Startseite Naturwissenschaften Broadband graphene-on-silicon modulator with orthogonal hybrid plasmonic waveguides
Artikel Open Access

Broadband graphene-on-silicon modulator with orthogonal hybrid plasmonic waveguides

  • Mingyang Su , Bo Yang , Junmin Liu , Huapeng Ye , Xinxing Zhou , Jiangnan Xiao , Ying Li , Shuqing Chen ORCID logo EMAIL logo und Dianyuan Fan
Veröffentlicht/Copyright: 18. Mai 2020

Abstract

Graphene, a two-dimensional nanomaterial, possess unique photoelectric properties that have potential application in designing optoelectronic devices. The tunable optical absorption is one of the most exciting properties that can be used to improve the performance of silicon modulators. However, the weak light–matter interaction caused by the size mismatch between the optical mode fields and graphene makes the graphene-on-silicon modulator (GOSM) has large footprint and high energy consumption, limiting the enhancement of modulation efficiency. Here, we propose a broadband GOSM with orthogonal hybrid plasmonic waveguides (HPWs) at near-infrared wavelengths. The orthogonal HPWs are designed to compress the interaction region of optical fields and enhance the light-graphene interaction. The results show that the GOSM has a modulation depth of 26.20 dB/μm, a footprint of 0.33 μm2, a 3 dB modulation bandwidth of 462.77 GHz, and energy consumption of 2.82 fJ/bit at 1.55 μm. Even working at a broad wavelength band ranging from 1.3 to 2 μm, the GOSM also has a modulation depth of over 8.58 dB/μm and energy consumption of below 4.97 fJ/bit. It is anticipated that with the excellent modulation performance, this GOSM may have great potential in broadband integrated modulators, on-chip optical communications and interconnects, etc.

1 Introduction

Electro-optic modulator, converting electronic signals into high-bit-rate photonic data, is a crucial device in integrated optoelectronics and has been widely applied in optical interconnect [1], [2], environmental monitoring [3], and optical sensing [4], etc. The signals can be modulated to the phase, amplitude, and polarization of lights by electro-absorption effect [5], magneto-optic effect [6], and thermo-optic effect [7]. Due to the fast response speed and low power consumption, electro-absorption modulators that exchange the electrical and optical signals via adjusting the optical absorption of materials with external voltages has attracted extensive attention [8], [9]. According to the electro-absorption materials, these modulators can be classified into two categories: III–V material based-modulators and IV material based-modulators [10], [11], [12], [13], [14]. III–V based-modulators are the most common modulators and have been extensively used in photoelectric information processing because of the high modulation efficiency, better temperature stability, and low bias. But the large footprint, high cost and fabrication difficulty make it hard in optoelectronic integrated circuits [10], [11]. Recently, IV material based-modulators, such as silicon-based modulators, attracted research interests for the high integration of photoelectric devices. However, the silicon-based modulator always has a low modulation rate and a large footprint because of the weak electro-absorption effect in silicon [13], [14]. Thus, how to design a silicon-based modulator with ultra-compact device footprint, low energy consumption, high modulation speed, and broad working bandwidth is always a challenge in on-chip optical communications and interconnects.

To decrease the footprint and improve the modulation efficiency, many silicon-based modulators with novel structures have been proposed, such as high-Q resonators [15], [16], hybrid semiconductors [17], [18], and surface plasmon polariton structures [19]. The high-Q optical resonator may increase the modulation depth and reduce the footprint, but it suffers from narrow working bandwidth and poor thermal stability [15]. Integrating semiconductors into the silicon-based modulator may partially resolve these issues, but the difficult fabrication process will enlarge the size of the modulator to micrometers [20], [21]. Surface plasmon polariton modulators can improve the modulation efficiency and reduce the footprint that makes the modulator compact, but the large loss and low modulation speed are inevitable because of the additional losses induced by surface plasmon resonance [19]. One of the most enticing promises that delimit further development of silicon-based modulators might come from the lack of suitable optical materials which have high electric-to-optic conversion efficiency.

Graphene, a two-dimensional nanomaterial composed of carbon atoms with hexagonal lattices, has aroused growing interest by virtue of its excellent electrical, optical, thermal, and mechanical properties, which may have great potential in designing graphene-based optoelectronic devices [22], [23], [24], [25], [26]. Because of the zero-band-gap structure, graphene has a wide response range, and its optical absorption ranges from the visible to the infrared band [27], [28]. Electrons in graphene behave as massless particles, and the mobility is up to 200,000 cm2 V−1 s−1 at room temperature, implying that the graphene-based devices may have a large theoretical modulation bandwidth [29], [30]. Furthermore, the complex conductivity of graphene can be easily adjusted by chemical and electrostatic doping, which can be realized by applying a gate bias voltage to the graphene layers [31]. These exceptional electrical and optical properties, excellent thermal conductivity, stability, and carrier mobility, make graphene a priority electro-optic material for electro-absorption modulators [32], [33], [34]. In 2011, Liu et al. proposed the first graphene-on-silicon modulator (GOSM) by placing single-layer graphene on a silicon waveguide, and the 3 dB modulation bandwidth up to 1 GHz and modulation depth of 0.18 dB/μm are achieved [24]. Subsequently, various structures have been proposed to enhance the modulation efficiency of GOSM, such as silicon-based slot waveguides [35], double-layer graphene waveguides [36], and Mach-Zehnder interferometers [37], etc. However, because the optical mode field is much larger than the thickness of graphene, the overlap between graphene and the optical field is quite small, resulting in a weak light-graphene interaction and limiting the further enhancement of modulation efficiency. Researchers have devoted to improving the efficiency of light-graphene interaction, including micro-ring resonances [38], [39], localized plasmon resonances [40], and plasmonic waveguides [41], [42]. For the micro-ring resonance modulators, the interaction only happens in a few micro-meter resonator footprints, but the modulation depth is usually restricted to ∼10 dB [43]. Moreover, it always suffers from working bandwidth limitations and temperature instability. Localized plasmon resonance modulators may reduce the sensitivity to temperature and increase the modulation rate to several hundred gigahertz, but the working bandwidth is limited [44]. Using the field confinement of subwavelength structure, plasmonic waveguide modulators have a high light–matter interaction, but the requirement of a few micro-meters active graphene footprints will cause a high capacitance and energy consumption [45]. Thus, it is still a challenge to achieve broadband GOSM with high modulation depth, small footprint, and low energy consumption.

