* Your assessment is very important for improving the workof artificial intelligence, which forms the content of this project
Download Field Computational Aspects of Wireless Power Transfer
Mechanical filter wikipedia , lookup
Chirp spectrum wikipedia , lookup
Opto-isolator wikipedia , lookup
Switched-mode power supply wikipedia , lookup
Electrification wikipedia , lookup
Loudspeaker wikipedia , lookup
Power engineering wikipedia , lookup
Mains electricity wikipedia , lookup
Distribution management system wikipedia , lookup
Rectiverter wikipedia , lookup
Lumped element model wikipedia , lookup
Distributed element filter wikipedia , lookup
Nominal impedance wikipedia , lookup
Electromagnetic compatibility wikipedia , lookup
Loading coil wikipedia , lookup
Utility frequency wikipedia , lookup
Mathematics of radio engineering wikipedia , lookup
Alternating current wikipedia , lookup
Galvanometer wikipedia , lookup
Zobel network wikipedia , lookup
RLC circuit wikipedia , lookup
Network analysis (electrical circuits) wikipedia , lookup
Two-port network wikipedia , lookup
Field Computational Aspects of Wireless Power Transfer Szabolcs Gyimóthy1 , Sándor Bilicz1 , József Pávó1 , László Tóth1 , Péter Kis2 , Gábor Varga1 and László Szűcs1 1 Budapest University of Technology and Economics, Egry J. u. 18, H-1111 Budapest, Hungary 2 Furukawa Electric Institute of Technology, Késmárk u. 28/A, H-1158 Budapest, Hungary E-mail: [email protected] Abstract—The aim of this paper is, on one hand, to give an overview of the fast developing and diversified field of wireless power transfer (WPT) from the point of view of the inevitable numerical modeling of electromagnetic fields. On the other hand, we propose a fast method by which the full WPT system based on inductive resonant coupling (IRC) can be analyzed or even optimized very efficiently. The proposed method has been validated by alternative methods and measurements. Index Terms—wireless power transfer, inductive resonant coupling, method of moments I. I NTRODUCTION No doubt, the revolution of wireless power transfer is facing us. Electric vehicles, mobile devices and several industrial and medical applications are raring to use this new feature. Among various techniques making such energy-exchange possible, the inductive resonant coupling [1] is preferred for the time being. Although the capabilities of alternative techniques are far from being exhaustively researched [2], yet in some cases the “big idea” is waited for. Apparently Tesla’s original method (1894) for wireless energy transfer based on inductive resonant coupling was kind of reinvented around 2007, in fact it had never been really forgotten (cf. Fig. 1). Although the technique of IRC WPT has become more or less mature for the mass application, there is still a lot of work to do in order to improve factors like efficiency, working distance, transmitted power and safety. Moreover, standardization, regulations and guidelines of the field are still incomplete. As a matter of fact, the numerical simulation of electromagnetic (EM) fields has currently growing interests in this development. Besides its use for component (e.g. coil) design [4], issues like adaptive coupling [5], foreign object detection [6] and human exposure [7] have to be dealt with, too. Hot topics are among others multifrequency coupling [8], meta-materials for enhancing mutual coupling of coils [9], repeaters for relaying power to larger distances [10] and resonant reactive shield- Figure 1. Illustration of an experiment for inductive resonant wireless power transfer as can be found in a Hungarian schoolbook [3] published in 1954. ing [11]. Some novel applications requiring sophisticated EM simulation are “charge while driving” systems [12] and board-to-board level interconnection in highly integrated packages [13]. In this paper we first review the most prevalent modeling approaches of resonant WPT and the role of EM simulation in each of them. Then we propose a lightweight computational method based on integral equations, by which the full IRC WPT system can be simulated and optimized very efficiently. On the one part, the results computed for a simple configuration are verified with those obtained by the finite element method (FEM) and compared with measured data as well. On the other part, it is shown that full-wave modeling of the whole WPT system, which is achievable by the proposed method, can better predict the conditions of optimal power transfer than other methods based on component analysis and equivalent circuits. II. T RANSFER MODELS AND FIELD SIMULATION The technology of wireless non-radiative mid-range energy transfer is