1: \documentstyle[aps,preprint]{revtex}
2: %\documentclass[prl]{revtex4}
3: %\documentstyle{article}
4: \begin{document}
5: \draft
6: \title{Image resolution depending on slab thickness and object distance in a two-dimensional photonic-crystal-based superlens}
7: \author{Xiangdong Zhang}
8: \address{Department of Physics, Beijing Normal University, Beijing
9: 100875, P. R. China}
10:
11: \maketitle
12:
13: \begin{abstract}
14: Based on the exact numerical simulation and physical analysis, we
15: have demonstrated all-angle single-beam left-handed behavior and
16: superlens for both TE and TM modes in a two-dimensional coated
17: photonic crystals. The imaging behaviors by two-dimensional
18: photonic-crystal-based superlens have been investigated
19: systematically. Good-quality images and focusing, with relative
20: refractive index of -1, have been observed in these systems for
21: both polarized waves. In contrast to the images in near-field
22: region for the lowest valence band, non-near-field images,
23: explicitly following the well-known wave-beam negative refraction
24: law, have been demonstrated. The absorption and compensation for
25: the losses by introducing optical gain in these systems have also
26: been discussed. Thus, extensive applications of such a phenomenon
27: to optical devices are anticipated.
28:
29: \end{abstract}
30: \pacs{PACS numbers: 78.20.Ci, 42.70.Qs, 41.20.Jb
31: }
32: \narrowtext
33:
34: \section{INTRODUCTION}
35:
36: Recently there has been a great deal of interest in studying a
37: novel class of media that has become known as the left-handed
38: materials (LHMs)[1-20]. These materials are characterized by
39: simultaneous negative permittivity $\epsilon$ and negative
40: permeability $\mu$. Properties of such materials were analyzed
41: theoretically by Veselago over 30 years ago[1]. As was shown by
42: veselago, the LHMs possess some peculiar electromagnetic
43: properties such as inverse Snell's law, reversed Doppler shift,
44: and reversed Cherenkov radiation. It had also been suggested that
45: slab of the LHM could be employed as an unconventional flat lens.
46:
47: Due to the absence of naturally occurring materials having both
48: negative $\epsilon$ and negative $\mu$, Veselago's prediction did
49: not receive much attention until recently, when a system
50: consisting an array of split-ring resonators and metallic wires
51: was prepared and demonstrated to have negative refractive index
52: experimentally [3,4]. Subsequently, some physical properties of
53: the LHMs were analyzed by many authors [5-20]. Pendry [5]
54: predicted that the LHM slab can amplify the evanescent waves and
55: the flat lens constructed from such a material with
56: $\epsilon=\mu=-1$ could in principle work as ``perfect" lens
57: (superlens). Although the concept of superlens was questioned by a
58: number of authors [21], more detailed physical considerations had
59: shown that the construction of ``almost perfect" lens is indeed
60: possible [22-27]. Recently, such image behaviors have been
61: observed by some numerical simulations [22-24] and experimental
62: measurements [24-27]. However, only near-field images were
63: demonstrated and extensive applications of such a phenomenon were
64: limited [22-27].
65:
66: It was shown that the negative refraction could also occur in
67: photonic crystal (PC)[28-39]. The physical principles that allow
68: negative refraction in them arise from the dispersion
69: characteristics of wave propagation in a periodic medium, which
70: can be well described by analyzing the equifrequency surface (EFS)
71: of the band structures [28-37]. In the PC structures, there are
72: two kinds of cases for negative refraction occuring [31]. The
73: first is the left-handed behavior as being described above[29-32].
74: In this case, $\vec{k}$, $\vec{E}$ and $\vec{H}$ form a
75: left-handed set of vectors (i.e., $\vec{S}\cdot\vec{k}<0$, where
76: $\vec{S}$ is the Poynting vectors). Another case is that the
77: negative refraction can be realized without employing a negative
78: index or a backward wave effect[33-37]. In this case, the PC is
79: behaving much like a uniform right-handed medium(i.e.
80: $\vec{S}\cdot\vec{k}>0$). Recently, Luo $et$ $al.$[33] have shown
81: that all-angle negative refraction could be achieved at the lowest
82: band of two-dimensional(2D) PC in the case of
83: $\vec{S}\cdot\vec{k}>0$. The advantages of the negative refraction
84: in the lowest valence band are single-mode and high transmission.