Hybrid plasmonic waveguide (HPW) with great potential in confining optical mode field at subwavelength scale has been applied in many fields, such as optical absorber [46], optical sensor [47], and optical tweezers [48]. Its sandwich structure composes a metal layer, a high index material layer, and a low index material layer, and the optical mode field is highly confined in the low-index layer due to the high index difference between the layers [49], [50]. Recently, HPW has been used in GOSM to improve modulation efficiency, but the requirement of larger active graphene footprint remains a constraint [51]. In this work, we propose a broadband GOSM with orthogonal HPWs. The HPWs are arranged in the two orthogonal directions on the cross-section of the modulator to confine optical mode fields and lower the requirement of the active area of graphene, which can reduce the capacitance and energy consumption. After optimizing the parameters, this modulator has a modulation depth of 26.20 dB/μm for transverse magnetic and achieves a 3 dB modulation bandwidth of 462.77 GHz with only 0.33 μm2 active region and 2.82 fJ/bit energy consumption at 1.55 μm. Besides, further studies also suggest that this modulator has a wide working wavelength ranging from 1.3 to 2 μm. By virtue of the excellent modulation performance, this GOSM is expected to be applied in optical interconnects, photonic integrated chip, and other fields.

2 Design of the GOSM with orthogonal HPWs

The 3D schematic configuration and the cross-sectional view of the proposed GOSM are displayed in Figures 1(a) and (b), respectively. The modulator is composed of a graphene-hBN-graphene layer, two high index silicon (Si) layers, and two metallic silver (Ag) layers, which are inserted into a silica (SiO2) waveguide acting as the protective and substrate layer. The two Si and two Ag layers with the same width w and thickness t1 are interleaved in the modulator and separated by a low index SiO2 layer (the thickness is t2) to construct orthogonal HPWs. This orthogonal structure can compress the light field twice on the cross section of modulator to achieve the strong confinement of optical mode fields due to the high index difference between the layers. The two graphene layers are separated by hBN and placed in the middle of the modulator, serving as the active medium for light absorption. The gold is used as the electrode to introduce external electric fields. Because the optical absorption of graphene relates to the chemical potential μc (Fermi level) which can be adjusted by external voltages, we can modulate optical signals by regulating the voltage on gold electrodes. The refractive index of the materials, Ag, Si, and SiO2, are measured by Palik and can be obtained from the FDTD Solutions directly [52]. At the working wavelength of 1.55 μm, the relative permittivity of hBN (εhBN) is 2.94 measured by Woessner, which can be used to prevent graphene from being chemically doped by other materials and enhanced the charge mobility [53].

Figure 1: Structure of the GOSM with orthogonal HPWs. (a) Three-dimensional schematic illustration and (b) cross-section of the modulator.
Figure 1:

Structure of the GOSM with orthogonal HPWs. (a) Three-dimensional schematic illustration and (b) cross-section of the modulator.

Light coupling input/output of the modulator can be realized by conventional waveguide grating couplers, which couples lights through the Bragg diffraction of grating if the light incidents on the grating with a specific angle [54]. After lights are coupled into the modulator, the orthogonal HPWs confines the optical mode field to the size of subwavelength to enhance the light-graphene interaction, and the optical absorption changes with the shift of the chemical potential μc adjusted by external voltages. Figure 2 (a) shows the real and imaginary part changes of the effective refractive index (neff) when the chemical potential changes from 0.2 to 0.8 eV. As shown in Figure 2(a), the imaginary part of neff increases sharply first and then decreases, which has a maximum at the chemical potential of μc = 0.51 eV and a minimum at the chemical potential of μc = 0.41 eV, suggesting that the proposed modulator achieves the maximum and minimum absorption at μc = 0.51 eV and μc = 0.41 eV. Because the real part difference of neff between μc = 0.51 eV and μc = 0.41 eV is 0.027, the phase change of the incident light is small and can be ignored. Thus, μc = 0.51 eV and μc = 0.41 eV are chosen as the ‘OFF’ state and ‘ON’ states at the working wavelength of 1.55 μm. Figure 2(b) is the corresponding optical absorption curve, the light field distribution of ‘ON’ and ‘OFF’ states in the illustration shows the light field is well confined to the subwavelength region, which can enhance the light-graphene interaction to improve the performance of modulator. Hence, the loss of modulator mainly includes the insertion loss caused by the optical absorption at ‘ON’ states and the coupling loss of light, the coupling loss of light consists the coupling loss of waveguide grating couplers and the connection loss between the coupler and the modulator. The connection loss can be calculated by 10log10(S1/S2), where S1 is the light input areas of the modulator, and S2 is the light output areas of the waveguide grating coupler. If the coupler is exactly aligned with the modulator and S2=S1=2wt1, the connection loss can be ignored [55]. Because the coupling efficiency of conventional waveguide grating couplers can higher than 50%, the coupling loss of modulator may below 6 dB, including coupling input loss and coupling output loss [55], [56].

Figure 2: (a) Real and Imaginary part of neff. (b) Attenuation constants at different μc with the width of w = 100 nm and the thickness of t1 = 20 nm.
Figure 2:

(a) Real and Imaginary part of neff. (b) Attenuation constants at different μc with the width of w = 100 nm and the thickness of t1 = 20 nm.

Here, graphene is modeled as an anisotropic material because of the single layer of carbon atoms and the π electrons, which make the electric conduction can only achieve in the plane [1]. Thus, the conductivity of graphene can be divided into the in-plane conductivity σ|| and the out of-plane conductivity σ. The in-plane conductivity σ|| is parallel to the graphene plane, which can be approximately calculated by the Kubo formalisms as follows [51], [57], [58]:

(1)σ||(ω)=2ie2kBTπ2(ω+iτ1)ln[2cosh(μc2kBT)]+e24{12+1πtan1(ω2μc2kBT)i2πln[(ω+2μc)2(ω2μc)2+(2kBT)2]}

where e is the charge of electrons, kB is the Boltzmann constant, T is the temperature, is the reduced Planck constant, ω is the angular frequency, and τ is the interband relaxation time.

The permittivity used for the out-of-plane direction in graphene ε(ω) is 2.5, while the in-plane relative permittivity ε||(ω) of graphene as the function of chemical potential can be calculated by [37], [59]:

(2)ε||(ω)=1+iσ||(ω)ε0ωd

where ε0 is the permittivity in vacuum, and d is the thickness of graphene. The calculation results with λ = 1550 nm, T = 300 K, τ = 0.1 ps are shown in Figure 3. We can see that the permittivity ε(ω) dramatically changes as the increase of chemical potential, which means that the effective mode index (neff) of the modulator can be adjusted by altering the chemical potential of graphene.

