optica participates in Crossref's Cited-By Linking service. Citing articles from Optica Publishing Group journals and other participating publishers are listed here.

Alert me when this article is cited.


Figures (6)

Fig. 1
Fig. 1 Geometry of an infinitely long slit under illumination of THz waves. The slit width is w, which is filled with gain dielectric material in region 2 and bounded by semi-infinite regions of heavily doped InSb in region 1 and 3.
Fig. 2
Fig. 2 The effective indices and propagation lengths of GSPPs mode versus slit width for different dielectric materials, the radiation frequency is 1.0 THz; the carrier concentration of InSb is 8.0×1016 cm−3. (a) The dielectric materials filling in the slit are air, polyethylene, water, and GaAs, respectively. (b) The dielectric materials filling in the slit is GaAs, the real part of dielectric constant keeps 3.4, the imaginary part are 0.0, 0.02, 0.05, 0.10, 0.20 and 0.50, respectively.
Figures 3(a) and 3(b)
Figures 3(a) and 3(b) show the effective indices and propagation lengths of GSPPs mode versus slit width, respectively. The radiation frequency is 1.0 THz, the carrier concentration are 2.0×1016 cm−3, 4.0×1016 cm−3, 6.0×1016 cm−3, and 8.0×1016 cm−3, respectively, air is filled in the slit. The insets show the effective indices and propagation lengths contour for slit width and carrier concentration.
Fig. 4
Fig. 4 The GSPPs mode of effective indices and propagation lengths versus slit width, the dielectric material filling in the slit is water with different radiation frequencies and temperatures, the doping concentration of InSb is 8.0×1016 cm−3. (a) The water temperature is 292.3 K, the radiation frequencies are 0.1 THz, 0.3 THz, 0.5 THz, and 1.0 THz, respectively. (b) The water temperature are 278.8 K, 292.3 K, 315.0 K, and 366.7 K, respectively.
Fig. 5
Fig. 5 The field distribution of electric component of GSPPs mode along the x-direction. The slit material is heavily doped InSb with carrier concentration of 8.0×1016 cm−3, the radiation frequency is 1.0 THz. (a) The dielectric materials filling in the slit are air, polyethylene, water, and GaAs, respectively, the slit width is 20 μm . (b) The slit width are 1.0 μm , 10.0 μm , 20.0 μm , 50.0 μm , 100.0 μm , and 200.0 μm , respectively, air is filled in the slit.
Fig. 6
Fig. 6 The electric and magnetic energy densities in the metal and slit versus frequency for different slit width are shown in Fig. 6(a)-6(d), respectively. Air is filled in the slit; the carrier concentration of InSb is 8.0×1016 cm−3; the slit width are 1 μm , 2 μm , 10 μm , 20 μm , 50 μm , and 100 μm , respectively.

Equations (22)

Equations on this page are rendered with MathJax. Learn more.

γ2γ3γNε1ε2εN1[HN+HN]N,N1=(M11M12M21M22)[H1+H1]2,1,
(M11M12M21M22)=MN1MN2M1,
Mi=(εi1εiεi1εi)(γiγiγi1γi1)(exp(γidi)00exp(γidi)),
γj2=β2k02εi,i=1,2,3,
Exm(r,ω)=exp(α1,3|x|)×exp(i(βzωt)),
Hym(r,ω)=ωε0εmβexp(α1,3|x|)×exp(i(βzωt)),
Ezm(r,ω)=iα1,3β|x|xexp(α1,3|x|)×exp(i(βzωt)),
(|x|<w/2)
Exd(r,ω)=2cosh(α2|x|)×exp(i(βzωt)),
Hyd(r,ω)=2ωε0ε2βcosh(α2|x|)×exp(i(βzωt)),
Ezd(r,ω)=2iα2βsinh(α2|x|)×exp(i(βzωt)),
uei(x,ω)=14ε0εi|εm(ω)εiexp(kmT/2)cosh(kiT/2)|2|A|2(|cosh(kix)|2+|kiksinh(kix)2|),
umi(x,ω)=14ε0|ωεm(ω)ckexp(kmT/2)cosh(kiT/2)|2|A|2(|cosh(kix)|2),
uem(x,ω)=14ε0Re((ωεm(ω))ω)|Aexp(km|x|)|2(1.0+|kmk|2),
umm(x,ω)=14ε0|Aωεm(ω)ckexp(km|x|)|2,
ε(ω)=(εωp2ω2+ωτ2)+iωτωp2ω(ω2+ωτ2),
ε(ω)=ε+εsε11+iωτD+ε1ε1+iωτ2,
εs=87.91e0.00458T,
ε(ω)=ε+ine2τωm(1iωτ),
δsemic0ω(Re(εsemi)+1Re(εsemi)2)1/2.
neff=Re(β)/k0,
L=[2Im(β)]1.
Select as filters


Select Topics Cancel
© Copyright 2024 | Optica Publishing Group. All rights reserved, including rights for text and data mining and training of artificial technologies or similar technologies.