85: These can help us to design microsuperlens and realize the
86: focusing of the wave. Very recently, the subwavelength focusing
87: and image by 2D PC slab have been observed experimentally [34-36].
88: Absolute negative refraction and imaging of unpolarized
89: electromagnetic wave by 2D PC slabs have also been obtained [37].
90: However, due to the anisotropy of dispersion in 2D PC, such images
91: only appear in near-field region[38,39].
92:
93: How to realize a good-quality non-near-field image becomes an
94: important issue. The prerequisite condition to realize such a
95: phenomenon is the negative refraction which possesses the
96: single-mode and high transmission. Although some works[29] have
97: shown that the left-hand behavior and focusing exist in the 2D PC,
98: the multiple-mode and low transmission in high frequencies affect
99: the features of focusing[31,32]. It is well known that the
100: electromagnetic (EM) wave can be decomposed into TM modes (S wave)
101: and TE modes (P wave) for the 2D PC structures [40]. However, the
102: investigations[29] have shown that air-hole-type 2D PCs possess
103: good left-handed behavior for TM modes, and a pillar type 2D PC
104: prefers TE modes. In this paper, we will demonstrate that coated
105: cylinder PCs with triangular lattice posses double features. They
106: have not only good left-handed behavior for the TM modes, but also
107: for the TE modes. Most interestingly, the all-angle single-beam
108: left-handed behaviors with relative refractive index of -1 for
109: both polarized waves have been found in these systems. Thus,
110: high-quality focusing and imaging behavior have been obtained. In
111: contrast to the images in the near-field region for the lowest
112: valence band, non-near-field images, explicitly following the
113: well-known wave-beam negative refraction law, have been
114: demonstrated. The absorption and compensation for the losses by
115: introducing optical gain in these systems have also been
116: discussed.
117:
118: The rest of this paper is arranged as follows. In Sec. II, we
119: demonstrate the all-angle single-beam left-handed behaviors for
120: both TE and TM modes in a two-dimensional coated photonic
121: crystals. The good-quality non-near-field imaging behaviors are
122: discussed in Sec. III. In Sec. IV, we analyze the effect of
123: absorption and gain. The conclusions are given in Sec. V.
124:
125: \section{LEFT-HANDED BEHAVIOR in 2D COATED CYLINDER PC}
126:
127: We consider a 2D triangular lattice of coated cylinders immersed
128: in a air background with lattice constant $a$. The coated
129: cylinders have metallic cores coated with a dielectric coating.
130: The radii of metallic core and coated cylinder are 0.25a and
131: 0.45a, respectively. The dielectric constants of dielectric
132: coating are taken as $11.4$ for the S wave and $7.0$ for the P
133: wave. For the metallic component, we use the frequency-dependent
134: dielectric constant [41],
135: \begin{equation}
136: \epsilon=1-\frac{f_{p}^2}{f(f+i\gamma)},
137: \end{equation}
138: where $f_{p}$ and $\gamma$ are the plasma frequency and the
139: absorption coefficient. Following Ref.41, for all numerical
140: calculations carried out in this work, we have chosen
141: $f_{p}=3600$THZ and $\gamma=340$THZ, which corresponds to a
142: conductivity close to that of Ti. However, our discussion and
143: conclusions given below can apply to other metal parameters as
144: well. In order to simplify the problem, we first consider the
145: cases without absorption ($\gamma=0$). The effect of absorption
146: will be discussed at the latter part.
147:
148: For all calculations throughout this paper, we adopt the
149: multiple-scattering Korringa-Kohn-Rostoker method [42] as our main
150: computational tool, both to calculate the photonic band structure
151: in the reciprocal space and to perform numerical simulations for
152: wave propagating in the finite real space. The multiple-scattering
153: method is not only success in the calculations of band structure,
154: it is also best suited for a finite collection of cylinders with a
155: continuous incident wave of fixed frequency. For circular
156: cylinders, the scattering property of the individual cylinder can
157: be obtained analytically, relating the scattered fields to the
158: incident fields. The total field, which includes the incident plus
159: the multiple-scattered field, can then be obtained by solving a
160: linear system of equations, whose size is proportional to the
161: number of cylinders in the system. Both near field and far field
162: radiation patterns can be obtained straightforwardly. So, such a
163: method is a very efficient way of handling the scattering problem
164: of a finite sample containing cylinders of circular cross
165: sections, and it is capable of reproducing accurately the
166: experimental transmission data, which should be regarded as exact
167: numerical simulation. The detailed description of this method has
168: been given in Ref. [42].