Figure 3: Permittivity of graphene as the function of chemical potential.
Figure 3:

Permittivity of graphene as the function of chemical potential.

More interesting, the chemical potential of graphene will shift with external voltages and can be expressed as [1], [31]:

(3)μc=vFπε0εhBNde|VgV0|

where vF = 0.9106 m/s is the Fermi velocity, Vg is the applied basis voltage, V0 is the voltage offset, d is the thickness of hBN, e is the charge of electrons, and εhBN is the permittivity of hBN. Thus, the neff of the modulator can be adjusted by the applied voltage Vg, which can be used to modulate optical signals.

To quantify the performance of the modulator, we define the neff, attenuation constant (α), and modulation depth (MD), as follows [60], [61]:

(4)neff=βλ02π
(5)α=10Im(neff)4πλ0ln10
(6)MD = α(OFF)α(ON)

where, λ0 is the wavelength in free space, β is the propagation constant, Im(neff) is the imaginary part of neff. α (ON) and α (OFF) are the values of α at ‘ON’ and ‘OFF’ states respectively.

3 Simulation results and analysis

3.1 Effect of the geometric parameters of HPWs

The HPW usually composes of three individual layers: a high-index layer, a metal layer, and a low-index layer. Here, we will discuss the effects of the thickness t1 and the width w of the high-index layer and metallic layer (the high-index layer and metallic layer have the same width and thickness), which is critical in determining the neff of modulator. In the simulations, the thickness of graphene and hBN are respectively fixed as 0.7 nm (the thickness of single graphene layer [1], [37], [62], [63], [64]) and 5 nm, while the width and the thickness of hBN are respectively set as 100 and 5 nm. Figure 4(a) shows the imaginary part changes of neff with the increase of t1 and μc (w is fixed to 100 nm). With the t1 increases from 10 to 55 nm, the imaginary part of neff increases sharply and then decreases with a peak value at t1 = 12 nm because the large t1 will induce a weak optical mode confinement. When the t1 is set as 20 nm, the neff changes with the increase of w and μc, as presented in Figure 4(b). Similar to Figure 4(a), the imaginary part of neff also has a peak value at w = 30 nm, but the changes are smaller than t1, which means that the influence of t1 on the refractive index is much greater than w. To comprehensively consider the effect of the thickness t1 and width w on the modulation performance, we calculate the attenuation constants α at ‘ON’ and ‘OFF’ states. Figure 5(a) and (b) show the shift of α at the chemical potential of μc = 0.41 eV and μc = 0.51 eV with the increase of t1 and w, which can be used to calculate modulation depth. As shown in Figure 6, the maximum modulation depth is 16.55 dB/μm, where w = 240 nm, and t1 = 12 nm.

Figure 4: Imaginary part of neff (a) at different t1 and μc, when w = 100 nm and λ = 1550 nm (b) at different w and μc, when t1 = 20 nm and λ = 1550 nm.
Figure 4:

Imaginary part of neff (a) at different t1 and μc, when w = 100 nm and λ = 1550 nm (b) at different w and μc, when t1 = 20 nm and λ = 1550 nm.

Figure 5: Shifts of (a) α at μc = 0.41 eV, (b) α at μc = 0.51 eV with the increase of t1 and w.
Figure 5:

Shifts of (a) α at μc = 0.41 eV, (b) α at μc = 0.51 eV with the increase of t1 and w.

Figure 6: Shifts of modulation depth with the increase of t1 and w.
Figure 6:

Shifts of modulation depth with the increase of t1 and w.

Besides, the thickness of the low-index SiO2 layer (t2) also has a positive effect on the performance. Here, we set w = 240 nm, t1 = 12 nm and analyze the effect of the thickness of SiO2 layer (t2) on α and modulation depth. As presented in Figure 7(a), with the t2 increases from 13 to 25 nm, α(OFF) increases sharply first and then decreases with a peak value of 37.60 dB/μm at t2 = 16 nm, while α(ON) increases first and then decreases slowly. Thus, the maximum modulation depth is calculated to be 26.12 dB/μm at t2 = 16 nm, as shown in Figure 7(b).

Figure 7: (a) Changes of α at ‘ON’ and ‘OFF’ states, and (b) modulation depth with the increase of t2.
Figure 7:

(a) Changes of α at ‘ON’ and ‘OFF’ states, and (b) modulation depth with the increase of t2.

3.2 Effect of the width of graphene

As the active medium of GOSM, the geometric parameters of graphene will directly affect the performance because the optical signals are modulated by adjusting the permittivity of graphene with external voltages. Here, the width of graphene will be discussed, and the results are shown in Figure 8. The thickness of the Si, SiO2 and graphene are set as 12, 16 and 0.7 nm respectively, while the width of Si is fixed as 240 nm. Figures 8(a) and (b) show the variation curves of α and modulation depth with the increase of the width of graphene. As shown in Figure 8(a), with the width of graphene increases from 10 to 210 nm, α(OFF) increases first and then tends to be a stable value of 37.72 dB/μm, while α(ON) has an equilibrium value of 11.50 dB/μm. These indicating that the enhancement of modulation depth is very limited as the width greater than 110 nm, as shown in Figure 8(b). When the width of graphene is set to be 110 nm, the modulation depth of 26.20 dB/μm is obtained.

Figure 8: (a) Changes of α at ‘ON’ and ‘OFF’ states, and (b) modulation depth with the increase of the width of graphene.
Figure 8:

(a) Changes of α at ‘ON’ and ‘OFF’ states, and (b) modulation depth with the increase of the width of graphene.

