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

Optimized tapered dipole nanoantenna as efficient energy harvester

Open Access Open Access

Abstract

In this paper, a novel design of tapered dipole nanoantenna is introduced and numerically analyzed for energy harvesting applications. The proposed design consists of three steps tapered dipole nanoantenna with rectangular shape. Full systematic analysis is carried out where the antenna impedance, return loss, harvesting efficiency and field confinement are calculated using 3D finite element frequency domain method (3D-FEFD). The structure geometrical parameters are optimized using particle swarm algorithm (PSO) to improve the harvesting efficiency and reduce the return loss at wavelength of 500 nm. A harvesting efficiency of 55.3% is achieved which is higher than that of conventional dipole counterpart by 29%. This enhancement is attributed to the high field confinement in the dipole gap as a result of multiple tips created in the nanoantenna design. Furthermore, the antenna input impedance is tuned to match a wide range of fabricated diode based upon the multi-resonance characteristic of the proposed structure.

© 2016 Optical Society of America

1. Introduction

In early 1970s, Bailey has proposed the concept of rectenna in the field of solar energy harvesting [1] where rectenna stands for a rectifying diode coupled with antenna. Nanoantenna is responsible for transforming solar radiation into an AC electric field across the diode which rectifies the signal to obtain DC power. In recent years, due to the large demand on renewable clean energy, rectennas design and fabrication techniques are getting more mature as this approach promises to offer highly efficient solar energy harvesting systems. Referring to the radiation spectrum of the sun, maximum irradiance occurs at the visible wavelength range from 400 to 700 nm. Nonetheless, extending the utilization of rectenna systems to visible frequencies (PHz) resulted in several challenges. First, the challenge of scaling down antenna dimensions to submicrometer which is becoming more realizable with advanced technologies [2]. However, suitable rectifiers are still not available for this operation criteria. Metal\insulator\metal (MIM) diode topology is gaining interest as a candidate for high frequency operation [3] and an investigation of MIM diode limitations will be illustrated later.

The main factors contributing to the efficiency of nano-rectenna systems are:

  • (1) Harvesting efficiency of the nanoantenna device.
  • (2) The cut off frequency of the diode and responsivity at zero voltage bias.
  • (3) Impedance matching between the nanoantenna and the diode.

Vandenbosch and Ma [4,5] reported an upper bound for total solar harvesting efficiency of 64%, over wavelength range 400-1200 nm, for a silver dipole nanoantenna. This design offers a nominal input impedance of 250 Ω at resonance and harvesting efficiency of 26% at wavelength of 500 nm. Bowtie [6,7] and Vivaldi [8] nanoantennas offered a low input impedance of about 100 Ω at resonance. However, authors study didn’t provide calculations for harvesting efficiency at visible light range. Other design topologies were proposed, such as spiral, log-periodic [9], flower-shaped dipole and elliptic dipole nanoantenna [10]. While some of these designs offer higher efficiencies up to 90% at 500 nm wavelength, they suffer from very high design complexity and the lack of impedance analysis in the literature, especially over the visible light region. A tapered dipole nanoantenna topology was firstly introduced in [11] where two structures were presented, two-arm and cross arm dipole antennas with tapered end. The simulation results for these structures show high field confinement in the antenna gap where the cross dipole antenna exhibits the higher field enhancement. However, this study was not intended for solar harvesting application and thus no further analysis for the harvesting efficiency or input impedance was presented [11]. A study of the field distribution around nanoantenna and current induced on the surface was also carried out for the bowtie and spiral designs [6,9] where simulation and experimental results showed high enhancement over the conventional dipole design [5]. However, these rectenna devices are reported to suffer from low coupling efficiency between nanoantenna and rectifier due to high impedance mismatch.

To ensure highly efficient operation of rectenna system, MIM diode must obtain several characteristics. One of the main characteristics is the high responsivity which is a measure of the DC output with respect to the input power [3]. Another characteristic is the diode impedance (RD and CD) which defines the junction cut-off frequency [3] and plays the major role in coupling efficiency with the antenna device. While nanoantennas have a typical impedance of 100-200 Ω, the reported MIM diodes offer as low impedance as 0.5 – 3 kΩ [6,11] with thickness range from 4 nm down to 0.7 nm (MIM thickness is directly proportional to diode resistance [12]). This deviation between MIM diode and nanoantenna impedances reduces the coupling efficiency significantly as reported in [6,9].

It is worth emphasizing that this work doesn’t aim to model a new rectifier topology. This paper introduces a novel nanoantenna design which offers harvesting efficiencies up to 55.3% at wavelength λ of 500 nm with a field confinement 60 times greater than that of the conventional dipole design. This enhancement is attributed to high divergence of surface current over the nanoantenna surface as a result of the multiple thickness-grading introduced to the design. Moreover, the proposed structure can be tuned to perfectly match a wide range of fabricated rectifier impedances based on the multi-resonance characteristic of this novel design. Three design configurations are introduced which have 500Ω, 1kΩ and 2kΩ resonant impedances to match that of MIM diodes reported in [6,11].

