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

Design and analysis of perfect terahertz metamaterial absorber by a novel dynamic circuit model

Open Access Open Access

Abstract

Metamaterial terahertz absorbers composed of a frequency selective layer followed by a spacer and a metallic backplane have recently attracted great attention as a device to detect terahertz radiation. In this work, we present a quasistatic dynamic circuit model that can decently describe operational principle of metamaterial terahertz absorbers based on interference theory of reflected waves. The model comprises two series LC resonance components, one for resonance in frequency selective surface (FSS) and another for resonance inside the spacer. Absorption frequency is dominantly determined by the LC of FSS while the spacer LC changes slightly the magnitude and frequency of absorption. This model fits perfectly for both simulated and experimental data. By using this model, we study our designed absorber and we analyze the effect of changing in spacer thickness and metal conductivity on absorption spectrum.

©2013 Optical Society of America

1. Introduction

Terahertz (THz) radiation encompasses a range of frequencies in the electromagnetic spectrum between 0.1 THz and 10 THz. There has been an increasingly strong driving force to make use of this portion of the spectrum because of the potential benefit of THz technology in a range of application including, but not limited to, medical imaging [1], environmental monitoring of earth [2], remote sensing of explosives [3], and semiconductor electrical property determination [4,5], as terahertz radiation has the unique ability to safely penetrate a wide variety of non-conducting materials including clothing, paper, cardboard, wood, masonry, plastic and ceramics, and to interact with molecules without any ionizing effect [6]. However, this scientifically rich spectrum has been technologically underdeveloped. On the other hand, commonly used devices and techniques in the microwave and optical regime are not applicable to measure and manipulate terahertz radiation. THz detection techniques and devices like Terahertz Time Domain Spectroscopy (THz-TDS) and bolometer suffer from some major drawbacks like larger size, lower sensitivity at room temperature, higher cost or indirect measurement of THz radiation in frequency domain. Therefore, there exist great demands for a compact, low cost and sensitive THz detector capable of directly measuring THz radiation.

Metamaterial which is an artificial subwavelength composite material that can manipulate electromagnetic waves by proper designing its effective permittivity and permeability could be an alternative approach to meet this need [7,8]. Among different devices based on metamaterials, metamaterial terahertz absorbers (MMTA) have been recently attracted attentions due to their potential which promise a novel compact and perfect THz detector. The first MMTA proposed by Hu Tao et al. [9], was based on three major components including a patterned metallic layer as a frequency selective surface (FSS), a dielectric spacer and a metallic backplane, which absorbs THz radiation in a very narrow bandwidth. Soon after, more research was done to enhance the bandwidth of the absorber. Similar structures were reported either to make multiband MMTAs by using different FSS for various resonance frequencies [1012],or to make broadband absorbers by bringing resonance frequencies closer to one another and by engineering the frequency dispersion of FSS to mimic an ideal absorbing sheet in infrared range [1316]. Other structures were also designed and fabricated to improve polarization insensitivity of MMTA by using four-fold symmetric FSS [11, 17].A thorough study of the recent advances in MMTAs has been recently reported by C. M. Watts et al [18].

There have been reports of a few different physical mechanisms to understand operational principles of the metamaterial terahertz absorbers; anti-parallel currents [9], out of phase currents inside the absorber [19], standing waves resonances inside spacer [10],and destructive interference between reflected waves of the FSS and metallic back layer [20]. Several articles have reported modeling of split ring resonators (SRR) based on RLC resonator circuits by using either Transmission Line method (TL) or qausistatic approach [2125]. Lorentz oscillator model has been considered in other works for determining effective permittivity and permeability of metamaterials by quasistatic and nonquasistatic circuit models [16, 2229]. Inspired by SRR electric models, several researchers have proposed similar models applying for FSS and filters in microwave and to understand their principal function and improve their responses [3033]. In those models, metallic patterns have been considered as an inductor in series configuration with a resistor and the dielectric gap between metallic surfaces as a capacitor. Similarly, Y. Pang et al. have proposed a simple TL model to accurately describe the operation of their GHz absorber [34]. This model is composed of a series RLC circuit, to account for resonance and loss in the FSS, which is in parallel to a transmission line accounting for the dielectric spacer. A similar work has also been reported in more detail in THz domain [35]. Furthermore, Q. Wen et al modeled the first reported MMTA by considering both dipole and LC resonances in the FSS as well as the coupling interaction between the two [36]. However to the best of our knowledge, a quasistatic RLC circuit model has not been introduced yet for this type of absorbers to explain its functionality by considering the effect of both FSS and the spacer.