3.3 Performance of the modulator

To study the modulation performance, we discuss the operating voltage, 3 dB modulation bandwidth, and energy consumption of the modulator at the working wavelength of 1.55 μm. In the simulations, Si layers have a thickness of 12 nm and a width of 240 nm, and the SiO2 has a thickness of 16 nm, while the graphene has a width of 110 nm. As mentioned before, the chemical potential can be flexibly regulated by external voltages, and the voltage difference of ΔV = 2.56 V can switch the ‘ON’ (μc = 0.41 eV) and ‘OFF’ (μc = 0.51 eV) states. To estimate 3 dB modulation bandwidth and energy consumption, the modulator is regarded as an equivalent graphene-hBN-graphene capacitor model [24], [58]. So, the 3 dB modulation bandwidth and energy consumption can be calculated by f3dB=1/2πRC and Ebit=1/4CΔV2, respectively [44], [51], [61]. Here, C is the capacitance of capacitor, which can be obtained from C=ε0εhBNS/d. ε0 is the permittivity of vacuum, d is the thickness of hBN, and S is the footprint of modulator (the overlap area of two graphene layers separated by hBN). Due to the length of the active region in the modulator is 3 μm, the S of 0.33 μm2 (0.11 × 3 μm) and the C of 1.72 fF are obtained. R is the resistance, which mainly comes from gold-graphene contact resistance Rc and graphene sheet resistance Rg. Because the graphene sheet resistance is very small and can be safely neglected for the tiny footprint, we chose R=Rc=200Ω as the contact resistance to facilitate the comparison with recently reported GOSMs [51]. Hence, the 3 dB modulation bandwidth and energy consumption are calculated to be 462.77 GHz and 2.82 fJ/bit, respectively.

Besides, the characteristic of working wavelengths also plays a crucial role in evaluating the performance of modulators. To verify the broadband property, we investigate the wavelength-dependent response of the modulator. Figure 9(a–d) show the characteristics of modulation depth, voltage difference, energy consumption, and 3 dB modulation bandwidth with the working wavelength ranges from 1.3 to 2 μm, respectively. As shown in Figure 9, the modulation depth, voltage difference, and energy consumption decrease with the increase of working wavelengths, while the 3 dB modulation bandwidth increases. These changes mainly depend on the permittivity of graphene. As the working wavelength increases from 1.3 to 2 μm, the modulation depth decreases from 38.40 to 8.58 dB/μm, which still higher than conventional modulators and satisfies the needs of practical applications. The voltage difference and the energy consumption decrease from 3.40 to 1.39 V, and 4.97 to 0.83 fJ/bit respectively because the working point of graphene will shift towards smaller voltage, which means that the optical signals at higher wavelength requires less energy for modulation. Unlike modulation depth, voltage difference, and energy consumption, the 3 dB modulation bandwidth mainly relates to the permittivity of hBN. Thus, the variation of 3 dB modulation bandwidth is very small because the permittivity of hBN is approximately maintained a constant over the whole wavelength range. These results indicate that the proposed modulator has an excellent modulation performance and a wide wavelength tolerance, which are greatly superior to conventional GOSMs.

Figure 9: Broadband working characteristics of the modulator. (a) Modulation depth, (b) voltage difference between ‘OFF’ and ‘ON’, (c) energy consumption, and (d) 3 dB modulation bandwidth.
Figure 9:

Broadband working characteristics of the modulator. (a) Modulation depth, (b) voltage difference between ‘OFF’ and ‘ON’, (c) energy consumption, and (d) 3 dB modulation bandwidth.

3.4 Discussion

In this work, we propose a performance-enhanced GOSM with orthogonal HPWs. HPW usually includes a metal layer, a high index material layer, and a low index material layer, which can strongly confine optical mode fields in the low-index layer because of the high index difference between the layers. With the strong light-graphene interaction caused by the orthogonal HPWs, the proposed GOSM shows an excellent performance with the modulation depth of 26.20 dB/μm, the 3 dB modulation bandwidth of 462.77 GHz, and the energy consumption of 2.82 fJ/bit at 1.55 μm. Table 1 presents the performance comparison of some recently reported GOSMs. As shown in Table 1, Liu et al. designed GOSMs through placing single- or double-layer graphene on a silicon waveguide. However, to enhance modulation performance, it often requires a longer interaction length because the modulation of optical absorption mainly based on the evanescent field, which will cause vast energy consumption [24], [36]. Improving light-graphene interaction may be an effective way to enhance the performance of GOSM. Ye et al. increased the light-graphene interaction by strong coupling of HSPPs between the metal nanoribbons, leading to a modulation depth of 11.3 dB/μm, but the large 3 dB modulation bandwidth (380.23 GHz) mainly depends on the low contact resistance which is difficult in actual processing [61]. Huang et al. enhanced the performance of GOSM by localized plasmon resonance, but the working bandwidth is limited [44]. To improve the performance of modulator with wide working bandwidth, Chen et al. introduced HPW into modulators, but it is still limited by the weaker light-graphene [51]. Here, we construct two HPWs at the orthogonal directions on the cross-section of the silicon modulator. Because the optical mode field can be highly confined in the low-index layer due to the high index difference between the layers of HPWs, the interaction region between the optical mode field and graphene is compressed twice, and the light-graphene interaction is significantly enhanced, which make the modulator has a high modulation depth and a small capacitance. Because the 3 dB modulation bandwidth is inversely proportional to the capacitance while the energy consumption is proportional to the capacitance, the 3 dB modulation bandwidth and energy consumption of the modulator can be further improved and reduced, respectively, suggesting that the modulation performance can be flexibly regulated according to actual requirements through controlling the capacitance.

Table 1:

Comparison of recently reported GOSMs.

ReferenceModulation depth (dB/μm)Footprint (μm2)3 dB modulation bandwidth (GHz)Energy consumption (fJ/bit)Working bandwidth (μm)
Liu et al. [24]0.1251.28801.35 ∼ 1.6
Liu et al. [36]0.168011000NA
Ye et al. [61]11.31.8380.2329.391.3 ∼ 1.8
Huang et al. [44]6 (dB)0.24000.51.52 ∼ 1.56
Chen et al. [51]13.750.6190.57.681.2 ∼ 1.8
This work26.200.33462.772.821.3 ∼ 2

The loss of modulator mainly comes from the insertion loss and coupling loss of light. The result suggests that the modulator has an insertion loss of 11.5 dB/μm caused by the optical absorption on the ‘ON’ state, which is smaller than the modulation depth and has a less impact on the performance of modulator. The light coupling loss includes the coupling loss of waveguide grating coupler and the connecting loss between the coupler and the modulator. For conventional waveguide grating couplers, the coupling efficiency is almost 50%, suggesting the modulator may have a coupling loss of 6 dB. But recent researches also manifest that the coupling loss of waveguide grating coupler can be reduced by improving the structure. Li et al. achieved a coupling loss lower than 1.5 dB by designing a grating coupler on a silicon layer of 340 nm [56], and Ding et al. realized the coupling loss of 0.58 dB by using aluminum material as metal reflector to design subwavelength photonic crystal grating structure [65], which suggest that the coupling loss of modulator can reduce ∼3 dB. Considering that the light input area of the modulator is equal to the light output area of the waveguide grating coupler (S=S2=S1=2wt1), the connection loss comes from the mismatch between the coupler and the modulator. If the offset is ΔS, the connection loss can be calculated by 10log10((SΔS)/S), which relates to the machining error and can be solved by improving the manufacturing processes. Thus, we analyze the performance of the modulator under ideal conditions and ignore the coupling loss in the simulation, which may lead to a small deviation but does not affect the overall performance.