Following this introduction, the theory and design consideration will be presented in section II. A description of the simulation environment and optimization technique are considered in section III. In section IV, detailed simulation results are reported which quantify the benefits of the proposed nanoantenna design topology. Finally, conclusion will be drawn in section V.

2. Theory and design consideration

This research work focuses on overcoming the drawbacks of previously studied nanoantenna designs such as: low field confinement at the gap, small nominal input impedance and low harvesting efficiency at visible light frequencies. Moreover, this study introduces a novel design that offers relatively high manufacture feasibility as will be discussed later.

For an efficient operation of rectenna system, high intensity field should exist at the surface of the MIM diode δs so that sufficient electrons can tunnel through the diode. Therefore, high field confinement at the gap of the nanoantenna represents a major requirement in the nanoantenna design. According to the electromagnetic theory, electric field tends to be accumulated at the metal tips of the antenna structure [6,13]. Based on this phenomenon, the proposed design was constructed by inserting more steps of smaller dimensions into the conventional dipole design, as shown in ‎Fig. 1, which increases the number of tips and consequently the electric field confinement by the nanoantenna. This is directly related to the antenna efficiency where the harvested electric energy is transformed to AC power at the antenna port. Figure 1 demonstrates the design parameters of the conventional dipole and the two proposed nanoantennas topologies, type 1 and type 2. Type 1 nanoantenna is considered a modified version of the conventional dipole where one more step of smaller dimensions is inserted. On the other hand, type 2 nanoantenna represents a 3-steps tapered dipole. Silver material is selected for all the nanoantennas studied in this work as it’s reported to offer higher efficiency when compared to other metals [5]. The proposed design parameters are chosen so that the overall nanoantenna volume is less than that of the conventional dipole reported in [4] to obtain a shorter device length and to have a fair comparison with the conventional dipole as larger material volume will increase the harvested light. Additionally, a minimum feature size of 5 nm is maintained in grading adjacent dipole steps to reduce design complexity.

 figure: Fig. 1

Fig. 1 (a) Conventional dipole [‎4], (b) Type 1 (2-steps tapered dipole), (c) Type 2 (3-steps tapered dipole).

Download Full Size | PDF

Further study of this design was executed by analyzing the surface current distribution over the nanoantenna surface as shown in ‎Fig. 2 at λ=500nm. The figure shows that the divergence of the current at the tips of the tapered dipole is much higher than any point in the middle which indicates high accumulation of charges and thus a greater field confinement [14]. Moreover, it’s found that a higher field confinement at the gap region is achieved by grading dipole dimensions so that the smallest stub is located towards the gap.

 figure: Fig. 2

Fig. 2 Surface current distribution over nanoantenna structures at 500 nm wavelength (the bigger arrows indicates higher divergence of current) for (a) conventional dipole, (b) type 1, and (c) type 2.

Download Full Size | PDF

As the wave propagates through the tapered dipole antenna structure, it will encounter different interfaces between adjacent dipoles stubs as shown in Fig. 1. Consequently, the number of current paths on the antenna surface increases, as illustrated in Fig. 2, which results in multiple impedance resonances [7,15]. For a multi-resonance antenna, the input impedance at a single resonant frequency can be expressed as a function of all the resonant impedances within the antenna bandwidth [16,17]. The current paths excited in the antenna are directly related to the design dimensions. Therefore, the proposed design can be configured so that its input impedance can match perfectly with wide range of fabricated diode impedances.

The suggested design can be fabricated using top-down methods such as electron-beam lithography (EBL) and focused ion beam (FIB) techniques [8]. The EBL is employed to define desired patterns down to the nanometer scale in resists. Therefore, it is normally combined with stripping (lift-off) or etching to obtain the desired patterns in the target materials. For a specific EBL technique, reported in [18], 4–8 nm patterning and lift-off were achieved for Au nanoparticles. On contrast to EBL, the FIB technique can define patterns down to the nanoscale without using masks [19]. Most widespread instruments of FIB use gallium ion source which is capable of fabricating sub-5nm holes [20]. Another technique of FIB is called Helium ion beam milling which is inherently less damaging to the sample than Ga ions but ideal for structuring thin slabs of material with high precision [21]. The helium milling was reported to achieve 5–10 nm resolution in patterning a thin gold film using milling from sides to center [21] .These fabrication methods have been commonly used in manufacturing nanoantennas with highly controllable parameters. However, the fabrication of several nanoantenna structures (such as spiral [9], elliptical and flower [10]) suffers from the complexity of the lift-off process for ultra-small features and curvatures since resists are difficult to strip away. The proposed design offers high feasibility since it doesn’t include sharp tips or curvature. Furthermore, type 1 and type 2 structures have a minimum feature size of 5 nm, to be compatible with the techniques stated above and a large gap size to facilitate the fabrication process [8]. Moreover, a true smooth taper can be more feasible than the reported graded tapered dipole. However, the stepped thickness-grading introduced in the proposed design allows for impedance configurability, as discussed in section 4.5, and a higher harvesting efficiency due to field accumulation around metal tips of the nanoantenna structure.