In this work we present a simple and straightforward quasistatic dynamic RLC model that can fit well with absorption response of MMTAs. Based on Lorentzian shape of the absorption spectrum, a simple RLC band-pass circuit is first employed to model the perfect absorption (100%) from the absorber. Then, in order to understand the effect of dielectric spacer on absorber function, a parallel RLC loop is inserted into circuit. In addition to simplicity, our model describes well the physics governing the operation of MMTAs based on destructive interference theory [20].

2. RLC circuit model of MMTA

The circuit model of MMTA was originated from our observation that absorption spectrum of MMTA is a Lorentzian function that can be described as,

f(x)=I(γ2[(xx0)2+γ2])
where x0 is the resonance frequency, γ is half width at half maximum bandwidth (HWHM) and I is the peak of resonance. On the other hand, it is clear from circuit theory that square magnitude of transfer function in a band-pass filter can accurately mimic a Lorentzian function. Therefore a simple band-pass RLC circuit shown in Fig. 1(a) can be used to model an absorption spectrum of MMTA. To design this model, we start with the case of a perfect MMTA which has 100% absorption. As shown in the Fig. 1(a), LC determines the resonance frequency in FSS of MMTA and R determines the resistivity of metal in FSS and backplane. We derived magnitude square of transfer function of the circuit, where
|VoVi|2=(RCω)2(1LCω2)2+(RCω)2
and fitted this to the simulated data from our absorber to find L, C and R. Bandwidth of the band-pass filter is proportional to R, where increasing R means increase in bandwidth which agrees well with the results previously reported [36]. Figures 1(b) and 1(c) illustrate an MMTA structure studied in this paper, which was designed by our group to exhibit polarization independent and wide incident angle response. Details of the simulation and experimental results of polarization insensitivity and incident angle dependence have been reported elsewhere [3739]. In this structure, copper and polyimide have been used as the metal and the spacer material respectively. The simulation has been done through COMSOL Multiphysics by using Finite Element Method (FEM) method. In the origin of designing our perfect absorber structures, we have used Comsol’s standard material library for the material parameters, including a copper and polyimide conductivity of 6 × 107 (S/m) and 6.67 × 10−16 (S/m), respectively. The polyimide complex permittivity of 3.15 + j0.1 was used in our simulations, which was empirically determined by both ellipsometry and THz time domain spectrometer. For the finite element simulation, the total number of free tetrahedral mesh elements is 786971.We have utilized a continuous plane-wave source and periodic boundary conditions to simulate the structure. Due to the metal backplane, there is no transmission through the structure and hence the absorption is calculated by A = 1 – R, where R is the reflection. In this work, absorbers were studied under normal incidence and with a vertical polarization excitation.

 figure: Fig. 1

Fig. 1 (a) Quasistatic RLC model for a perfect MMTA with 100% absorption. (b) Schematic of the FSS of absorber which gray band is copper deposited on polyimide (c) three dimensional picture of the absorber. The space between patterned plane and metal backplane is polyimide.

Download Full Size | PDF

If the MMTA absorption is not perfect, such a simple circuit model will not suffice, as it will always give 100% absorption. To make the circuit model realistic and accommodate the possibility for non-perfect absorption, we should add a component to account for the spacer layer to the model. In order to decrease the peak of absorption for non-perfect cases, another resistor between LC and R should be considered. The peak of absorption changes based on the value of this resistor, and perfect absorption will be obtained when it is equal to zero. In actual MMTA, spacer thickness and permittivity of dielectric tend to change the absorption frequency slightly, while a resistance of the spacer does not change the resonance frequency of the model. Therefore, variation in resistance of the spacer cannot be sufficient to account for the spacer. In order to realize the spacer in the model another series LC component which is parallel to the resistor should be added. Figure 2(a) depicts the complete circuit model that is proposed in this work. Rp and LpCp account for the spacer so that LpCp can bypass the resistor Rp in the perfect case compared to the non-perfect cases, where LpCp can vary both absorption amplitude and frequency slightly. The magnitude square of the transfer function for this circuit shown in Fig. 2 can be written as,

 figure: Fig. 2

Fig. 2 (a) Complete quasistatic RLC model for a MMTA with current loops corresponding to first reflected beam (I1) and collective reflection from spacer cavity (I2). (b) Schematic of reflection beams in a MMTA. I1 is the first reflected beam and I2 is collective reflection spacer cavity. In perfect absorption case, I1 and I2 are out of phase and equal-magnitude.