169:
170: The calculated results of band structure for the S wave and the P
171: wave are plotted in the Fig.1(a) and (b), respectively. We focus
172: on the problems of wave propagation in the second bands marked by
173: dotted lines in Fig.1(a) and (b). Owing to the strong scattering
174: effects, it is generally difficult to describe the propagation
175: behavior of EM wave in the PC in a simple yet accurate way.
176: However, a lot of theoretical and experimental practices[28-39]
177: have shown that the overall behavior of the wave propagation
178: within a PC can be well described by analyzing the equifrequency
179: surface (EFS) of the band structures, because the gradient vectors
180: of constant-frequency contours in k-space give the group
181: velocities of the photonic modes. Thus, the propagation direction
182: of energy velocity of EM wave can be deduced from them. The EFS
183: contours of the above system for the S wave and the P wave at
184: several relevant frequencies are demonstrated in Fig.2(a) and (b),
185: respectively.
186:
187: It is clear from the figures that some EFS contours such as
188: $\omega=0.42-0.47(2\pi c/a)$ for the S wave and the P wave are
189: very close to a perfect circle, indicating that the crystal can be
190: regarded as an effective homogeneous medium at these frequencies.
191: At the same time, we also notice that the frequencies increase
192: inwards for both cases, meaning that $\vec{S}\cdot\vec{k_{i}}<0$
193: and the group velocities ($v_{g}$) are opposite to the phase
194: velocity. Here, $\vec{S}$ and $\vec{k_{i}}$ represent the Poynting
195: vector and wave vector, respectively. These indicate that the
196: transmitting features of the wave in the above PC structures are
197: the left-handed behavior. The conservation of the surface-parallel
198: wave vector would result in the negative refraction effect in
199: these cases, which the direction of the refracted wave inside the
200: PC can also estimated from the EFS. Then, we apply Snell's law in
201: these cases, the negative indexes of refraction can also be
202: obtained. Fig.3(a) and (b) show the negative effective indexes as
203: function of frequencies for the S wave and the P wave,
204: respectively. It is interesting that the cases with relative
205: refractive index of -1 for the S wave at $\omega=0.42(2\pi c/a)$
206: and the P wave at $\omega=0.43(2\pi c/a)$ have been found.
207:
208: In order to test the above analysis, we do numerical simulations
209: in the present systems. We take the slab samples which consist of
210: 13-layer coated cylinders in the air background with a triangular
211: arrays. The surface normal of the PC slab is along $\Gamma K$
212: direction. The parameters of coated cylinders are the same to the
213: cases in Fig.1. When a slit beam with a half width $2a$ goes
214: through the slab material, it will be refracted two times by two
215: interfaces of the slab. The shape of the sample and a snapshot of
216: the refracted process are shown on the top of Fig.4. There are
217: different ray-traces for the wave transmitting through the slab
218: sample with various effective indexes, when the wave does not
219: incident on the interface with normal direction. From the
220: ray-traces, we can deduce the effective refraction index of the
221: slab material.
222:
223: The simulations are based on a highly efficient and accurate
224: multiple-scattering method [42]. In our calculations, the widths
225: of the samples are taken enough large, such as 40a, to avoid the
226: edge diffraction effects. The calculated results for the S wave
227: and the P wave are plotted in Fig.4(a) and (b), respectively. The
228: field energy patterns of incidence and refraction are shown in the
229: figures. The arrows and texts illustrate the various beam
230: directions. The geometries of the slab are also displayed. It can
231: be clearly seen that the energy fluxes of refraction wave travel
232: following the negative refraction law for both polarized waves.
233: The negative refraction indexes for the S wave and the P wave
234: obtained from the exact numerical simulation are marked as dark
235: dot in Fig.3(a) and (b), respectively. Comparing them with the
236: estimated results (solid lines in Fig.3(a) and (b)) from the EFS ,
237: we find that the agreements between them are well around the
238: region of effective refractive index of -1 for both polarized
239: waves. When the effective refractive indexes deviate largely from
240: -1, some differences can be found. This is due to the Goos-Hanchen
241: effect, which had been discussed in Ref.[43].