3. Simulation environment and numerical methodology

The proposed nano dipole design is analyzed using finite element frequency domain (FEFD) method via Comsol Multiphysics software [23] where the harvesting efficiency, return loss and input impedance are calculated at a wavelength of 500 nm. FEFD technique is widely used in analyzing optical and electric structure with high precision [24,25]. The constructed model uses a fine mesh with a minimum and maximum element size of 1 nm and 10 nm respectively in order to resolve the skin depth of silver (3 nm at 500 nm wavelength). The studied nanoantennas has a fixed thickness T of 40 nm, indicated in ‎Fig. 1, and surrounded by free space everywhere as in [4,26]. The boundary conditions are set to perfect matched layer (PML) and the permittivity of silver introduced in the model is taken from Johnson and Christy (1972) [27]. According to the reciprocity theorem, the efficiency of an antenna in the transmission mode ηrad is equal to the efficiency in receiving mode. The radiation efficiency of an antenna is calculated as [4]:

ηrad=PradPin=PradPrad+Ploss
wherePradis the total radiated power, Pinis the power input at the antenna port, and Ploss represents the power dissipated in the nanoantenna material. ‎Table 1 shows the equations used to calculatePradandPlosswhereSscis the scattered intensity poynting vector, Jis the antenna surface current, and Qrltogether with Qml represent the resistive losses and magnetic losses respectively.

Tables Icon

Table 1. Radiated Power and Power Loss Equations Extracted from the Simulated Electric and Magnetic Fields [‎23,‎‎31]

The input impedance of the proposed nanoantenna design can be configured to match a wide range of fabricated rectifier impedances. In this approach, particle swarm optimization (PSO) [28] algorithm is utilized to search for the best nanoantenna dimensions values that give the highest radiation efficiency and the desired value for the input impedance. This process is done by linking a multi-objective PSO algorithm with the external FEFD analyzer via Comsol Multiphysics software where the PSO algorithm updates the nanoantenna dimensions values and the FEFD analyzer acts as a fitness function evaluator. The fitness function of the PSO algorithm is defined as:

f= c1×ηrad+c2×|S11|
where S11 represents a measure of the reflected power at the antenna port which is directly related to the antenna impedance. The S11 is calculated in dB and has typical values of several decades in negative. Further, ηrad is the ratio presented in Eq. (1) and it’s less than unity in magnitude. Additionally, c1 and c2 are constants that are chosen as weighting factors to compensate the magnitude variation between S11 and ηrad so that both parameters have the same influence on the fitness function. Consequently, the optimization process is not carried away by a single parameter value. These constants are chosen through an iterative process where the optimization algorithm is executed repeatedly then c1 and c2 values are adjusted after each iteration in order to optimize both S11 and ηradfairly using the PSO algorithm.

For each candidate design, some constraints must be present in order to pass the values to the FEFD analyzer. These constraints are described as:

Wi> Wi+1
Wi Wi+1>5 nm
nLi×Wi const ,    n=1,2,3
where Eq. (3) ensures that the confinement occurs at the gap, while Eq. (4) is introduced to increase design feasibility and reduces manufacture cost. Further, Eq. (5) provides that optimized designs don’t have exaggerated overall. The constant value in the last equation represents a measure of the volume of the conventional dipole design in [4].

4. Simulation results

4.1 Model validation

In order to check the validity of the constructed FEFD model, a full analysis was executed for the conventional silver dipole nanoantenna [4] with L= 250 nm, W=40 nm, G=10 nm, andT=40 nm. Figure 3 shows the radiation efficiency versus the wavelength for the FEFD constructed model and that reported in [4] by the finite difference time domain (FDTD) method. It’s evident from this figure that a good agreement between both models is obtained especially at wavelength of 500 nm which is used as the operating wavelength of the optimized tapered nanoantenna. The slight deviation between the two curves can be attributed to two reasons. First, the constructed model in this study is based on FEFD solver, via Comsol Multiphysics software, which is a well-known method for its accuracy and high precision. However, the reported results by Vandenbosch and Ma [4] are based on the FDTD technique. Moreover, the meshing element used in this study has a minimum size of 1 nm which is 5 times smaller than that used in [4]. Figure 4 shows the variation of the harvesting efficiency with the wavelength at different minimum meshing sizes. It’s evident from this figure that the simulation results are independent of the mesh size which ensures the accuracy of the FEFD results. It’s also worth noting that the choice of the silver permittivity can affect the harvesting efficiency as shown in Fig. 5. The constructed model of the tapered nanoantenna is based on the permittivity values of Johnson and Christy model [27] which shows a good agreement with the experimental results [27]. Therefore, it’s concluded that the slight deviation between our results and that reported in [4] can be due to the choice of the silver permittivity model and the meshing capabilities of the FEFD and the FDTD [4].

 figure: Fig. 3

Fig. 3 Wavelength dependent harvesting efficiency for conventional dipole by the FDTD [4] and FEFD method.