based on the so-called long-lived oscillatory resonant electromagnetic modes with localized slowly-evanescent field patterns in the regime of “strong coupling”. This is something peculiar from the point of view of computational electromagnetism in the sense that quite different approaches —e.g. static analysis, wave propagation or mode analysis— can provide useful results. At the same time one or two methods may neglect or ignore important features. For example, static and quasi-static analyses certainly do not account for the radiated losses. Indeed, when using inductive coupling, these losses are almost negligible from the point of view of transfer efficiency. On the other hand, this radiation phenomenon may contribute to interferences and human exposure, which can be important factors too. Yet in practice the preferred type of EM analysis is primarily determined by the mathematical model used for describing the energy transfer (see below). Zs Vs Figure 2. I1 V1 I2 s+1 s-1 Two-port Sik or Zik s+2 V s-2 2 Zl Simplified two-port representation of the WPT chain. A. Two-port representation From an electrical engineer’s point of view the most general description of power transfer can be given by a two-port (see Fig. 2), which is characterized either with impedance parameters Zik or with scattering parameters Sik , all complex and frequency dependent. Note that the full WPT chain is normally set up with sophisticated electronics on both the transmitter (Tx) and the receiver (Rx) side, e.g. converters, filters, matching circuits, as well as elements for control and communication. In our model the “two-port box” contains merely the “field part” i.e. the two resonators (most often capacitively loaded coils) together with some coupling elements. At the same time the circuit environment is greatly simplified for facilitating field analysis: the system is fed by a generator of source voltage Vs and intrinsic impedance Zs = Rs + jXs , whereas the load is modeled as an adjustable impedance Zl = Rl + jXl . Adequate computation of the frequency characteristics of the two-port in a broad frequency range requires full-wave EM analysis, which is expensive as a rule. Some attempts have already been made for reducing the computational burden of FEM simulation, by means of e.g. homogenization [4][14], domain decomposition[15] and “multi-slicing” [16]. B. Coupled mode theory (CMT) CMT describes the behavior of two (or more) resonant objects having (about) the same resonance frequency and being in weak coupling with each other. It is based on perturbation theory and is thoroughly described in e.g. [17]. The governing equations of a two-resonator system are as follows: da1 = jω1 a1 (t) − (Γext,1 + Γ1 )a1 (t) + jκa2 (t) dt p + 2Γext,1 s+1 (t) da2 = jω2 a2 (t) − (Γext,2 + Γ2 )a2 (t) + jκa1 (t) dt p s−1 (t) = 2Γext,1 a1 (t) − s+1 (t) p s−2 (t) = 2Γext,2 a2 (t) (1) in which a1,2 are so-called complex mode amplitudes, the squared magnitude of which gives the total energy stored in the resonator; s±1,2 are incident and reflected wave amplitudes, the square of which give the respective powers. The system is characterized by the self-resonant frequencies ω1,2 , the coupling coefficient κ, as well as the internal and external losses Γ1,2 and Γext,1,2 , respectively. As referring to the scheme of Fig. 2, s+1 and Γext,1 together represent the generator, while Γext,2 represents the load. Equations (1) are not confined to electromagnetics — they can also describe coupled pendulums, for example— thus CMT is very general in this respect. On the other hand, its validity is limited to weak coupling. Moreover, while CMT explains the transient flow of energy between resonators very plausibly, in the steady state it provides merely a transfer function, which is valid only in the vicinity of the resonance frequency indeed. CMT parameters of a coupled resonant electromagnetic system can be best obtained by means of mode analysis. The coupling coefficient κ is simply equal to the difference between the eigenfrequencies of the two fundamental modes of the coupled system. In turn, the attenuation coefficients Γ can be determined from the respective quality factors of the resonators, as Q = ω/2Γ holds (for more details see [18]). C. Lumped circuit representation For a designer, lumped circuit is undoubtedly the readiest model of WPT. Moreover, an electrical engineer believes he/she can intuitively assign circuit elements to EM field phenomena in most cases. Note that — unlike when using CMT model— we can have different circuit models depending on the coupling phenomena and the actual resonator configuration (cf. circuit model in Section V). Lumped