242:
243: Varying the angle of inclination of the sample, we have checked
244: the cases with various incident angles. The calculated and
245: analytical results of refracted angle $\theta$ versus incident
246: angle $\theta_0$ at $\omega=0.42(2\pi c/a)$ for the S wave and
247: $\omega=0.43(2\pi c/a)$ for the P wave are summarized in Fig.5 by
248: the circle and triangular dark dots, respectively. Because the
249: frequencies are below 0.5$(2\pi c/a)$, all-angle single-beam
250: negative refractions have been observed at these frequencies. More
251: interestingly, $\theta$ is linearly proportional to $\theta_0$ in
252: the whole angle region for both polarized waves. This feature is
253: close to the ideal LHM system that can serve as a perfect
254: superlens [5].
255:
256:
257:
258: \section{IMAGE DEPENDING ON SLAB THICKNESS AND OBJECT DISTANCE}
259:
260: It is well known that an important application of negative refraction
261: materials is the microsuperlens [5]. Ideally, such a superlens can
262: focus a point source on one side of the lens into a real point
263: image on the other side even for the case of a parallel sided slab
264: of material. It possesses some advantages over conventional
265: lenses. For example, it can break through the traditional
266: limitation on lens performance and focus light on to an area
267: smaller than a square wavelength.
268:
269: In order to model such a superlens, we
270: take a slab sample with 40a width and 11a thickness. A
271: continuous-wave point source is placed at a distance 5.5a (half
272: thickness of the sample) from the left surface of the slab. We
273: first discuss the case for the point source of the S wave. The
274: frequency of the incident wave emitting from such a point source
275: is $0.42(2\pi c/a)$, which is corresponded to the case with
276: relative refractive index of -1. If the wave transmits in such a
277: 2D PC slab according to the well-known wave-beam refraction law,
278: one should observe the focusing point in the middle of the slab
279: and the image at the symmetric position in the opposite side of
280: the slab, as being depicted by the simple picture on the top of
281: Fig.6.
282:
283: To see whether or not such a phenomenon exists, we employ the
284: multiple-scattering method [42] to calculate the propagation of
285: waves in such a system. A typical field intensity pattern for the
286: S wave across the above slab sample is plotted in Fig. 6(b).
287: X and Y present vertical and transverse direction of wave propagating,
288: respectively. The field intensity in figure is over $30a\times
289: 30a$ region around the center of the sample. The geometry of the
290: PC slab is also displayed for clarity of view. The high quality
291: image in the opposite side of the slab and the focusing in the
292: middle of the slab according to the wave-beam refraction law are
293: observed clearly. A closer look at the data reveals a transverse
294: size (full size at half maximum) of the image spot as $0.7a$ (or
295: $0.3\lambda$), which is well below the conventional diffraction
296: limit.
297:
298: In order to clarify the sample thickness dependence of the image
299: and focusing, we have also checked a series of slab samples with
300: various thickness. Similar phenomena have also been observed. For
301: example, Fig. 6(a) shows the calculated field energy pattern for a
302: 7a thick sample. A monochromatic point source with
303: $\omega=0.42(2\pi c/a)$ is placed at a distance of half thickness
304: of the sample (3.5a) from the left surface of the slab and its
305: image is found again near the symmetric position in the opposite
306: side of the slab.
307:
308: To have a more complete vision on the imaging effect of this type
309: of superlens, we move the light source and see what happens to the
310: imaging behavior. We first put a point source near the left
311: surface of the slab. In this case, the refracted process following
312: wave-beam negative refraction law is shown on the top of Fig.7.
313: The calculated intensity distribution is plotted in Fig.7. In our
314: simulations, the point source is placed at a distance of 2.0a from
315: the left surface of the sample and the image is found near 9a from
316: the right surface. The excellent agreements between the
317: simulations and the estimated results from the rules of geometric
318: optics are obvious. The corresponded result that a point source is
319: far from the left surface is displayed in Fig.8. In this case, the
320: point source is placed at a distance of 9.0a from the left surface
321: of the sample and the image is found near 2a from the right
322: surface. Comparing the simulations with the snapshot of refracted
323: process according to the wave-beam negative refraction law shown
324: on the top of Fig.8, we again find the excellent agreements
325: between them.