Download Full Size | PDF

 figure: Fig. 4

Fig. 4 Wavelength dependent harvesting efficiency for the conventional dipole at different minimum mesh element sizes.

Download Full Size | PDF

 figure: Fig. 5

Fig. 5 Wavelength dependent harvesting efficiency for the conventional dipole using different permittivities based on experimental data [4], Johnson and Christy model [27], Palik model [‎29] along with that reported by Vandenbosch and Ma [4].

Download Full Size | PDF

4.2 Efficiency and return loss analysis

In this analysis, the dipole nanoantenna is excited at its gap by a voltage imposed between the two conducting arms of the dipole and thus corresponds to Thevinin equivalent circuit. This approach, also known as gap excitation, is previously reported to result in antenna impedance very similar to that calculated using conventional feeding line excitation [30]. A nominal diode impedance of 500 Ω was chosen which corresponds to the fabricated diode reported in [6]. This diode impedance acts as the Thevinin impedance of the source feeding the nanoantenna at the gap. Additionally, this analysis is carried out at λ of 500 nm where the maximum irradiance of the sun occurs.

Type 1 and type 2 nanoantennas are optimized to have resonance at λ=500nmwith an input impedance of 500 Ω to match the diode impedance introduced to the model. The optimization process follows the PSO algorithm to maximize the suggested fitness function (Eq. (2)). In the case of type 1 nanoantenna, the optimizer works through a 5-dimensional solution space (L1, W1,  L2, W2, G) while type 2 structure has a 7-dimensional solution space (L1, W1,  L2, W2,  L3,  W3, G). Regarding the optimization of type 2 nanoantenna design, the PSO algorithm converged to maximum value of fitness function in 72 iterations. This is illustrated in ‎Fig. 6 along with the corresponding values of S11 and ηrad at 3 points which show the effectiveness of applying PSO technique in nanoantenna design problems.

 figure: Fig. 6

Fig. 6 Fitness function (Eq. (2) values for type 2 nanoantenna versus number of iterations of the PSO algorithm. The corresponding values of S11 and ηrad are demonstrated for 3 points.

Download Full Size | PDF

The optimized tapered dipoles dimensions, introduced to the FEFD analyzer, are given in ‎Table 2 along with the dimensions of the dipole reported in [4]. Additionally, the optimized designs have smaller overall volume than the conventional dipole leading to smaller size and weight of the device, especially when used in arrays. The proposed designs show high enhancement over the conventional dipole design whereas type 1 design offers ηrad=47% and S11=23.6 dB at λ=500nm. However, type 2 design achieves ηrad=55.3% and S11=28.1 dB at the same wavelength. On the other hand, the conventional dipole [4] design offers ηrad=26% and a near zero value of return loss at λ=500nm. This is clearly illustrated in ‎Fig. 7 and ‎Fig. 8 where the harvesting efficiency and return loss of the tapered dipole designs are put in comparison with that of the conventional dipole design.

Tables Icon

Table 2. Dipole Designs Dimensions in Nanometer Introduced to the FEFD analyzer

 figure: Fig. 7

Fig. 7 Variation of harvesting efficiency versus wavelength for conventional dipole [‎‎4] and the suggested designs of type 1 and type 2 nanoantennas.

Download Full Size | PDF

 figure: Fig. 8

Fig. 8 Wavelength dependent return loss at the nanoantenna port for conventional dipole [‎‎4] and the proposed designs of type 1 and type 2 nanoantennas.

Download Full Size | PDF

Additionally, total harvesting efficiency calculations are performed for the proposed nanoantenna designs to obtain a measure of the ultimate optical efficiency over the total wavelength range of the sun irradiance. The total harvesting efficiency can be defined as [4]:

ηtotal=0P(λ,T)× ηrad(λ) dλ0P(λ,T) dλ
where λ is the wavelength, T is the absolute temperature of the black body (in K), and P is the Planck’s law for black body radiation defined in [4]. The total harvesting efficiency values for type 2, type 1 and conventional dipole nanoantenna are equal to 79.2%, 62.9%, and 64.1%, respectively. It’s worth noting that the aim of our study is to improve the nanoantenna operation at single frequency (λ=500nm) where the input impedance and field confinement are optimized. However, calculating the total harvesting efficiency for the proposed design ensures a highly efficient operation at other frequencies.