Download Full Size | PDF

|VoVi|=R2[(RpCpCω2)2+(CωLpCpCω3)2][1+LCLpCpω4(LC+LpCp+RRpCCp)ω2]2+......[(RpC+RC+RpCp)ω(RpCpLC+RpLpCpC+RLpCpC)ω3]2

In this model, the absorption frequency is dominantly determined by FSS as a given LC. LpCp and Rp play the role of resonance inside spacer, between metal backplane and FSS which its resonance frequency is determined by LpCp. In more details, Rp becomes parallel with impedance ZLpCp and represents any loss in the absorber energy due to reflection. Whenever LC and LpCp resonance frequencies match, Rp will get short-circuited and the structure will operate as a perfect absorber with 100% absorption. In this case, the currents I1 and I2, have the same magnitude but are out of phase so that they will cancel out each other in Rp. If we view I2 as the reflected wave of the resonance inside spacer toward the source and I1 as the first reflected beam on FSS interface and air, this model works well to explain interference theory as shown in Fig. 2(b).

In the case of a non-perfect absorber, the LC and LpCp resonance frequencies cannot be equal anymore, thus impedance Z|| = Rp || (ZLpCp) will lead to a drop in the output voltage VO. It means that the reflected wave I2 will no longer be out of phase with I1 and consequently the total reflected power will not be zero (absorption not 100%). Non-perfect absorption can come from two sources: one is the dielectric spacer layer where the thickness or permittivity varies, and the other is variation in metal conductivity. The former case is explicitly explained by changes in LpCp and Rp directly. Unfortunately, effects of changes in metal conductivity are more subtle.

If the metal conductivity remains close to a perfect conductor, the resonance frequency associated with the FSS will be most likely fixed since it comes from L and C of the FSS layer which are independent of metal conductivity. Therefore the dominant resonance frequency of the absorber will not be directly changed by metal conductivity as we observed through simulation of the FSS resonance frequency by using different metal conductivities. However, instead, it is important to note that any change in metal conductivity will eventually change the resonance condition inside the cavity of spacer. Since the permittivity of a metal is related to its conductivity through εr=ε'+jσ/ω and its refractive index is related to the permittivity through n=εr, the conductivity of a metal changes its refractive index. This change in metal refractive index will then alter the phase and amplitude of the reflected and transmitted waves at both facets inside the spacer cavity resulting in a change in the resonance condition of the cavity and hence LpCp resonance frequency. The LpCp resonance frequency will shift further from LC (the major and dominant resonance frequency associated with the FSS) and the impedance (ZLpCp) of LpCp will no longer be zero. According to the change in cavity condition, ZLpCp will be a positive or negative imaginary number resulting in a phase shift in I2. Therefore reflected waves of I1 and I2 cannot be completely out of phase anymore. In this case, Rp will get involved in the circuit and will decrease absorption and LpCp will shift the structure absorption resonance frequency.

3. Results and discussion

Table 1 summarizes parameters used to simulate two different MMTAs, A and B which are studied. In this table, the parameters are used to simulate MMTAs for perfect absorber and they are indicated in Fig. 1. Both structures are simulated to obtain absorption magnitude and resonance frequencies and we applied our model to fit the absorption spectra.

Tables Icon

Table 1. Dimensions of the simulated and fabricated absorbers

Figure 3 shows the results of fitting Lorentzian Eq. (1) to the simulated data for the absorber A with 7 µm spacer thickness. Polyimide with 7 µm thickness gives nearly perfect absorption (99.8%) at 0.87 THz. We obtained x0 = 0.8713 THz, γ=0.01967 THz, and I = 1 from the fitting results respectively, which confirms our assumption for perfect absorber discussed in previous section. Additional simulations confirmed that reasonable changes in spacer thickness from 5.5 up to 15 µm, spacer permittivity from 3 up to 3.5 or spacer conductivity do not change the resonance frequency more than half width at half maximum bandwidth of the absorber. Thus, we used this limit on LpCp and Rp to avoid undesired values for parameters. If the small change in LpCp and Rp results in large deviation in resonance frequency, those set of values are not acceptable so that fitting will get continued to find acceptable values.

 figure: Fig. 3

Fig. 3 Lorentzian fitting of the simulation data of absorber A with 7 µm polyimide thickness.

Download Full Size | PDF

To examine non perfect case in circuit model, we first started with fitting absorber A (perfect case with 7 µm spacer thickness) with Eq. (3). Results of the obtained fitting parameters are summarized in Table 2. Then we varied the spacer thickness to both 6 µm and 9 µm in the same absorber A structure to achieve non-perfect cases. Since everything is same except the polyimide thickness, we can use the values of L, C and R obtained for the 7 µm case in the Table 2 as a reference. Once L, C and R are fixed, the Lp, Cp and Rp can be determined by fitting Eq. (3) again for absorber A with 6 and 9 µm. ZLpCp is almost zero for 7 µm but it is positive and negative imaginary numbers for 6 and 9 µm (non-perfect case) respectively, confirming that I2 is not out of phase with I1 anymore, hence the absorber loses some energy by non-zero reflection.