326:
327: The above results are only for the S wave, in the following, we
328: will investigate the case of the P wave. We take the PC slab
329: samples with 40a width and various thicknesses. A point source of
330: the P wave is placed at a distance of half thickness of the sample
331: from the left surface of the slab. The frequency of the incident
332: wave emitting from such a point source is $0.43(2\pi c/a)$, which
333: is corresponded to the case with relative refractive index of -1
334: for the P wave. The propagation behaviors of the P waves in such
335: systems are still calculated by the multiple-scattering method
336: [42]. Fig.9(a) and (b) show the cases with 7a and 13a thick slab,
337: respectively. Similar features to Fig.7 for the S wave are found.
338: If we move the source position, the effect of source position on
339: the image for the P wave can also be checked. The calculated
340: results under two kinds of source position for the P wave are
341: plotted in Fig.(10)(a) and (b), respectively. The cases for the P
342: waves are similar to those of the S waves in Fig.8.
343:
344: These observations indicate clearly that the imaging behaviors
345: depend on the slab thickness and the object distance, explicitly
346: following the well-known wave-beam negative refraction law.
347: Therefore, such PC slabs in some frequencies can be considered as
348: homogeneous effective medium with effective refraction index of
349: -1. The high-quality focusing and images can be realized in these
350: PC systems for both S wave and P wave.
351:
352: \section{EFFECT OF ABSORPTION AND GAIN}
353:
354: The above investigations have shown that the PC slab consisting of
355: coated cylinders is actually considered as a good superlens for
356: both polarized waves. The common features of these coated systems
357: are that they all include metal components. Therefore, the
358: absorption for these systems is inevitable. Lou $et$ $al.$ [38]
359: have pointed that the central image peak disappear and the image
360: degrade gradually with the increase of absorption. However,
361: fortunately, the loss can be overcome by introducing the optical
362: gain in the systems. Recently, Ramakrishna and Pendry [19] have
363: suggested a method to remove the absorption by introducing optical
364: gain into the lens made from a multilayers stack of thin
365: alternating layers of silver and dielectric medium. Here, we
366: borrow their idea and introduce the optical gain in the 2D PC
367: superlens.
368:
369: Fig.11(a) shows the intensity distribution as a function of
370: transverse coordinate ($y/a$) for the S wave at the image plane
371: (5.5a away from the second interface). Curve {\it A} is
372: corresponded to the perfect case without absorption and curve {\it
373: B} to that with absorption, in this case $\gamma$ is taken as
374: $340$THZ. Comparing the curve {\it A} with the curve {\it B}, we
375: find that the central peak of image decrease with the introducing
376: of absorption, which is agree with the analysis of Ref.[38]. This
377: is also consistent with the numerical studies of left-handed
378: structures constructed from split-ring resonators in Ref.[7,8].
379: The result by introducing gain to remove the absorption for the
380: corresponding case is plotted in Fig.11(b). Curves {\it B} in the
381: Fig.11(a) and (b) are the same one, and curve {\it C} in the
382: Fig.11(b) is the result with the dielectric constant
383: $\epsilon=11.4-0.08i$ for the dielectric part of coated cylinder.
384: We do not find any difference between curve {\it A} in the
385: Fig.11(a) and curve {\it C} in the Fig.11(b).
386:
387: Similar phenomenon can also be found for the P wave. Fig.12(a) and
388: (b) show the corresponded case of the P wave. Fig.12(a) represents
389: the intensity distribution as a function of transverse coordinate
390: ($y/a$) for the P wave at the image plane (5.5a away from the
391: second interface). Comparing the curve {\it A} without absorption
392: with the curve {\it B} in presence of absorption in Fig.12(a), we
393: find that the absorption decreases the central peak of the image
394: as the case of the S wave. However, with the introducing of gain
395: such as $\epsilon=11.4-0.06i$ for the dielectric part of coated
396: cylinder, the lose due to absorption can be compensated
397: completely. Curve C in Fig.12(b) represents such a case. In fact,
398: for any cases of absorption, the losses can always be compensated
399: by introducing fitted gain for both polarized waves. Thus, the
400: lens based on the above 2D PC can work well even in presence of
401: absorption.