4.3 Electric field confinement analysis

To obtain the electric field distribution around the nano dipole, the whole structure is illuminated by a normally incident plane wave. The incident electric field has an intensity of 1 V/m and is polarized linearly parallel to the antenna axis. The dipoles design dimensions are listed in Table 2 as in the previous analysis. However, the gap size is held constant during this investigation at 10 nm for all design topologies due to its major effect on the field confinement through the gap. Figure 9 illustrates the electric field profile for conventional dipole, type 1 and type 2 nanoantennas at λ=500nm. As demonstrated in the figures, electric field is accumulated with high intensity around the tips of the three nanoantennas whereas type 2 topology produces the maximum field confinement in the gap. To obtain a measure of the field enhancement, the electric field vector was integrated over the gap region for the three design topologies. The simulation results show a relative intensity enhancement of 10 times of magnitude for type 1 structure and 60 times of magnitude for type 2 structure over the conventional dipole design. This enhancement is attributed to inserting more tips in the nanoantenna design where a high divergence of surface current occurs.

 figure: Fig. 9

Fig. 9 Electric field distribution over dipole structures at λ=500nm and G=20nm for the (a) conventional dipole, (b) type 1, and (c) type 2 nanoantennas.

Download Full Size | PDF

4.4 Gap study for type 2 nanoantenna

For type 2 nanoantenna, a study of the effect of the gap size on different antenna parameters (such as input impedance, harvesting efficiency, and total current at the gap) was carried out. The input impedance of the nanoantenna, given in Eq. (6), consists of three parts: the loss resistance  Rloss,the radiation resistance  Rrad, and the reactive component X which represents the energy stored in the near field [31].

Z=Rrad+Rloss+jX
The loss resistance is primarily determined by the resistive losses of the antenna material with slight dependence on antenna geometry [32]. However, the radiation resistance depends mainly on the antenna geometry and the surface current distribution where the energy lost in radiation resistance is transformed to electromagnetic radiation [31,33]. Based on the dipole antenna theory, radiation resistance is directly proportional to the total electrical length of the antenna which is function of total length, diameter and gap size [30,34]. This is compatible with the results plotted in ‎Fig. 10 which shows small changes in the reactive part of impedance while the resistance is significantly increased, contributed to the increase of radiation and loss resistances.

 figure: Fig. 10

Fig. 10 Variation of real and imaginary parts of the antenna input impedance with the gap size.

Download Full Size | PDF

The total current induced at the antenna/rectifier interface represents the major factor in the rectenna system operation as it gives a measure of the DC power output from the rectenna device. In this analysis, the total current is calculated by integrating the electric surface current J over the rectifier port surface δs placed at the gap as illustrated in ‎Fig. 1. The values of the total current and harvesting efficiency for each gap size are plotted in ‎Fig. 11. It is evident from this figure that the total current has relatively small variation with the gap size change which indicates that the total field at the antenna port is almost constant. It is worth noting that the values of flux density (C/m2) and surface current (A/m2) at the gap decrease due to larger gap size. However, the total field enclosed in the gap region and the total current present at the rectifier port have relatively small changes in magnitudes. The harvesting efficiency can be expressed as the ratio of  Rradto the total input resistance [30]. Therefore, it’s expected that the harvesting efficiency will have relatively small variations, as evident from ‎Fig. 10, since  Rrad and  Rlosschange conjointly with the gap size as shown in ‎Fig. 10. It’s worth mentioning that the ripples in the numerical results, shown in Fig. 10 and Fig. 11, are due to internal resonances introduced when changing the gap size quite similar to Fabry-Perot interference pattern [35,36].

 figure: Fig. 11

Fig. 11 Variation of the harvesting efficiency and the total current at the nanoantenna port with the change of the gap size.

Download Full Size | PDF

4.5. Configurability of type 2 nanoantenna input impedance

Due to different surface current paths over type 2 nanoantenna, it acquires a multi impedance resonance behavior. This is clearly illustrated in ‎Fig. 12 where the input impedance for type 2 nanoantenna, with the previously studied dimensions, is plotted versus the free space wavelength. The figure shows four resonance points, indicated by the black dots, at different wavelengths with the corresponding resonance resistance values. Since this study targets an optimum operation at 500 nm wavelength, type 2 nanoantenna dimensions are tuned to match a 500 Ω impedance at resonance (indicated by point A). This impedance corresponds to the impedance of the fabricated MIM diode in [6].

 figure: Fig. 12

Fig. 12 Real and imaginary parts of the input impedance for type 2 nanoantenna with 500 Ω resonance impedance.

Download Full Size | PDF

As discussed above, the value of the input impedance at a single resonance point depends on the corresponding values for all resonance points over the antenna bandwidth. Moreover, type 2 nanoantenna can have up to 7 degrees of freedom in the design parameters as shown in ‎Fig. 1. Therefore, these parameters can be tuned to allow various resonant impedance values and hence, a wide range of impedance matching can be obtained at the desired operation wavelength. Figure 13 shows the input impedance for two optimized designs of type 2 nanoantenna which have resonant impedances of 1 KΩ (‎Fig. 13(a)) and 2 kΩ (‎Fig. 13(b)) at λ=500nmcorresponding to the fabricated MIM diode resistances reported in [11]. The optimized dimensions of type 2 designs are listed in ‎Table 3 which shows relatively small changes in the overall volume for the different structures. Figure 14 shows the harvesting efficiencies for type 2 nanoantennas with the different resonant impedances. It’s found that, at λ=500nm, optimizing the antenna design parameters to have higher resonance impedance results in lower harvesting efficiency.

 figure: Fig. 13

Fig. 13 Real and imaginary parts of the input impedance for optimized type 2 nanoantennas with resonance impedance of (a) 1 kΩ and (b) 2 kΩ.