Tables Icon

Table 2. Parameters extracted from fitting for simulation data of the absorber A with three different polyimide thicknesses

Figure 4 illustrates fitting results from the absorber A and compares those three absorbers with different polyimide thicknesses. The dashed curve is the fitting result for the 7 µm or perfect case while the solid curves belong to the fitting results for non-perfect absorber with 6 and 9 µm spacer thickness in Figs. 4(a) and 4(b), respectively. Circular dots represent data obtained from FEM simulation for 7µm case and square dots are the data from simulation for non-perfect MMTAs. We also simulated the absorber with finer frequency step sizes around the peak to verify the matching peak position between simulation data and fitting curves. We confirmed resonance frequency and absorption value at peak matched very well for both fitting and finer simulation. From the Fig. 4, we found fres is 0.867 THz with 97.00% absorption for 6 µm MMTA, and fres is 0.875 THz with 97.5% absorption for 9 µm MMTA.

 figure: Fig. 4

Fig. 4 Fitting results for absorber A by using RLC model. Circular dots are simulation data for 7 µm MMTA and square dots are simulation data for 6 and 9 µm in figures a and b respectively. Solid line is the fittings for non-perfect MMTA and dashed line is the fitting for perfect case. Insets are fitting of simulation with finer frequency at peaks by using the electric model.

Download Full Size | PDF

Finally the circuit model was applied to the experimental results. Absorber C in Table 1 describes the dimensions of an absorber structure fabricated using standard photolithography. The Cu backplane and FSS Cu elements were deposited by electron beam evaporation with the FSS elements patterned by UV exposure of a positive photoresist through a dark-field photomask. After metal deposition of the FSS elements, excess Cu was removed by lift-off. The polyimide dielectric layer between the ground plane and FSS layer was deposited by spin coating consecutive layers to achieve the desired thickness. Once applied to the ground plane, the polyimide was baked in nitrogen at 350°C for one hour to ensure complete curing. It should be noted that variation in the polyimide curing process can lead to slight differences in the optical properties of the polyimide film. For this work, the polyimide curing procedure was strictly maintained constant to avoid this issue. THz time domain spectroscopy in reflection mode with normal incident beam was used for the experimental absorption measurement. The details of the measurement were reported in previous work [39].

Figure 5(a) shows the absorption spectrum and fitting results to Eq. (3) for both experimental and simulation data (absorbers B and C). The circular dots show the data from simulation for the perfect absorber case (absorber B) as the reference and the square dots correspond to experimentally measured data for the same absorber with 10.7 µm polyimide thickness (absorber C, supposed to be non-perfect). It is clearly observed that the resonance frequency was shifted to the left and the absorption bandwidth was significantly broadened compared to the resonance frequency of 0.899 THz with 99.99% absorption of the perfect absorber with polyimide thickness of 7.5 µm. The fitting parameters which show large variations in R, Rp and LpCp are summarized in Table 3. Similarly to what was done for the fitting procedure for absorber A, we first fitted absorption spectrum of the perfect case (absorber B) and determined the values of R, L and C. Then we used the same values of R, L and C for absorber C to fit absorption spectrum of absorber C using Eq. (3) finding Rp,Lp and Cp.

 figure: Fig. 5

Fig. 5 (a) Fitting results for absorber B by using RLC model. Circular dots are simulation data for 7.5 µm polyimide thickness and square dots are measured data for 10.7 µm polyimide thickness in the sample. Solid and dashed lines are the fitting results. (b) Effect of the changing in conductivity of copper on FSS reflection. Solid line is for σ=6×107S/mand dashed line is for σ=6×106S/m

Download Full Size | PDF

Tables Icon

Table 3. Parameters extracted from fitting for simulation and experiment data of the absorber B and C.