According to the discussion above, the proposed GOSM has an excellent performance, which can be operated with advanced silicon-based optoelectronic components on a unified silicon-based platform for reducing the volume of photoelectric devices.

4 Conclusion

In conclusion, we have proposed a GOSM based on orthogonal HPWs at near-infrared wavelengths. With the strong optical mode field confinement of HPWs, the light-graphene interaction is significantly improved, which directly affect the modulation performance. The results show that the proposed modulator has a modulation depth of 26.20 dB/μm and a 3 dB modulation bandwidth of 462.77 GHz with only 0.33 μm2 footprint and 2.82 fJ/bit energy consumption at the working wavelength of 1.55 μm. Even working in a wide wavelength range from 1.3 to 2 μm, the modulator also has excellent modulation performance with the voltage difference is below 3.40 V, the modulation depth is over 8.58 dB/μm, the 3 dB modulation bandwidth is over 462.42 GHz, and the energy consumption is below 4.97 fJ/bit. Compared with conventional GOSMs, this modulator presents priority characteristics in operation bandwidth, energy consumption, and 3 dB modulation bandwidth, which may have great potential prospects in various integrated interconnects and optoelectronic devices.


Corresponding author: Shuqing Chen, International Collaborative Laboratory of 2D Materials for Optoelectronics Science & Technology of Ministry of Education, Engineering Technology Research Center for 2D Material Information Function Devices and Systems of Guangdong Province, Institute of Microscale Optoelectronics, Shenzhen University, Shenzhen, 518060, PR China, E-mail:

Award Identifier / Grant number: 61805149

Award Identifier / Grant number: 61805087

Funding source: Guangdong Natural Science Foundation

Award Identifier / Grant number: 2020A1515011392

Award Identifier / Grant number: 2016A030310065

Award Identifier / Grant number: 2018A030313368

Funding source: Program of Fundamental Research of Shenzhen Science and Technology Plan

Award Identifier / Grant number: JCYJ20180507182035270

Funding source: Science and Technology Planning Project of Guangdong Province

Award Identifier / Grant number: 2016B050501005

Funding source: Science and Technology Project of Shenzhen

Award Identifier / Grant number: ZDSYS201707271014468

Funding source: International Collaborative Laboratory of 2D Materials for Optoelectronics Science and Technology

Award Identifier / Grant number: 2DMOST2018003

Acknowledgments

National Natural Science Foundation of China (NSFC) (61805149 and 61805087); Guangdong Natural Science Foundation (2020A1515011392, 2016A030310065, and 2018A030313368); Program of Fundamental Research of Shenzhen Science and Technology Plan (JCYJ20180507182035270); Science and Technology Planning Project of Guangdong Province (2016B050501005; Science and Technology Project of Shenzhen (ZDSYS201707271014468); International Collaborative Laboratory of 2D Materials for Optoelectronics Science and Technology (2DMOST2018003).

  1. Conflicts of interest: There are no conflicts to declare.

References

[1] X. Hu and J. Wang, “Design of graphene-based polarization-insensitive optical modulator,” Nanophotonics, vol. 7, pp. 651–658, 2018, https://doi.org/10.1515/nanoph-2017-0088.Suche in Google Scholar

[2] D. A. B. Miller, “Are optical transistors the logical next step?” Nat. Photonics, vol. 4, pp. 3–5, 2010, https://doi.org/10.1038/nphoton.2009.240.Suche in Google Scholar

[3] Z. P. Sun, A. Martinez, and F. Wang, “Optical modulators with 2D layered materials,” Nat. Photonics, vol. 10, pp. 227–238, 2016, https://doi.org/10.1038/nphoton.2016.15.Suche in Google Scholar

[4] C. Haffner, D. Chelladurai, Y. Fedoryshyn, et al., “Low-loss plasmon-assisted electro-optic modulator,” Nature, vol. 556, pp. 483–486, 2018, https://doi.org/10.1038/s41586-018-0031-4.Suche in Google Scholar

[5] G. T. Reed, G. Mashanovich, F. Y. Gardes, and D. J. Thomson, “Silicon optical modulators,” Nat. Photonics, vol. 4, pp. 518–526, 2010, https://doi.org/10.1038/nphoton.2010.179.Suche in Google Scholar

[6] W. E. Ross, D. Psaltis, and R. H. Anderson, “Two-dimensional magneto-optic spatial light modulator for signal processing,” Opt. Eng., vol. 22, pp. 485–490, 1983, https://doi.org/10.1117/12.933713.Suche in Google Scholar

[7] G. Cocorullo, M. Iodice, and I. Rendina, “All-silicon Fabry–Perot modulator based on the thermo-optic effect,” Opt. Lett., vol. 19, pp. 420–422, 1994, https://doi.org/10.1364/ol.19.000420.Suche in Google Scholar

[8] J. F. Liu, M. Beals, A. Pomerene, et al., “Waveguide-integrated, ultralow-energy GeSi electro-absorption modulators,” Nat. Photonics, vol. 2, pp. 433–437, 2008, https://doi.org/10.1038/nphoton.2008.99.Suche in Google Scholar

[9] B. Sensale-Rodriguez, R. S. Yan, S. Rafique, et al., “Extraordinary control of terahertz beam reflectance in graphene electro-absorption modulators,” Nano Lett., vol. 12, pp. 4518–4522, 2012, https://doi.org/10.1021/nl3016329.Suche in Google Scholar

[10] E. L. Wooten, K. M. Kissa, A. Yi-Yan, et al., “A review of lithium niobate modulators for fiber-optic communications systems,” IEEE J. Sel. Top. Quantum Electron., vol. 6, pp. 69–82, 2000, https://doi.org/10.1109/2944.826874.Suche in Google Scholar

[11] R. A. Griffin, S. K. Jones, N. Whitbread, S. C. Heck, and L. N. Langley, “InP Mach–Zehnder modulator platform for 10/40/100/200-Gb/s operation,” IEEE J. Sel. Top. Quantum Electron., vol. 19, p. 3401209, 2013, https://doi.org/10.1109/JSTQE.2013.2270280.Suche in Google Scholar