402:
403: \section{SUMMARY}
404:
405: Through the exact numerical simulation and physical analysis, we
406: have demonstrated all-angle single-beam left-handed behavior for
407: both TE and TM modes in the 2D coated photonic crystals. More
408: interestingly, the relative refractive index of -1 for both
409: polarized waves have been found. Furthermore, the refracted angle
410: is linearly proportional to the incident angle in the whole angle
411: region for both polarized waves have been demonstrated. These
412: feature are close to the ideal LHM system that can serve as a
413: perfect superlens. The imaging behaviors by 2D coated
414: photonic-crystal-based superlens have been investigated
415: systematically. Good-quality images and focusing, with relative
416: refractive index of -1 and explicitly following the well-known
417: wave-beam negative refraction law, have been observed in these
418: systems for both polarized waves.
419:
420: Our above results are in contrast to the previous investigations
421: about the image behaviors for the LHM-based superlenes and the
422: PC-based microsuperlenes at the lowest valence band. For these
423: cases, the images only appear in the near-field region, which does
424: not follow rules of geometric optics [22-27,38,39]. According to
425: Pendry's analysis [5] , the ``perfect" image arises from the
426: enhancement of evanescent components of incoming waves, and
427: surface plasmons play an important role in the ``perfect" imaging
428: [5,21]. So, the quality of the image is affected by many factors
429: such as interface feature, finite-size and cavity resonance
430: [22-27]. These will limit further applications. Here, our
431: superlens based on the coated PC systems with triangular lattice
432: can overcome these restrictions. Thus, extensive applications of
433: such a phenomenon to optical devices are anticipated.
434:
435: In addition, the absorption and compensation for the losses by introducing
436: optical gain in these systems have also been discussed. In
437: general, for the PC structures with metal components, increased
438: absorption in metals prohibits the scaling of these structures to
439: the optical wavelengths. However, since the losses by absorption
440: can always be compensated by introducing fitted gain in our coated
441: PC systems, many negative refraction phenomena that have been
442: observed in the microwave regime can also be found in the optical
443: wavelengths. These features make the PC slabs consisting of coated
444: cylinders promising for application in a range of optical devices,
445: such as a superlens for visible light.
446:
447:
448:
449: \begin{center}
450: Acknowledgments
451: \end{center}
452: This work was supported by the National Natural Science Foundation
453: of China (Grant No.10374009) and the National Key Basic Research
454: Special Foundation of China under Grant No.2001CB610402. The
455: project sponsored by SRF for ROCS, SEM and the Grant from Beijing
456: Normal University.
457: \newpage
458: \begin{references}
459: \bibitem{1}V.G.Veselago, Sov.Phys.Uspekhi {\bf 8}, 2854 (1967)[Sov. Phys. Usp. 10, 509(1968)].
460: \bibitem{2}J. B. Pendry, A.J.Holden, D.J.Robbins, and W.J.Stewart, IEEE Trans. Microwave Theory
461: Tech. {\bf 47}, 2075 (1999).
462: \bibitem{3}D.R.Smith, W.J.Padilla, D.C.View, S.C.Nemat-Nasser, and S.Schultz,
463: Phys. Rev. Lett. {\bf 84}, 4184 (2000); D.R.Smith and N. Kroll,
464: Phys. Rev. Lett. {\bf 84}, 2933 (2000).
465: \bibitem{4}R. A. Shelby, D.R.Smith, and S.Schultz, Science {\bf 292}, 77 (2001).
466: \bibitem{5}J. B. Pendry, Phys. Rev. Lett. {\bf 85}, 3966 (2000).
467: \bibitem{6}P.Markos and C.M.Soukoulis, Phys. Rev. E {\bf 65}, 036622(2002); Phys. Rev. B {\bf 65}, 033401(2002).
468: \bibitem{7} Markos, I. Rousochatzakis and C. M. Soukoulis, Phys. Rev. E 66 045601(R) (2002).
469: \bibitem{8} R.B. Greegor, C.G.Parazzoli, K.Li and M.H.Tanielian, Appl. Phys. Lett. 82, 2356(2003).
470: \bibitem{9}S. Foteinopoulou, E.N. Economou and C.M.Soukoulis,
471: Phys. Rev. Lett. {\bf 90}, 107402(2003).
472: \bibitem{10}J. Pacheco, Jr., T.M.Grzegorczyk, T B.I.Wu, Y.Zhang and J.A.Kong,
473: Phys. Rev. Lett. {\bf 89}, 257401(2002).
474: \bibitem{11}Y. Zhang, B. Fluegel, and A. Mascarenhas, Phys. Rev. Lett. {\bf 91}, 157404(2003).
475: \bibitem{12}D.R.Smith and D.Schurig, Phys. Rev. Lett. {\bf 90}, 077405 (2003).