Download Full Size | PDF

Tables Icon

Table 3. Type 2 Design Configurations for Different Input Impedance at λ = 500 nm

 figure: Fig. 14

Fig. 14 Variation of harvesting efficiency versus wavelength for type 2 nanoantennas with different resonant impedance.

Download Full Size | PDF

4.6. Fabrication tolerance analysis for type 2 nanoantenna

While modern techniques are available for fabricating nanoantenna devices, fabrication processes at nano scale always convey some perturbation. Therefore, a study of the fabrication tolerance is performed for type 2 nanoantenna structure. This study is directed to calculate the sensitivity of the nanoantenna performance while introducing minor changes to the structure dimensions (L1, W1,  L2, W2,  L3,  W3). It’s worth noting that the tolerance of a specific parameter is studied while other parameters of the design are held constant at their optimum values (L1=103.5 nm,   W1=51 nm,   L2=70 nm,  W2=19.5 nm,   L3=5 nm,   W3=13.4 nm). Figures 15 and 16 show the variation of the harvesting efficiency and the resonance impedance with the studied parameters W1andL1, respectively. Additionally, the summary of the tolerance study is listed in Table 4. It’s evident from the figures and Table 4 that the proposed design has a tolerance of ± 5% at which the harvesting efficiency is still higher than 53.8% with a maximum deviation of ± 58 Ω from the optimum input impedance of the nanoantenna. Additionally, a maximum wavelength resonance shift of 10 nm is obtained within the tolerance of ± 5% of the studied parameters. Therefore, it’s evident from the above results that the proposed nanoantenna design bears high robustness for fabrication imperfection.

 figure: Fig. 15

Fig. 15 Values of harvesting efficiency and resonance impedance with 5% variation of W1 parameter.

Download Full Size | PDF

 figure: Fig. 16

Fig. 16 Values of harvesting efficiency and resonance impedance with 5% variation of L1 parameter.

Download Full Size | PDF

Tables Icon

Table 4. Fabrication Tolerance for Type 2 Design Parameters at λ = 500 nm

5. Conclusion

In this paper, a full study for a novel tapered dipole design has been implemented where antenna input impedance, return loss, harvesting efficiency, surface current and field confinement were calculated using FEFD method. Simulation detailed results have demonstrated that the proposed nanoantenna structure can achieve a harvesting efficiency of 55.3% at wavelength of 500 nm and field confinement 60 times higher than the conventional dipole structure. Additionally, the configurability of type 2 nanoantenna resonant impedance has been proven and three different design configurations were proposed to match fabricated diode impedances.

References and links

1. R. L. Bailey, “A Proposed new concept for a solar-energy converter,” J. Eng. Power 94(2), 73–77 (1972). [CrossRef]  

2. D. K. Kotter, S. D. Novack, W. D. Slafer, and P. J. Pinhero, “Theory and manufacturing processes of solar nanoantenna electromagnetic collectors,” J. Sol. Energy Eng. 132(1), 011014 (2010). [CrossRef]  

3. S. Grover and G. Moddel, “Applicability of metal/insulator/metal (MIM) diodes to solar rectennas,” IEEE J. Photovolt. 1(1), 78–83 (2011). [CrossRef]  

4. G. A. E. Vandenbosch and Z. Ma, “Upper bounds for the solar energy harvesting efficiency of nano-antennas,” Nano Energy 1(3), 494–502 (2012). [CrossRef]  

5. Z. Ma and G. A. E. Vandenbosch, “Optimal solar energy harvesting efficiency of nano-rectenna systems,” Sol. Energy 88, 163–174 (2013). [CrossRef]  

6. M. N. Gadalla, M. Abdel-Rahman, and A. Shamim, “Design, optimization and fabrication of a 28.3 THz nano-rectenna for infrared detection and rectification,” Sci. Rep. 4, 4270 (2014). [CrossRef]   [PubMed]  

7. H. Fischer and O. J. F. Martin, “Engineering the optical response of plasmonic nanoantennas,” Opt. Express 16(12), 9144–9154 (2008). [CrossRef]   [PubMed]  

8. Z. Iluz and A. Boag, “Dual-Vivaldi wideband nanoantenna with high radiation efficiency over the infrared frequency band,” Opt. Lett. 36(15), 2773–2775 (2011). [CrossRef]   [PubMed]  