[12] C. R. Doerr, L. Zhang, R. J. Winzer, et al., “Compact high-speed InP DQPSK modulator,” IEEE Photon. Technol. Lett., vol. 19, pp. 1184–1186, 2007, https://doi.org/10.1109/LPT.2007.901588.Suche in Google Scholar

[13] A. S. Liu, R. Jones, L. Liao, et al., “A high-speed silicon optical modulator based on a metal–oxide–semiconductor capacitor,” Nature, vol. 427, pp. 615–618, 2004, https://doi.org/10.1038/nature02310.Suche in Google Scholar

[14] J. F. Ding, R. Q. Ji, L. Zhang, and L. Yang, “Electro-optical response analysis of a 40 Gb/s silicon Mach-Zehnder optical modulator,” J. Lightwave Technol., vol. 31, pp. 2434–2440, 2013, https://doi.org/10.1109/JLT.2013.2262522.Suche in Google Scholar

[15] Q. Xu, B. Schmidt, S. Pradhan, and M. Lipson, “Micrometre-scale silicon electro-optic modulator,” Nature, vol. 435, pp. 325–327, 2005, https://doi.org/10.1038/nature03569.Suche in Google Scholar

[16] Q. F. Xu, S. Manipatruni, B. Schmidt, J. Shakya, and M. Lipson, “12.5 Gbit/s carrier-injection-based silicon micro-ring silicon modulators,” Opt. Express, vol. 15, pp. 430–436, 2007, https://doi.org/10.1364/OE.15.000430.Suche in Google Scholar

[17] Y. H. Kuo, Y. K. Lee, Y. S. Ge, et al., “Strong quantum-confined Stark effect in germanium quantum-well structures on silicon,” Nature, vol. 437, pp. 1334–1336, 2005, https://doi.org/10.1038/nature04204.Suche in Google Scholar

[18] Y. B. Tang, J. D. Peters, and J. E. Bowers, “Over 67 GHz bandwidth hybrid silicon electroabsorption modulator with asymmetric segmented electrode for 1.3 μm transmission,” Opt. Express, vol. 20, pp. 11529–11535, 2012, https://doi.org/10.1364/OE.20.011529.Suche in Google Scholar

[19] J. A. Dionne, K. Diest, L. A. Sweatlock, and H. A. Atwater, “PlasMOStor: a metal− oxide− Si field effect plasmonic modulator,” Nano Lett., vol. 9, pp. 897–902, 2009, https://doi.org/10.1021/nl803868k.Suche in Google Scholar

[20] P. Chaisakul, D. Marris-Morini, M. S. Rouifed, et al., “23 GHz Ge/SiGe multiple quantum well electro-absorption modulator,” Opt. Express, vol. 20, pp. 3219–3224, 2012, https://doi.org/10.1364/OE.20.003219.Suche in Google Scholar

[21] S. J. Zhang, Y. Liu, R. G. Lu, B. Sun, and L. S. Yan, “Heterogeneous III–V silicon photonic integration: components and characterization,” Front. Inform. Tech. EI., vol. 20, pp. 472–480, 2019, https://doi.org/10.1631/FITEE.1800482.Suche in Google Scholar

[22] F. Bonaccorso, Z. Sun, T. Hasan, and A. C. Ferrari, “Graphene photonics and optoelectronics,” Nat. Photonics, vol. 4, pp. 611–622, 2010, https://doi.org/10.1038/NPHOTON.2010.186.Suche in Google Scholar

[23] L. L. Miao, Y. Q. Jiang, S. B. Lu, et al., “Broadband ultrafast nonlinear optical response of few-layers graphene: toward the mid-infrared regime,” Photonics Res., vol. 3, pp. 214–219, 2015, https://doi.org/10.1364/PRJ.3.000214.Suche in Google Scholar

[24] M. Liu, X. B. Yin, E. Ulin-Avila, et al., “A graphene-based broadband optical modulator,” Nature, vol. 474, pp. 64–67, 2011, https://doi.org/10.1038/nature10067.Suche in Google Scholar

[25] J. Zou, S. L. Wu, Y. Liu, et al., “An ultra-sensitive electrochemical sensor based on 2D g-C3N4/CuO nanocomposites for dopamine detection,” Carbon, vol. 130, pp. 652–663, 2018, https://doi.org/10.1016/j.carbon.2018.01.008.Suche in Google Scholar

[26] J. Zou, D. P. Mao, A. T. S. Wee, and J. Z. Jiang, “Micro/nano-structured ultrathin g–C3N4/Ag nanoparticle hybrids as efficient electrochemical biosensors for L-tyrosine,” Appl. Surf. Sci., vol. 467, pp. 608–618, 2019,.10.1016/j.apsusc.2018.10.187Suche in Google Scholar

[27] R. R. Nair, P. Blake, A. N. Grigorenko, et al., “Fine structure constant defines visual transparency of graphene,” Science, vol. 320, pp. 1308–1308, 2008, https://doi.org/10.1126/science.1156965.Suche in Google Scholar

[28] C. Y. Guan, J. H. Shi, J. L. Liu, et al., “Recent progress in photonic synapses for neuromorphic systems,” Laser Photonics Rev. , vol. 13, p. 1800242, 2019, https://doi.org/10.1002/aisy.201900136.Suche in Google Scholar

[29] K. I. Bolotin, K. J. Sikes, Z. Jiang, et al., “Ultrahigh electron mobility in suspended graphene,” Solid State Commun., vol. 146, pp. 351–355, 2008, https://doi.org/10.1016/j.ssc.2008.02.024.Suche in Google Scholar

[30] X. Du, I. Skachko, A. Barker, and E. Y. Andrei, “Approaching ballistic transport in suspended graphene,” Nat. Nanotechnol., vol. 3, pp. 491–495, 2008, https://doi.org/10.1038/nnano.2008.199.Suche in Google Scholar

[31] F. Wang, Y. B. Zhang, C. S. Tian, et al., “Gate-variable optical transitions in graphene,” Science, vol. 320, pp. 206–209, 2008, https://doi.org/10.1126/science.1152793.Suche in Google Scholar

[32] X. T. Gan, R. J. Shiue, Y. D. Gao, et al., “High-contrast electrooptic modulation of a photonic crystal nanocavity by electrical gating of graphene,” Nano Lett., vol. 13, pp. 691–696, 2013, https://doi.org/10.1021/nl304357u.Suche in Google Scholar