476: \bibitem{13}A.A.Houck, J. B. Brock and I. L. Chuang,
477: Phys. Rev. Lett. {\bf 90}, 137401(2003); C.G.Parazzoli, R. B.
478: Greegor, K. Li, B. E. C. Koltenbah, and M. Tanielian, Phys. Rev.
479: Lett. {\bf 90}, 107401 (2003).
480: \bibitem{14}J. Li, Lei Zhou, C.T.Chan and P.Sheng,
481: Phys. Rev. Lett. {\bf 90}, 083901(2003).
482: \bibitem{15}R. Ziokowski and E. Heyman, Phys. Rev. E {\bf 64}, 056625(2001).
483: \bibitem{16}Focus Issue ``Negative refraction and metamaterials", Optics Express {\bf vol. 11},no.7 (2003).
484: \bibitem{17}G. Shvets, Phys. Rev. B {\bf 67},035109(2003).
485: \bibitem{18}V. A. Podolskiy, A.K.Sarychev and V.M. Shalaev, Optics Express {\bf 11},no.735 (2003).
486: \bibitem{19}S. A. Ramakrishna and J.B. Pendry, Phys. Rev. B {\bf 67}, 201101(R)(2003).
487: \bibitem{20} A.A.Zharov, I.V.Shadrivov and Y. S. Kivshar, Phys. Rev. Lett. {\bf 91},
488: 037401(2003); V. M. Agranovich, Y.R.Shen, R.H.Baughman and A.A.
489: Zakhidov, Phys. Rev. B {\bf 69}, 165112 (2004).
490: \bibitem{21}G.W.'t Hooft, Phys. Rev. Lett. {\bf 87}, 249701(2001);
491: J. M. Williams, Phys. Rev. Lett. {\bf 87}, 249703(2001); N.Garcia
492: and M.Nieto-Vesperinas, Phys. Rev. Lett. {\bf 88}, 207403(2002);
493: A.L.Pokrovsky and A. L. Efros, Phys. Rev. Lett. {\bf 89}, 093901
494: (2002).
495: \bibitem{22}R.Merlin, Appl.Phys.Lett. {\bf 84}, 1290(2004).
496: \bibitem{23}L. Chen, S. He and L. Shen, Phys. Rev. Lett. {\bf 92}, 107404 (2004).
497: \bibitem{24} D.R. Smith, D. Schurig, J. J. Mock, P. Kolinko, and P. Rye, Appl. Phys. Lett. 84, 2244 (2004).
498: \bibitem{25} Zhaowei Liu, Nicholas Fang, Ta-Jen Yen, and Xiang Zhang, Appl.Phys.Lett. {\bf 83}, 5184(2003).
499: \bibitem{26}A. Grbic and G.V.Eleftheriades, Phys. Rev. Lett. {\bf 92}, 117403 (2004).
500: \bibitem{27}A.N.Lagarkov and V.N.Kissel, Phys. Rev. Lett. {\bf 92}, 07701 (2004).
501:
502: \bibitem{28}H. Kosaka, T. Kawashima, A. Tomita, M.Notomi, T.Tamamura, T.Sato, and S.Kawakami,
503: Phys. Rev. B {\bf 58}, 10096(1998).
504: \bibitem{29}M.Notomi, Phys. Rev. B {\bf 62}, 10696(2000).
505: \bibitem{30}B.Gralak, S.Enoch, and G.Tayeb, J. Opt. Soc. Am. A {\bf 17}, 1012(2000).
506: \bibitem{31}S. Foteinopoulou and C.M.Soukoulis,
507: Phys. Rev. B {\bf 67}, 235107(2003).
508: \bibitem{32}P.V.Parimi, W.T.Lu, P.Vodo, J.Sokoloff, J.S. Derov and S. Sridhar, Phys. Rev. Lett. {\bf 92}, 127401 (2004).
509:
510: \bibitem{33}C.Luo, S.G.Johnson, J.D.Joannopoulos, J.B.Pendry, Phys. Rev. B {\bf 65},
511: 201104(R)(2002); optics express, {\bf 11}, 746(2003); C.Luo,
512: S.G.Johnson, J.D.Joannopoulos, Appl.Phys.Lett. {\bf 83},
513: 2352(2002).