parameters are commonly extracted from static or quasi-static analysis, and sometimes from analytical formulas [19]. Self and mutual inductances L, M are usually obtained from static magnetic analysis. Wire resistance Rwire at high frequency can be approximated analytically by considering the skin effect (but ignoring the proximity effect). Radiation losses are represented by Rrad and may be approximated analytically, too, by modeling the coil as a current dipole. Finally, self capacitance of the coil C is usually negligible to that of the external capacitor Cext connected to the coil, but if not, still an approximation can be given by assuming a kind of sinusoidal current distribution along the coil [1]. Although widespread, this practice can be criticized in many respects. Above all, once having obtained the (approximate) lumped parameters, one is tempted to treat the equivalent circuit as the primary model. For example, parameters of the CMT model are usually derived from one such circuit model — which method is indeed adequate, provided the lumped parameters had been calculated at the resonance frequency. Far less acceptable however, if one derives the impedance or scattering parameters of the circuit as a two-port, and applies it in a broader frequency range (see Section V-B for a comparison). Even if the frequency dependence of lumped parameters is taken into account (e.g. R = R(ω)), it is the network structure that cannot reflect all possible couplings, some of which might be significant under unforeseen or unexamined circumstances. III. I NTEGRAL FORMULATION Computing “frequency scan” within a wide band, performing parameter studies, or optimizing WPT systems, all of these call for an efficient numerical method. A fastto-evaluate integral formalism can be proposed, in case the following conditions hold: 1) The system is made up of air-cored coils, optionally loaded with elements that can be considered lumped. 2) Each coil is made of a homogeneous and thin wire (compared to the radius of the turns) with circular cross-section. 3) The coil configuration is surrounded by air. Note that the extension of the method for handling conducting extraneous objects in the proximity is straightforward, but is not dealt with hereafter. The numerical simulation is based on the discretization of the wire, resulting in a much smaller number of degrees of freedom (DoF) than using FEM. The method is similar to the classical scheme of wire-antenna calculation based on the method of moments (MoM). However, it goes beyond that by exploiting the special geometry of coils. In addition to this, a quasi-static approximation —to be used for the extraction of lumped parameters, for instance— can be given in an elegant way. The equations are summarized below for only one single coil in order to keep simplicity, but they can be easily generalized for an arbitrary number of coils (for more details see [20]). In the proposed model, the wire of each coil is represented by a 1-dimensional curve, along which we define the coordinate ζ, so that ζ = 0 corresponds to the beginning, and ζ = l to the end of the wire, respectively. Sinusoidal time dependence with angular frequency ω is assumed for all quantities (the exp(jω) convention is used). The magnetic vector potential A and the electric scalar potential Φ along the wire are expressed with the retarded source quantities, i.e. current I and charge per unit length q, respectively, as follows: Z µ0 l I(ζ ′ )dζ ′ −jβχ(ζ,ζ ′ ) Aζ (ζ) = êζ e (2) 4π 0 χ(ζ, ζ ′ ) Z l 1 q(ζ ′ )dζ ′ −jβχ(ζ,ζ ′ ) Φ(ζ) = e (3) 4πε0 0 χ(ζ, ζ ′ ) where β = ω/c, and êζ denotes the unit vector tangential to the wire (note that we address only the ζcomponent of A). Then we apply the differential Ohm’s law, Eζ = rI(ζ) to express the tangential electric field on the surface of the conductor, where r is the resistance per unit length of the wire. Hence using the potentials one can write dΦ − jωAζ (4) rI(ζ) = − dζ Figure 4. The studied helical coil. additional relationships have to be given for the terminals of each coil. In our simulations, simply the current is prescribed, i.e. I(0) = I(l) = Icoil , but this is also the point where more complex lumped port boundary conditions can be entered. Once the system of equations is solved (cf. [20] for the details of discretization and numerical implementation) one can compute the complex amplitude of voltage between the coil terminals —which makes sense only if their separation is small— as Vcoil = Φ(l) − Φ(0), and then find the impedance. Alternatively, the complex power flowing through the coil terminals can be expressed. A notable advantage of the formulation from the point of view of EMC studies is that the radiated part of complex power can be easily expressed by the Poynting’s theorem: Z l 1 [Aζ (ζ)I ∗ (ζ) − q(ζ)Φ∗ (ζ)] dζ (6) Smedium = jω 2 0 in which superscript (*) stands for the complex conjugate. IV. A NALYSIS OF A SINGLE COIL For evaluating the above introduced numerical method, we have examined the impedance of a helical coil made of copper, in the frequency range 10 . . . 