[33] Q. L. Bao and K. P. Loh, “Graphene photonics, plasmonics, and broadband optoelectronic devices,” ACS Nano, vol. 6, pp. 3677–3694, 2012, https://doi.org/10.1021/nn300989.Suche in Google Scholar

[34] N. Youngblood, Y. Anugrah, R. Ma, S. J. Koester, and M. Li, “Multifunctional graphene optical modulator and photodetector integrated on silicon waveguides,” Nano Lett., vol. 14, pp. 2741–2746, 2014, https://doi.org/10.1021/nl500712u.Suche in Google Scholar

[35] Z. L. Lu and W. S. Zhao, “Nanoscale electro-optic modulators based on graphene-slot waveguides,” J. Opt. Soc. Am. B, vol. 29, pp. 1490–1496, 2012, https://doi.org/10.1364/JOSAB.29.001490.Suche in Google Scholar

[36] M. Liu, X. B. Yin, and X. Zhang, “Double-layer graphene optical modulator,” Nano Lett., vol. 12, pp. 1482–1485, 2012, https://doi.org/10.1021/nl204202k.Suche in Google Scholar

[37] A. Phatak, Z. Z. Cheng, C. Y. Qin, and K. Goda, “Design of electro-optic modulators based on graphene-on-silicon slot waveguides,” Opt. Lett., vol. 41, pp. 2501–2504, 2016, https://doi.org/10.1364/OL.41.002501.Suche in Google Scholar

[38] M. Midrio, S. Boscolo, M. Moresco, et al., “Graphene-assisted control of coupling between optical waveguides,” Opt. Express, vol. 20, pp. 28479–28484, 2012, https://doi.org/10.1364/OE.20.028479.Suche in Google Scholar

[39] T. Pan, C. Y. Qiu, J. Y. Wu, et al., “Analysis of an electro-optic modulator based on a graphene-silicon hybrid 1D photonic crystal nanobeam cavity,” Opt. Express, vol. 23, pp. 23357–23364, 2015, https://doi.org/10.1364/OE.23.023357.Suche in Google Scholar

[40] Y. Yao, R. Shankar, M. A. Kats, et al., “Electrically tunable metasurface perfect absorbers for ultrathin mid-infrared optical modulators,” Nano Lett., vol. 14, pp. 6526–6532, 2014, https://doi.org/10.1021/nl503104n.Suche in Google Scholar

[41] D. Ansell, I. P. Radko, Z. Han, F. J. Rodriguez, S. I. Bozhevolnyi, and A. N. Grigorenko, “Hybrid graphene plasmonic waveguide modulators,” Nat. Commun., vol. 6, p. 8846, 2015, https://doi.org/10.1038/ncomms9846.Suche in Google Scholar

[42] J. S. Shin and J. T. Kim, “Broadband silicon optical modulator using a graphene-integrated hybrid plasmonic waveguide,” Nanotechnology, vol. 26, p. 365201, 2015, https://doi.org/10.1088/0957-4484/26/36/365201.Suche in Google Scholar

[43] C. Y. Qiu, W. L. Gao, R. Vajtai, P. M. Ajayan, J. Kono, and Q. F. Xu, “Efficient modulation of 1.55 μm radiation with gated graphene on a silicon microring resonator,” Nano Lett., vol. 14, pp. 6811–6815, 2014, https://doi.org/10.1021/nl502363u.Suche in Google Scholar

[44] B. H. Huang, W. B. Lu, Z. G. Liu, and S. P. Gao, “Low-energy high-speed plasmonic enhanced modulator using graphene,” Opt. Express, vol. 26, pp. 7358–7367, 2018, https://doi.org/10.1364/OE.26.007358.Suche in Google Scholar

[45] Z. Z. Ma, M. H. Tahersima, S. Khan, and V. J. Sorger, “Two-dimensional material-based mode confinement engineering in electro-optic modulators,” IEEE J. Sel. Top. Quantum Electron., vol. 23, pp. 81–88, 2017, https://doi.org/10.1109/JSTQE.2016.2574306.Suche in Google Scholar

[46] L. Huang, G. H. Hu, C. Y. Deng, Y. Zhu, B. F. Yun, and R. H. Zhang, “Realization of mid-infrared broadband absorption in monolayer graphene based on strong coupling between graphene nanoribbons and metal tapered grooves,” Opt. Express, vol. 26, pp. 29192–29202, 2018, https://doi.org/10.1364/OE.26.029192.Suche in Google Scholar

[47] L. Chen, Y. M. Liu, Z. Y. Yu, et al., “Numerical investigations of a near-infrared plasmonic refractive index sensor with extremely high figure of merit and low loss based on the hybrid plasmonic waveguide-nanocavity system,” Opt. Express, vol. 24, pp. 23260–23270, 2016, https://doi.org/10.1364/OE.24.023260.Suche in Google Scholar

[48] X. D. Yang, Y. M. Liu, R. F. Oulton, X. B. Yin, and X. A. Zhang, “Optical forces in hybrid plasmonic waveguides,” Nano Lett., vol. 11, pp. 321–328, 2011, https://doi.org/10.1021/nl103070n.Suche in Google Scholar

[49] R. F. Oulton, V. J. Sorger, D. A. Genov, D. F. P. Pile, and X. Zhang, “A hybrid plasmonic waveguide for subwavelength confinement and long-range propagation,” Nat. Photonics, vol. 2, pp. 496–500, 2008, https://doi.org/10.1038/nphoton.2008.131.Suche in Google Scholar

[50] J. T. Kim, J. J. Ju, S. Park, M. Kim, S. K. Park, and S. Y. Shin, “Hybrid plasmonic waveguide for low-loss lightwave guiding,” Opt. Express, vol. 18, pp. 2808–2813, 2010, https://doi.org/10.1364/OE.18.002808.Suche in Google Scholar

[51] X. Chen, Y. Wang, Y. J. Xiang, et al., “A broadband optical modulator based on a graphene hybrid plasmonic waveguide,” J. Lightwave Technol., vol. 34, pp. 4948–4953, 2016, https://doi.org/10.1109/JLT.2016.2612400.Suche in Google Scholar

[52] E. D. Palik. Handbook of Optical Constants of Solids. San Diego, USA: Academic Press, 1998.Suche in Google Scholar

[53] A. Woessner, M. B. Lundeberg, Y. Gao, et al., “Highly confined low-loss plasmons in graphene-boron nitride hetero-structures,” Nat. Mater., vol. 14, pp. 421–425, 2014, https://doi.org/10.1038/NMAT4169.Suche in Google Scholar