However, we found the fitting in the case of absorber C was not straightforward and did not converge at first. The reason was because the values of L, C and R were fixed and changing Lp,Cp and Rp meant reasonable deviation in polyimide characteristics and thickness that could not shift the resonance frequency from 0.9 THz to 0.85 THz. In fact, based on our simulations, we showed that increasing the polyimide thickness (from the 7.5 µm of absorber B to the10.7 µm of absorber C) actually leads to a blue shift of the resonance from 0.9 THz to 0.91 THz, rather than the observed red shift. In order to proper fit the experimentally measured data of absorber C, we found that R should be higher than the value used in FEM simulation, and the resulting fit parameters are shown in Table 3. Because R is related to the resistivity of the copper, the obtained higher R means a lower conductivity of copper, which is the main reason for bandwidth broadening and confirms the similar result reported by other researcher [36]. A reduction in the conductivity of copper might be due to the thin copper layer thickness which makes carrier scattering by lattice defects much larger than scattering by phonons [40].To confirm the impact of a lower copper conductivity, and to show that FEM simulation using a lower conductivity leads to a more reasonable match to experiment, we measured the conductivity of the copper in our structure and obtained a value of 3 × 107 (S/m), which is half the one originally used from the Comsol standard material library. FEM simulation using this reduced conductivity in the geometry of absorber C then led to a red shift in the resonance frequency to 0.87 THz (from 0.91 THz). This was accompanied by a broadening of the absorption peak to close to that in Fig. 5(a).Furthermore, as discussed in the previous section, the conductivity of copper should not affect the resonance frequency of the FSS itself but will change the amplitude and the phase of the reflected and transmitted waves of the FSS. To verify this, we simulated the reflection spectrum from the FSS alone without ground backplane for two different copper conductivities as shown in Fig. 5(b). It can be seen that reducing copper conductivity by 10 times does not change FSS resonance frequency at all. From the above reasoning, we conclude that the differences between the absorption spectra of perfect absorber B and experimentally fabricated absorber C are majorly due to the change in copper conductivity, which results in a resonance frequency red-shift, a peak broadening and leads to large changes in LpCp and Rp values in the fitting results.

To better understand the influence of metamaterial geometries on the RLC values, we can gain some insight on the physical origin of L and C by first observing in Fig. 6(a) the electric field strength and current density distribution on the front layer of perfect absorber B. For clarity, we have added white arrows to show the direction of current. It clearly shows that resonance in the FSS arises not only in the gaps between the two vertical legs but also as a dipole on the square ring. By contrast, the horizontal legs do not contribute to the resonance. From this observation, we can infer that the capacitance C originates primarily from the electric charges in upper and lower parts of the FSS pattern where the electric field looks most intense. On the other hand, the origin of inductance L should be straightforward to understand, since we know the path of the resonant current. It comes from the self-inductance of metal bars in which the current flows. According to the current path on the FSS, the simplified equivalent circuit model for the FSS should be like Fig. 6(b) which will result in the equivalent inductance:

 figure: Fig. 6

Fig. 6 (a) Electric field strength and current density distribution on FSS. White arrows indicate to the path of the resonance current. (b) Equivalent circuit model of the FSS with red arrows indicating to the resonance current path. (c) Current density and electric field profile on the backplane with black arrows showing the current path.

Download Full Size | PDF

L=0.5(2L2+L3)+2L1

The self-inductance of a metal wire is proportional to its length, therefore by increasing the length of the wire its inductance will increase. This is why the inductance L of absorber A is larger than B, because, in absorber A, wp is bigger than that of B and the trapezoidal part of vertical legs contributes to the resonance current, therefore the value of L1 and hence L increases in absorber A. Since the capacitance C is almost the same between absorbers A and B, the resulting resonance frequency decreases.

Figure 6(c) shows the current density and electric field profile on the backplane. As seen in the figure, the resonance current on backplane is in opposite direction to that in the FSS. These opposite currents make a mutual inductance between the FSS metal pattern and the backplane which is at the origin of Lp in our model. We can use the method of image charges to understand this mutual inductance such that we can remove metal backplane and place the image of the FSS pattern with respect to the backplane. Whenever the distance between the FSS and its image increases, the mutual induction between them, i.e. Lp, decreases, which is consistent with the fitting results for Lp for absorber A when the spacer thickness was changed.

As for the resistance R in the model, it is the sum of the resistances from the FSS metal bars and the metal backplane, which can be calculated approximately through R = ρl/A on the FSS bars by considering skin depth at terahertz range. Because the current on the backplane is distributed on a large area, the backplane resistance should be much smaller and probably negligible compared to the resistance from the FSS.

4. Conclusion

In summary, we showed that MMTA can be fitted well to Lorentzian function and hence we can model it as a quaistatic RLC circuit model. Our RLC model is based on a band-pass circuit with an additional loop of LpCp || Rp as the polyimide positioned between LC and R as the FSS and loss in metal respectively. This model can explain well the physics of the MMTA, similar to the previously reported interference theory. We analyzed our designed and fabricated absorbers by this model and we investigated intentional changes in fabricated sample based on the electric elements values of the circuit model.

Acknowledgment

We would like to acknowledge NSF BRIGE 0824452 and NSF CAREER for supporting this work.