100 MHz. The radius of the coil is 110 mm, its height is 24.5 mm, the wire diameter is 1.5 mm and the number of turns is 6 (see Fig. 4). We used an Agilent/HP 4191A RF impedance analyzer for the measurement. The simulation of the impedance characteristics was carried out both by the integral formulation and by FEM — for the latter the RF Module of Comsol Multiphysics was used [21]. The performance of the two solutions are compared in Table I. For verification we have also computed the selfresonance frequencies of the coil by Comsol RF, using Finally, charge conservation law can be written as jωq(ζ) + dI(ζ) = 0. dζ (5) The two integral and two differential equations (2)-(5) form a coupled system for the unknown current, charge, and potential distributions. To provide a unique solution, Table I P ERFORMANCE OF NUMERICAL SIMULATIONS Degrees of freedom CPU time for 100 pts. [sec] FEM 3055664 18273 Integral formulation 1200 13 6 10 integral formulation FEM sweep FEM eigenfrequency measurement 5 impedance magnitude (kΩ) 10 4 10 X: 52 Y: 2704 3 10 2 10 1 10 10 30 frequency (MHz) its eigenfrequency application mode. Note that when considering open ended coil one obtains the resonance frequencies at which the impedance has a local maximum. In turn, with short circuited coil terminals one gets the so-called anti-resonance frequencies where the impedance tends to zero. All the simulated and measured results are plotted in Fig. 3. As the figure shows, the two simulated impedance curves agree well, and the computed resonance frequencies coincide with their extrema, too. However, the measured curve shows a systematic frequency shift from the computed ones in terms of the peaks. This may be originated either from calibration — although the measurement was carried out carefully enough— or from the glue material used to fasten the windings. More interestingly, the measured impedance curve shows a double peak at about 50 MHz, but this does not appear at all in either of the simulated curves. At the same time, FEM eigenfrequency analysis predicts both resonance and anti-resonance at about 58 MHz — a good example of how different types of analysis can catch or dismiss certain features. This resonance corresponds to a specific mode in which both the current and the voltage at the terminals are zero. We suspected that for a symmetric coil this singular behavior is confined to a such narrow frequency range, that with the chosen frequency steps we simply could not catch it. On the other hand, we believed this range is widened 0.02 0.02 0.015 0.015 z (m) z (m) 40 50 60 70 80 90 100 Comparison of the simulated and measured impedance curves, as well as the computed resonance frequencies of the helical coil. 0.01 0.01 and the effect becomes stronger, if the perfect symmetry of the coil is broken, as it can always happen with handmade coils. Therefore we modeled a coil having the same overall dimensions but a variable pitch of winding (see Fig. 5). The comparison of its impedance characteristics with that of the original coil (having uniform winding) clearly shows the difference: the expected double peak appears around 58 MHz (see Fig 6), while the remaining parts of the characteristics are not affected. V. A NALYSIS OF A WPT SYSTEM In order to point out the necessity of the full-wave field analysis of a WPT system, instead of resorting to its lumped circuit model, we have compared the efficiency predicted by the two models w.r.t. frequency and resonator distance. The configuration was adopted from [1]. It contains two identical coils of 30 cm radius and 20 cm height. There is a planar loop of 25 cm radius on the side of each resonator for the inductive coupling of the source and the load, respectively. All wires are made of copper having 6 mm diameter. The axes of the components are aligned (see Fig. 7). The distance range [0.5, 2.5]m and the frequency range [8, 11]MHz were examined in the simulations. graded winding regular winding impedance magnitude (kΩ) Figure 3. 20 100 10 X: 58.55 Y: 0.7432 1 0.1 0.005 0.005 0.01 0 −0.1 −0.05 0 x (m) 0.05 0.1 0 −0.1 −0.05 0 x (m) 0.05 0.1 Figure 5. Coil models of uniform and “graded” winding. Note that the two axes have different scaling for better visualization. 10 Figure 6. 