9. E. Briones, J. Alda, and F. J. González, “Conversion efficiency of broad-band rectennas for solar energy harvesting applications,” Opt. Express 21(S3Suppl 3), A412–A418 (2013). [CrossRef]   [PubMed]  

10. M. Hussein, N. F. Fahmy Areed, M. F. O. Hameed, and S. S. Obayya, “Design of flower-shaped dipole nano-antenna for energy harvesting,” IET Optoelectron. 8(4), 167–173 (2014). [CrossRef]  

11. J. L. Stokes, Y. Yu, Z. H. Yuan, J. R. Pugh, M. Lopez-Garcia, N. Ahmad, and M. J. Cryan, “Analysis and design of a cross dipole nanoantenna for fluorescence-sensing applications,” J. Opt. Soc. Am. B 31(2), 302–310 (2014). [CrossRef]  

12. S. Krishnan, “Design, fabrication and characterization of thin-film MIM diodes for rectenna array,” (University of South Florida, 2004).

13. F. F. K. Hussain, A. M. Heikal, M. F. O. Hameed, J. El-Azab, W. S. Abdelaziz, and S. S. A. Obayya, “Dispersion characteristics of asymmetric channel plasmon polariton waveguides,” IEEE J. Quantum Electron. 50(6), 474–482 (2014). [CrossRef]  

14. M. N. Gadalla, “Nano Antenna Integrated Diode (Rectenna) For Infrared Energy Harvesting,” (Kaust University, 2013).

15. S. Maci, G. B. Gentili, P. Piazzesi, and C. Salvador, “Dual-band slot-loaded patch antenna,” IEE Proc. Microw. Antenn. Propag. 142, 225–232 (1995) [CrossRef]  

16. H. R. Stuart, “Eigenmode analysis of small multi-resonant antennas,” in Proceeding of IEEE Antennas and Propagation Society International Symposium (IEEE, 2008), pp. 1–4.

17. M. Gustafsson and C. Sohl, “Summation rules for the antenna input impedance,” in Proceeding of IEEE Antennas and Propagation Society International Symposium (IEEE, 2008), pp. 1–4. [CrossRef]  

18. W. Hu, K. Sarveswaran, M. Lieberman, and G. H. Bernstein, “Sub-10 nm electron beam lithography using cold development of poly (methyl methacrylate),” J. Vac. Sci. Technol. B 22(4), 1711–1716 (2004). [CrossRef]  

19. J.-T. Lv, Y. Yan, W.-K. Zhang, Y.-H. Liu, Z.-Y. Jiang, and G.-Y. Si, “Plasmonic nanoantennae fabricated by focused Ion beam milling,” Int. J. Precis. Eng. Manuf. 16(4), 851–855 (2015). [CrossRef]  

20. C. J. Lo, T. Aref, and A. Bezryadin, “Fabrication of symmetric sub-5 nm nanopores using focused ion and electron beams,” Nanotechnology 17(13), 3264–3267 (2006). [CrossRef]  

21. P. F. A. Alkemade and E. van Veldhoven, “Deposition, Milling, and Etching with a Focused Helium Ion Beam,” in Nanofabrication: Techniques and Principles, M. Stepanova and S. Dew, eds. (Springer, 2012), pp. 275–300.

22. O. Muskens, Y. Wang, M. Abb, S. Boden, C. H. Groot, and J. Aizpurua, “Helium ion beam milling for plasmonic nanoantennas,” SPIE Newsroom 1407, 5553 (2014). [CrossRef]  

23. Comsol Multiphysics software.

24. B. M. Rahman, D. M. Leung, S. S. Obayya, and K. T. Grattan, “Numerical analysis of bent waveguides: bending loss, transmission loss, mode coupling, and polarization coupling,” Appl. Opt. 47(16), 2961–2970 (2008). [CrossRef]   [PubMed]  

25. M. Rajarajan, S. Obayya, B. Rahman, K. Grattan, and H. El-Mikali, “Characterization of low-loss waveguide bends with offset-optimization for compact photonic integrated circuits,” IEE P-Optoelecton. 147(6), 382–388 (2000). [CrossRef]  

26. E. A. Soliman, M. O. Sallam, and G. A. E. Vandenbosch, “Plasmonic grid array of gold nanorods for point-to-point optical communications,” J. Lightwave Technol. 32(24), 4898–4904 (2014). [CrossRef]  

27. P. B. Johnson and R. W. Christy, “Optical constants of noble metals,” Phys. Rev. B 6(12), 4370–4379 (1972). [CrossRef]  

28. J. Kennedy and R. Eberhart, “Particle swarm optimization,” in Proceedings of IEEE International Conference on Neural Networks (IEEE, 1995), pp. 1942–1948. [CrossRef]  

29. E. D. Palik, Handbook of Optical Constants of Solids (Academic Press, 1985).

30. A. E. Alù and N. Engheta, “Input impedance, nanocircuit loading, and radiation tuning of optical nanoantennas,” Phys. Rev. Lett. 101(4), 043901 (2008). [CrossRef]  

31. C. A. Balanis, Antenna Theory (John Wiley & Sons, 1997).

32. P. Banerjee and T. Bezboruah, “Theoretical study of radiation characteristics of short dipole antenna,” in Proceedings of The International MultiConference of Engineers and Computer Scientists, (2014), pp. 785–790.