References and links

1. R. M. Woodward, B. E. Cole, V. P. Wallace, R. J. Pye, D. D. Arnone, E. H. Linfield, and M. Pepper, “Terahertz pulse imaging in reflection geometry of human skin cancer and skin tissue,” Phys. Med. Biol. 47(21), 3853–3863 (2002). [CrossRef]   [PubMed]  

2. . W. Waters, L. Froidevaux, R. S. Harwood, R. F. Jarnot, H. M. Pickett, W. G. Read, P. H. Siegel, R. E. Cofield, M. J. Filipiak, D. A. Flower, J. R. Holden, G. K. Lau, N. J. Livesey, G. L. Manney, H. C. Pumphrey, M. L. Santee, D. L. Wu, D. T. Cuddy, R. R. Lay, M. S. Loo, V. S. Perun, M. J. Schwartz, P. C. Stek, R. P. Thurstans, M. A. Boyles, K. M. Chandra, M. C. Chavez, B. V. Gun-Shing Chen, R. Chudasama, R. A. Dodge, M. A. Fuller, J. H. Girard, Jiang, B. W. Yibo Jiang, R. C. Knosp, J. C. LaBelle, K. A. Lam, D. Lee, J. E. Miller, N. C. Oswald, D. M. Patel, O. Pukala, D. M. Quintero, W. Scaff, M. C. Van Snyder, P. A. Tope, Wagner, and M. J. Walch, “The earth observing system microwave limb sounder (EOS MLS) on the Aura satellite,” IEEE Trans. Geosci. Rem. Sens. 44(5), 1075–1092 (2006). [CrossRef]  

3. R. Appleby, “Standoff detection of weapons and contraband in the 100 GHz to 1 THz region” IEEE Trans. AntennasPropag. 55(11), 2944–2956 (2007).

4. N. Nagai, M. Sumitomo, M. Imaizumi, M. Imaizumi, and R. Fukasawa, “Characterization of electron- or proton-irradiated Si space solar cells by THz spectroscopy,” Semicond. Sci. Technol. 21(2), 201–209 (2006). [CrossRef]  

5. S. Balci, W. Baughman, D. S. Wilbert, G. Shen, P. Kung, and S. M. Kim, “Characteristics of THz carrier dynamics in GaN thin film and ZnO nanowires by temperature dependent terahertz time domain spectroscopy measurement,” Solid-State Electron. 78, 68–74 (2012). [CrossRef]  

6. N. Laman, S. S. Harsha, D. Grischkowsky, and J. S. Melinger, “High-resolution waveguide THz spectroscopy of biological molecules,” Biophys. J. 94(3), 1010–1020 (2008). [CrossRef]   [PubMed]  

7. J. Zhang, K. F. MacDonald, and N. I. Zheludev, “Controlling light-with-light without nonlinearity,” Nature Light: Science and Applications 1, 1–5 (2012).

8. M. Iwanaga, “Photonic metamaterials: a new class of materials for manipulating light waves,” Sci. Technol. Adv. Mater. 13(5), 053002–053019 (2012). [CrossRef]  

9. H. Tao, N. I. Landy, C. M. Bingham, X. Zhang, R. D. Averitt, and W. J. Padilla, “A metamaterial absorber for the terahertz regime: design, fabrication and characterization,” Opt. Express 16(10), 7181–7188 (2008). [CrossRef]   [PubMed]  

10. X. Y. Peng, B. Wang, S. Lai, D. H. Zhang, and J. H. Teng, “Ultrathin multi-band planar metamaterial absorber based on standing wave resonances,” Opt. Express 20(25), 27756–27765 (2012). [CrossRef]   [PubMed]  

11. X. Shen, T. J. Cui, J. Zhao, H. F. Ma, W. X. Jiang, and H. Li, “Polarization-independent wide-angle triple-band metamaterial absorber,” Opt. Express 19(10), 9401–9407 (2011). [CrossRef]   [PubMed]  

12. Q. Ye, Y. Liu, H. Lin, M. Li, and H. Yang, “Multi-band metamaterial absorber made of multi-gap SRRs structure,” Appl. Phys., A Mater. Sci. Process. 107(1), 155–160 (2012). [CrossRef]  

13. L. Huang, D. R. Chowdhury, S. Ramani, M. T. Reiten, S. N. Luo, A. J. Taylor, and H. T. Chen, “Experimental demonstration of terahertz metamaterial absorbers with a broad and flat high absorption band,” Opt. Lett. 37(2), 154–156 (2012). [CrossRef]   [PubMed]  

14. J. Sun, L. Liu, G. Dong, and J. Zhou, “An extremely broad band metamaterial absorber based on destructive interference,” Opt. Express 19(22), 21155–21162 (2011). [CrossRef]   [PubMed]  

15. Y. Liu, S. Gu, C. Luo, and X. Zhao, “Ultra-thin broadband metamaterial absorber,” Appl. Phys., A Mater. Sci. Process. 108(1), 19–24 (2012). [CrossRef]  