20 30 40 frequency (MHz) 50 60 70 100 Impedance characteristics of the coils shown in Fig 5. 1 z (m) 1.8 1.6 1.4 1.2 1 0.8 0.6 0.4 0.2 0 0.2 y (m) 0.8 0 −0.2 efficiency 0.7 −0.2 0 Figure 7. 0.2 x (m) fullwave 1m fullwave 2m lumped 1m lumped 2m 0.9 0.6 0.5 0.4 0.3 Configuration of the two-resonator WPT system. 0.2 0.1 For the full-wave frequency-domain modeling of the system we have used the method introduced in Section III. The lumped circuit model of the system is depicted in Fig. 8. Inductance (self and mutual) and capacitance parameters of the circuit were computed by the quasi-static approximation of the above mentioned method (cf. [20]). The resistance values representing joule loss and radiation, respectively, were calculated with analytical formulas as described in Section II-C. Both the full-wave model and the circuit model are represented by the impedance matrix of the two-port (symmetric, in this case) the entries of which, Z11 = R11 + jX11 and Z12 = R12 + jX12 , are frequency dependent. Note however, that this dependence differs for the two models. 0 8 8.5 9 9.5 10 frequency (MHz) 10.5 11 Figure 9. Maximum efficiency w.r.t. operating frequency at two fixed distances, as predicted by the two models. (The full-wave model suffers from numerical instabilities at the certain frequencies; some outlying data points were skipped.) decrease efficiency). Hence the three real parameters Rl , Xs and Xl are to be chosen such that η is maximal (cf. Fig. 2). Their “matched” values depend on the frequency and the coil distance as well. It can be shown that the maximum efficiency does not depend on Xs ; it can be used to compensate the input reactance at the primary port so as to the maximal power can enter the twoport at a given source voltage Vs . Thus, the optimization problem reduces to A. Impedance matching for optimal efficiency The conditions of optimal efficiency of WPT are usually determined with the assumption that the output power of the generator is maximized. This leads to the so-called double conjugate matching conditions [22]. At the same time, the internal losses of the generator are not dealt with. Since in our opinion these losses should be considered in the evaluation of the whole system, we define efficiency in a slightly different manner. We consider the net power coming from the source, 1 Re{Vs I1∗ } 2 and the power consumed by the load, Ps = (7) 1 1 (8) Pl = − Re{V2 I2∗ } = Rl |I2 |2 . 2 2 The efficiency of power transfer is defined as η = Pl /Ps . For the optimization of this efficiency we assume that the resistance of the source Rs is given (a lower bound exists in practice, and an artificial increase of Rs can only Ml-c Lloop V1 Rloop I1 Figure 8. Mc-c Lcoil Rcoil Ccoil Ml-c Lcoil Rcoil Ccoil Lloop Rloop V2 I2 Lumped circuit model of the WPT system shown in Fig. 7 (R̂l , X̂l ) = arg max η(Rl , Xl |Rs , Z11 , Z12 ) (9) The optimal choice of Rl and Xl is as follows (deduced simply by calculating the partial derivatives; details are omitted herein) p 2 − R2 + R R )(R2 + X 2 + R R ) (R11 s 11 s 11 12 11 12 R̂l = R11 + Rs R12 X12 − (R11 + Rs )X11 X̂l = . (10) R11 + Rs B. Results and comparison Figure 9 shows how much the maximum achievable efficiency depends on the frequency. Note that the two models suggest different operating frequencies (9.0 MHz by lumped, and 10.3 MHz by full-wave). At the same time, the distance does not have considerable effect on the location of the best operating frequency. In Fig. 10 the dependence of the maximum available efficiency on the coil distance is presented, provided that the best operating frequency has been chosen for each model (cf. Fig 9). In this figure, the measured data of [1] are also indicated, although the authors defined efficiency in the other way (see above). The predictions of the two models are different, again. Finally, the wide variation of the optimal load resistance is presented in Figure 11 as function of the distance (9.0 MHz frequency is set for the lumped model, and 10.3 MHz for the full-wave model). This is definitely an 1 R EFERENCES 0.9 0.8 efficiency 0.7 0.6 0.5 0.4 0.3 0.2 0.1 0 0.5 fullwave lumped measured* 1 1.5 distance (m) 2 2.5 Figure 10. Maximum efficiency w.r.t. coil distance predicted by the two models, and compared with measured data borrowed from [1]. 