[54] R. Hao, J. Y. Jiao, X. L. Pend, et al., “Experimental demonstration of a graphene-based hybrid plasmonic modulator,” Opt. Lett., vol. 44, pp. 2586–2589, 2019, https://doi.org/10.1364/OL.44.002586.Suche in Google Scholar

[55] F. van Laere, M. Ayre, D. Taillaert, D. van Thourhout, T. F. Krauss, and R. Baets, “Compact and highly efficient grating couplers between optical fiber and nanophotonic waveguides,” J. Lightwave Technol., vol. 25, pp. 151–156, 2007, https://doi.org/10.1109/JLT.2006.888164.Suche in Google Scholar

[56] C. Li, H. J. Zhang, M. B. Yu, and G. Q. Lo, “CMOS-compatible high efficiency double-etched apodized waveguide grating coupler,” Opt. Express, vol. 21, pp. 7868–7874, 2013, https://doi.org/10.1364/OE.21.007868.Suche in Google Scholar

[57] A. H. Castro-Neto, N. M. R. Peres, K. S. Novoselov, A. K. Geim, and F. Guinea, “The electronic properties of graphene,” Rev. Modern Phys., vol. 81, pp. 109–162, 2009, https://doi.org/10.1103/RevModPhys.81.109.Suche in Google Scholar

[58] L. A. Falkovsky, “Optical properties of graphene and IV–VI semiconductors,” Phys.-Usp., vol. 51, pp. 887–897, 2008, https://doi.org/10.1070/PU2008v051n09ABEH006625.Suche in Google Scholar

[59] X. L. Peng, R. Hao, Z. W. Ye, et al., “Highly efficient graphene-on-gap modulator by employing the hybrid plasmonic effect,” Opt. Lett., vol. 42, pp. 1736–1739, 2017, https://doi.org/10.1364/OL.42.001736.Suche in Google Scholar

[60] J. A. Dionne, L. A. Sweatlock, H. A. Atwater, and A. Polman, “Plasmon slot waveguides: towards chip-scale propagation with subwavelength-scale localization,” Phys. Rev. B: Condens. Matter. Mater. Phys., vol. 73, p. 035407, 2006, https://doi.org/10.1103/PhysRevB.73.035407.Suche in Google Scholar

[61] L. F. Ye, K. Sui, Y. Zhang, and Q. H. Liu, “Broadband optical waveguide modulators based on strongly coupled hybrid graphene and metal nanoribbons for near-infrared applications,” Nanoscale, vol. 11, p. 3229, 2019, https://doi.org/10.1039/C8NR09157A.Suche in Google Scholar

[62] Y. T. Hu, M. Pantouvaki, J. V. Campenhout, et al., “Broadband 10 Gb/s operation of graphene electro-absorption modulator on silicon,” Laser Photonics Rev., vol. 10, pp. 307–316, 2016, https://doi.org/10.1002/lpor.201500250.Suche in Google Scholar

[63] J. Lu, H. Z. Tsai, A. N. Tatan, et al., “Frustrated supercritical collapse in tunable charge arrays on graphene,” Nat. Commun., vol. 10, p. 477, 2019, https://doi.org/10.1038/s41467-019-08371-2.Suche in Google Scholar

[64] J. S. Shin and J. T. Kim, “Broadband silicon optical modulator using a graphene-integrated hybrid plasmonic waveguide,” Nanotechnology, vol. 26, p. 365201, 2015, https://doi.org/10.1088/0957-4484/26/36/365201.Suche in Google Scholar

[65] Y. H. Ding, C. Peucheret, H. Y. Ou, and K. Yvind, “Fully etched apodized grating coupler on the SOI platform with -0.58 dB coupling efficiency,” Opt. Lett., vol. 39, pp. 5348–5350, 2014, https://doi.org/10.1364/OL.39.005348.Suche in Google Scholar

Received: 2020-03-05
Accepted: 2020-04-15
Published Online: 2020-05-18

© 2020 Mingyang Su et al., published by De Gruyter, Berlin/Boston

This work is licensed under the Creative Commons Attribution 4.0 International License.

Artikel in diesem Heft

  1. Reviews
  2. Multiparticle quantum plasmonics
  3. Physics and applications of quantum dot lasers for silicon photonics
  4. Integrated lithium niobate photonics
  5. Subwavelength structured silicon waveguides and photonic devices
  6. Nonlinear optical microscopies (NOMs) and plasmon-enhanced NOMs for biology and 2D materials
  7. Enhancement of upconversion luminescence using photonic nanostructures
  8. 3D Nanophotonic device fabrication using discrete components
  9. Research Articles
  10. A flexible platform for controlled optical and electrical effects in tailored plasmonic break junctions
  11. Effects of roughness and resonant-mode engineering in all-dielectric metasurfaces
  12. Colloidal quantum dots decorated micro-ring resonators for efficient integrated waveguides excitation
  13. Engineered telecom emission and controlled positioning of Er3+ enabled by SiC nanophotonic structures
  14. Diffraction engineering for silicon waveguide grating antenna by harnessing bound state in the continuum
  15. On-chip scalable mode-selective converter based on asymmetrical micro-racetrack resonators
  16. High-Q dark hyperbolic phonon-polaritons in hexagonal boron nitride nanostructures
  17. Multilevel phase supercritical lens fabricated by synergistic optical lithography
  18. Continuously-tunable Cherenkov-radiation-based detectors via plasmon index control
  19. Cherenkov radiation generated in hexagonal boron nitride using extremely low-energy electrons
  20. Geometric phase for multidimensional manipulation of photonics spin Hall effect and helicity-dependent imaging
  21. Stable blue-emissive aluminum acetylacetonate nanocrystals with high quantum yield of over 80% and embedded in polymer matrix for remote UV-pumped white light–emitting diodes
  22. Pumping-controlled multicolor modulation of upconversion emission for dual-mode dynamic anti-counterfeiting
  23. Broadband graphene-on-silicon modulator with orthogonal hybrid plasmonic waveguides
  24. Non-noble metal based broadband photothermal absorbers for cost effective interfacial solar thermal conversion
  25. Metal-insulator-metal nanoresonators – strongly confined modes for high surface sensitivity
  26. Erratum
  27. Erratum to: Darkfield colors from multi-periodic arrays of gap plasmon resonators
Heruntergeladen am 22.12.2025 von https://www.degruyterbrill.com/document/doi/10.1515/nanoph-2020-0165/html
Button zum nach oben scrollen