16. Q. Feng, M. Pu, C. Hu, and X. Luo, “Engineering the dispersion of metamaterial surface for broadband infrared absorption,” Opt. Lett. 37(11), 2133–2135 (2012). [CrossRef]   [PubMed]  

17. Y. Ma, Q. Chen, J. Grant, S. C. Saha, A. Khalid, and D. R. S. Cumming, “A terahertz polarization insensitive dual band metamaterial absorber,” Opt. Lett. 36(6), 945–947 (2011). [CrossRef]   [PubMed]  

18. C. M. Watts, X. Liu, and W. J. Padilla, “Metamaterial electromagnetic wave absorbers,” Adv. Mater. 24(23), OP98–OP120, OP181 (2012). [CrossRef]   [PubMed]  

19. Y. Zeng, H. T. Chen, and D. A. R. Dalvit, “The role of magnetic dipoles and non-zero-order Bragg waves in metamaterial perfect absorbers,” Opt. Express 21(3), 3540–3546 (2013). [CrossRef]   [PubMed]  

20. H. T. Chen, “Interference theory of metamaterial perfect absorbers,” Opt. Express 20(7), 7165–7172 (2012). [CrossRef]   [PubMed]  

21. J. D. Baena, J. Bonache, F. Martin, R. M. Sillero, F. Falcone, T. Lopetegi, M. A. G. Laso, J. Garcia-Garcia, I. Gil, M. F. Portillo, and M. Sorolla, “Equivalent-circuit models for split-ring resonators and complementary split-ring resonators coupled to planar transmission lines,” IEEE Trans. Microw. Theory Tech. 53(4), 1451–1461 (2005). [CrossRef]  

22. Q. Wu, M. F. Wu, F. Y. Meng, J. Wu, and L. W. Li, “Modeling the effect of n individual SRR by equivalent circuit method” IEEE Antenna and Propagation Society International Symposium, (Washington DC, 2005), 1B. pp. 631–634.

23. Y. Nakata, T. Okada, T. Nakanishi, and M. Kitano, “Circuit model for hybridization modes in metamaterials and its analogy to quantum tight-bonding model,” Phys. Status Solidi B 249(11), 2293–2302 (2012). [CrossRef]  

24. F. Bilotti, A. Toscano, L. Vegni, K. Aydin, K. B. Alici, and E. Ozbay, “Equivalent circuit models for the design of metamaterial based on artificial magnetic inclusions,” IEEE Trans. Microw. Theory Tech. 55(12), 2865–2873 (2007). [CrossRef]  

25. A. K. Azad, A. J. Taylor, E. Smirnova, and J. F. Ohara, “Characterization and analysis of terahertz metamaterial based on rectangular split-ring resonators,” Appl. Phys. Lett. 92(1), 011119 (2008). [CrossRef]  

26. S. Linden, C. Enkrich, G. Dolling, M. W. Klein, J. Zhou, T. Koschny, C. M. Soukoulis, S. Burger, F. Schmidt, and M. Wegener, “Photonic metamaterials: magnetism at optical frequencies,” IEEE J. Sel. Top. Quantum Electron. 12(6), 1097–1105 (2006). [CrossRef]  

27. S. Linden, C. Enkrich, M. Wegener, J. Zhou, T. Koschny, and C. M. Soukoulis, “Magnetic response of metamaterials at 100 terahertz,” Science 306(5700), 1351–1353 (2004). [CrossRef]   [PubMed]  

28. T. P. Meyrath, T. Zentgraf, and H. Giessen, “Lorentz model for metamaterials: optical frequency resonance circuit,” Phys. Rev. B 75(20), 205102 (2007). [CrossRef]  

29. T. D. Karamanos, A. I. Dimitriadis, and N. V. Kantartzis, “Compact double-negative metamaterials based on electric and magnetic resonators,” IEEE Antennas Wirel. Propag. Lett. 11, 480–483 (2012). [CrossRef]  

30. R. J. Longley, “Double-square frequency-selective surfaces and their equivalent circuit,” Electron. Lett. 19(17), 675–677 (1983). [CrossRef]  

31. F. Medina, F. Mesa, and R. Maraques, “Extraordinary transmission through arrays of electrically small holes from a circuit theory perspective ” IEEE Trans. Microw. Theory and Techn . 56(12) 3108–3120 (2008).

32. M. D. Amore, V. D. Santis, and M. Feliziani, “Equivalent circuit modeling of frequency-selective surfaces based on nanostructured transparent thin films,” IEEE Trans. Magn. 48(2), 703–706 (2012). [CrossRef]  

33. X. Zhang, Q. Li, W. Cao, W. Yue, J. Gu, Z. Tian, J. Han, and W. Zhang, “Equivalent circuit analysis of terahertz metamaterial filters,” Chin. Opt. Lett. 9(11), 110012 (2011).