33. J.-J. Greffet and F. Marquier, “Impedance of a nanoantenna and a quantum emitter,” in Frontiers in Optics 2011/Laser Science XXVII, OSA Technical Digest (Optical Society of America, 2011), paper LWG4.

34. David Staelin, 6.661 Receivers, Antennas, and Signals, spring2003, (MIT OpenCourseWare), http://ocw.mit.edu (Accessed 1 Mar, 2016).

35. E. Cubukcu and F. Capasso, “Optical nanorod antennas as dispersive one-dimensional Fabry–Pérot resonators for surface plasmons,” Appl. Phys. Lett. 95(20), 201101 (2009). [CrossRef]  

36. H. Liu, M. Erouel, E. Gerelli, A. Harouri, T. Benyattou, R. Orobtchouk, L. Milord, A. Belarouci, X. Letartre, and C. Jamois, “Nanoantenna-induced fringe splitting of Fabry-Perot interferometer: a model study of plasmonic/photonic coupling,” Opt. Express 23(24), 31085–31097 (2015). [CrossRef]   [PubMed]  

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 (16)

Fig. 1
Fig. 1 (a) Conventional dipole [‎4], (b) Type 1 (2-steps tapered dipole), (c) Type 2 (3-steps tapered dipole).
Fig. 2
Fig. 2 Surface current distribution over nanoantenna structures at 500 nm wavelength (the bigger arrows indicates higher divergence of current) for (a) conventional dipole, (b) type 1, and (c) type 2.
Fig. 3
Fig. 3 Wavelength dependent harvesting efficiency for conventional dipole by the FDTD [4] and FEFD method.
Fig. 4
Fig. 4 Wavelength dependent harvesting efficiency for the conventional dipole at different minimum mesh element sizes.
Fig. 5
Fig. 5 Wavelength dependent harvesting efficiency for the conventional dipole using different permittivities based on experimental data [4], Johnson and Christy model [27], Palik model [‎29] along with that reported by Vandenbosch and Ma [4].
Fig. 6
Fig. 6 Fitness function (Eq. (2) values for type 2 nanoantenna versus number of iterations of the PSO algorithm. The corresponding values of S 11 and η rad are demonstrated for 3 points.
Fig. 7
Fig. 7 Variation of harvesting efficiency versus wavelength for conventional dipole [‎‎4] and the suggested designs of type 1 and type 2 nanoantennas.
Fig. 8
Fig. 8 Wavelength dependent return loss at the nanoantenna port for conventional dipole [‎‎4] and the proposed designs of type 1 and type 2 nanoantennas.
Fig. 9
Fig. 9 Electric field distribution over dipole structures at λ = 500 nm and G=20 nm for the (a) conventional dipole, (b) type 1, and (c) type 2 nanoantennas.
Fig. 10
Fig. 10 Variation of real and imaginary parts of the antenna input impedance with the gap size.
Fig. 11
Fig. 11 Variation of the harvesting efficiency and the total current at the nanoantenna port with the change of the gap size.
Fig. 12
Fig. 12 Real and imaginary parts of the input impedance for type 2 nanoantenna with 500 Ω resonance impedance.
Fig. 13
Fig. 13 Real and imaginary parts of the input impedance for optimized type 2 nanoantennas with resonance impedance of (a) 1 kΩ and (b) 2 kΩ.
Fig. 14
Fig. 14 Variation of harvesting efficiency versus wavelength for type 2 nanoantennas with different resonant impedance.
Fig. 15
Fig. 15 Values of harvesting efficiency and resonance impedance with 5% variation of W 1 parameter.
Fig. 16
Fig. 16 Values of harvesting efficiency and resonance impedance with 5% variation of L 1 parameter.

Tables (4)

Tables Icon

Table 1 Radiated Power and Power Loss Equations Extracted from the Simulated Electric and Magnetic Fields [‎23,‎‎31]

Tables Icon

Table 2 Dipole Designs Dimensions in Nanometer Introduced to the FEFD analyzer

Tables Icon

Table 3 Type 2 Design Configurations for Different Input Impedance at λ = 500 nm

Tables Icon

Table 4 Fabrication Tolerance for Type 2 Design Parameters at λ = 500 nm

Equations (7)

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

η rad = P rad P in = P rad P rad + P loss
f=  c 1 × η rad + c 2 ×| S 11 |
W i >  W i+1
W i   W i+1 >5 nm
n L i × W i  const ,    n=1,2,3
η total = 0 P( λ,T )×  η rad ( λ ) dλ 0 P( λ,T ) dλ
Z= R rad + R loss +jX
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.