1 optimal load resistance (kΩ) 0.9 0.8 0.7 0.6 0.5 0.4 0.3 fullwave lumped 0.2 0.1 0 0.5 1 1.5 distance (m) 2 2.5 Figure 11. Optimal load resistance w.r.t. coil distance, predicted by the two models issue for the design of the matching electronic circuit, which connects the receiver loop and the device to be supplied, in order that the loop is loaded with the optimal impedance. However, model predictions are slightly different. VI. C ONCLUSIONS We believe that different transfer models and field analysis methods can be applied complementary to have a better insight to the mechanism and characteristics of WPT. At the same time, we underline the superiority of the fullwave field computational model over the lumped circuit model. In this paper, at least the remarkable difference in their behavior and predictions was demonstrated. We have also proposed a fast numerical field simulation tool based on the classical method of moments by which the whole inductive resonant WPT system can be efficiently analyzed. The method has been verified by alternative methods and measurements; its extension to handling extraneous objects is underway. ACKNOWLEDGMENT This work was supported by the Hungarian Scientific Research Fund under grants K-105996 and K-111987. [1] A. Kurs, A. Karalis, R. Moffatt et al., “Wireless power transfer via strongly coupled magnetic resonances,” Science, vol. 317, no. 5834, pp. 83–86, July 2007. [2] M. Masuda, M. Kusunoki, D. Obara et al., “Wireless power transfer via electric coupling,” Furukawa Review, vol. 44, pp. 33– 38, 2013. [3] J. Öveges, Kis fizika II. Művelt Nép Könyvkiadó, 1954, (in Hungarian). [4] D. Klis, S. Burgard, O. Farle, and R. Dyczij-Edlinger, “A reducedorder fe method for wireless power transfer systems using homogenized litz wires,” in IGTE Symposium, Graz, Austria, Sept 2014, pp. P1–8. [5] W. Chen, S. Rickers, Z. Bai, G. H.Bruck, and P. Jung, “Design of spatial transmitter for freely positioned wireless power transfer,” in IEEE CEFC, Annecy, France, May 2014, pp. PB4–18. [6] T. Yamamoto, T. Ota, and K. Hirata, “Foreign object detection with sensing magnetic field variation for wireless power transfer,” in IEEE CEFC, Annecy, France, May 2014, pp. PE1–20. [7] P.-P. Ding, L. Bernard, L. Pichon, and A. Razek, “Evaluation of electromagnetic fields in human body exposed to wireless inductive charging system,” IEEE Transactions on Magnetics, vol. 50, no. 2, pp. 1037–1040, 2014. [8] M.-L. Kung and K.-H. Lin, “Investigation of dual-band coil module for near-field wireless power transfer systems,” in IEEE WPTC, May 2014, pp. 265–268. [9] G. Lipworth, J. Ensworth, K. Seetharam et al., “Magnetic metamaterial superlens for increased range wireless power transfer,” Scientific Reports, vol. 4, no. 3642, 2014. [10] I. Awai, Y. Ikuta, Y. Sawahara, Y. Thang, and T. Ishizaki, “Applicaions of a novel disk repeater,” in IEEE WPTC, May 2014, pp. 114–117. [11] H. Moon, S. Ahn, and Y. Chun, “Design of a novel resonant reactive shield for wireless charging system in electric vehicle,” in IEEE WPTC, May 2014, pp. 220–223. [12] D. Bavastro, A. Canova, V. Cirimele et al., “Design of wireless power transmission for a charge while driving system,” IEEE Transactions on Magnetics, vol. 50, no. 2, pp. 965–968, 2014. [13] S. Kim, D. H. Jung, J. J. Kim et al., “Modeling of electromagnetic interference shielding materials in wireless power transfer for board-to-board level interconnections,” in IEEE WPTC, May 2014, pp. 273–276. [14] D. Ferreira, R. Sabariego, L. Lebensztajn et al., “Homogenization methods in simulations of transcutaneous energy transmitters,” IEEE Transactions on Magnetics, vol. 50, no. 2, pp. 1017–1020, 2014. [15] M. Benhamouche, L. Bernard, L. Pichon, and M. Ibrahim, “Domain decomposition method for solving complex magnetodynamic problem: application to wireless charging system,” in IEEE CEFC, Annecy, France, May 2014, pp. PB2–2. [16] X. Zhang, H. L. Li, S. L. Ho, and W. N. Fu, “A multi-slice finite element model including distributive capacitances for wireless magnetic resonant energy transfer systems with circular coils,” IEEE Transactions on Magnetics, vol. 49, no. 5, pp. 1857–1860, 2013. [17] H. A. Haus, Waves and fields in optoelectronics. Prentice Hall, 1984. [18] A. Karalis, J. Joannopoulos, and M. Soljačić, “Efficient wireless non-radiative mid-range energy transfer,” Annals of Physics, vol. 323, no. 1, pp. 34–48, 2008. [19] A. Bodrov and S.-K. Sul, “Analysis of wireless power transfer by coupled mode theory (CMT) and practical considerations to increase power transfer efficiency,” in Wireless Power Transfer – Principles and Engineering Explorations, K. Y. Kim, Ed. InTech, 2012, pp. 19–50. [20] S. Bilicz, “High-frequency modeling of coils by integral formulations,” in IGTE Symposium, Graz, Austria, Sept 2014, pp. P1–12. [21] Introductory tutorial to the RF Module: Coil design, COMSOL, 2009, available online. [22] T. Ohira, “Angular expression of maximum power transfer efficiency in reciprocal two-port systems,” in IEEE WPTC, 2014, pp. 228–230.