34. Y. Q. Pang, Y. J. Zhao, and J. Wang, “Equivalent circuit model analysis of the influence of frequency selective surfaces on the frequency response of metamaterial absorbers,” J. Appl. Phys. 110(2), 023704 (2011). [CrossRef]  

35. F. Costa, S. Genovesi, A. Monorchio, and G. Manara, “A circuit based model for the interpretation of perfect metamaterial absorbers”IEEE Trans. AntennasPropag. 61(3), 1201–1210 (2013).

36. Q. Y. Wen, Y. S. Xie, H. W. Zhang, Q. H. Yang, Y. X. Li, and Y. L. Liu, “Transmission line model and fields analysis of metamaterial absorber in the terahertz band,” Opt. Express 17(22), 20256–20265 (2009). [CrossRef]   [PubMed]  

37. L. Butler, D. S. Wilbert, W. Baughman, S. Balci, P. Kung, S. M. Kim, M. S. Heimbeck, and H. O. Everitt, “Design, simulation and characterization of THz metamaterial absorber,” Proc. SPIE 8363, 83630J, 83630J-8 (2012). [CrossRef]  

38. L. A. Butler, “Design, simulation, fabrication, and characterizations of terahertz metamaterial devices” http://acumen.lib.ua.edu/content/u0015/0000001/0000899/u0015_0000001_0000899.pdf

39. D. S. Wilbert, M. P. Hokmabadi, J. Martinez, P. Kung, and S. M. Kim, “Terahertz metamaterial perfect absorbers for sensing and imaging,” Proc. SPIE 8585, 85850Y, 85850Y-6 (2013). [CrossRef]  

40. N. Laman and D. Grischkowsky, “Terahertz conductivity of thin metal films,” Appl. Phys. Lett. 93(5), 051105 (2008). [CrossRef]  

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 (a) Quasistatic RLC model for a perfect MMTA with 100% absorption. (b) Schematic of the FSS of absorber which gray band is copper deposited on polyimide (c) three dimensional picture of the absorber. The space between patterned plane and metal backplane is polyimide.
Fig. 2
Fig. 2 (a) Complete quasistatic RLC model for a MMTA with current loops corresponding to first reflected beam (I1) and collective reflection from spacer cavity (I2). (b) Schematic of reflection beams in a MMTA. I1 is the first reflected beam and I2 is collective reflection spacer cavity. In perfect absorption case, I1 and I2 are out of phase and equal-magnitude.
Fig. 3
Fig. 3 Lorentzian fitting of the simulation data of absorber A with 7 µm polyimide thickness.
Fig. 4
Fig. 4 Fitting results for absorber A by using RLC model. Circular dots are simulation data for 7 µm MMTA and square dots are simulation data for 6 and 9 µm in figures a and b respectively. Solid line is the fittings for non-perfect MMTA and dashed line is the fitting for perfect case. Insets are fitting of simulation with finer frequency at peaks by using the electric model.
Fig. 5
Fig. 5 (a) Fitting results for absorber B by using RLC model. Circular dots are simulation data for 7.5 µm polyimide thickness and square dots are measured data for 10.7 µm polyimide thickness in the sample. Solid and dashed lines are the fitting results. (b) Effect of the changing in conductivity of copper on FSS reflection. Solid line is for σ=6× 10 7 S/m and dashed line is for σ=6× 10 6 S/m
Fig. 6
Fig. 6 (a) Electric field strength and current density distribution on FSS. White arrows indicate to the path of the resonance current. (b) Equivalent circuit model of the FSS with red arrows indicating to the resonance current path. (c) Current density and electric field profile on the backplane with black arrows showing the current path.

Tables (3)

Tables Icon

Table 1 Dimensions of the simulated and fabricated absorbers

Tables Icon

Table 2 Parameters extracted from fitting for simulation data of the absorber A with three different polyimide thicknesses

Tables Icon

Table 3 Parameters extracted from fitting for simulation and experiment data of the absorber B and C.

Equations (4)

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

f(x)=I( γ 2 [ (x x 0 ) 2 + γ 2 ] )
| V o V i | 2 = (RCω) 2 (1LC ω 2 ) 2 + (RCω) 2
| V o V i |= R 2 [ ( R p C p C ω 2 ) 2 + ( Cω L p C p C ω 3 ) 2 ] [1+LC L p C p ω 4 (LC+ L p C p +R R p C C p ) ω 2 ] 2 +... ... [( R p C+RC+ R p C p )ω( R p C p LC+ R p L p C p C+R L p C p C) ω 3 ] 2
L=0.5(2 L 2 + L 3 )+2 L 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.