514: \bibitem{34}E. Cubukcu, K. Aydin, E. Ozbay, S. Foteinopoulou and
515: C.M.Soukoulis, Nature (London) {\bf 423}, 604(2003).
516: \bibitem{35} P. V. Parimi, W.T.Lu, P. Vodo and S. Sridhar, Nature {\bf 426}, 404(2003).
517: \bibitem{36}E. Cubukcu, K. Aydin, E. Ozbay, S. Foteinopoulou and
518: C.M.Soukoulis, Phys. Rev. Lett. {\bf 91}, 207401(2003).
519: \bibitem{37} Xiangdong Zhang, phys. Rev. B (accepted).
520: \bibitem{38}C.Luo, S.G.Johnson, J.D.Joannopoulos, J.B.Pendry, Phys. Rev. B {\bf 68},
521: 045115(2003).
522: \bibitem{39}Z.Y. Li and L.L. Lin, Phys. Rev. B {\bf 68},245110(2003).
523: \bibitem{40}J.D.Joannopolous, R.D.Meade, and J.N.Winn, Photonic Crystals (Princeton University, Princeton, 1995).
524: \bibitem{41}M. M. Sigalas, C. T. Chan, K. M. Ho and C.M. Soukoulis, phys. Rev. B {\bf 52}, 11744
525: (1995); L.M. Li, Z.Q. Zhang and X. Zhang, phys. Rev. B {\bf 58},
526: 15589 (1998).
527: \bibitem{42}L.M.Li and Z.Q.zhang,phys. Rev. B {\bf 58}, 9587
528: (1998); X.Zhang, Z.Q.Zhang, and L.M.Li, C. Jin, D. Zhang, B. Man
529: and B. Cheng, Phys. Rev. B {\bf 61}, 1892 (2000).
530: \bibitem{43} D. Felbacq and R. Smaali , Phys. Rev. Lett. {\bf 92}, 193902(2004).
531:
532:
533:
534:
535:
536:
537:
538:
539:
540:
541:
542: \end{references}
543: \newpage
544:
545: FIGURE CAPTIONS
546:
547: Fig.1, The calculated photonic band structures of a triangular
548: lattice of coated cylinder in air for S wave (a) and P wave (b).
549: The radii of the dielectric cylinder and inner metallic cylinder
550: are $R=0.45a$ and $r=0.25a$, respectively. The dielectric
551: constants are $\epsilon=11.4$ for the S wave and $\epsilon=7$ for
552: the P wave. Dotted lines mark the region for negative refraction.
553:
554: Fig.2,Several constant-frequency contours for S wave (a) and P
555: wave (b) of the second band of the 2D PCs which are corresponded
556: to the cases in Fig.1(a) and (b), respectively. The numbers in the
557: figure mark the frequencies in unit of $2\pi c/a$. $k_{i}$ and
558: $v_{g}$ represent the wave vector and the group velocity,
559: respectively.
560:
561: Fig.3, Effective indexes versus frequencies for S wave (a) and P
562: wave (b). The crystals and parameters are identical to those in
563: Fig.1. The positions of relative refractive index of -1 are
564: marked by dotted lines.
565:
566: Fig.4, Simulation of negative refraction. The shape of the sample
567: and a snapshot of refraction process are shown on top of the
568: figure. The intensities of electric field for S wave (a) and
569: magnetic field for P wave (b) for incidence and refraction are
570: shown. The 2D PC slabs with 13-layer are marked as dark dots in
571: figures. The frequencies of incident wave are $\omega=0.42(2\pi
572: c/a)$ for the S wave and $\omega=0.43(2\pi c/a)$ for the P wave.
573: The crystals and parameters in (a) and (b) are corresponded to
574: those in Fig.1(a) and (b), respectively.
575:
576: Fig.5, The angles of refraction ($\theta$) versus angles of
577: incidence ($\theta_0$) at $\omega=0.42(2\pi c/a)$ for S wave and
578: $\omega=0.43(2\pi c/a)$ for P wave. Circle dots are corresponded
579: to the S wave and triangular dots to the P wave.
580:
581: Fig.6, (a) The intensity distributions of point source and its
582: image across a 7a 2D PC slab at frequency $\omega=0.42(2\pi c/a)$
583: for S wave. (b) The corresponding case for a slab with 11a
584: thickness. Schematic picture depicting the lensing of a source by
585: a PC slab to an image are shown on top of the figure.
586:
587: Fig.7, The intensity distributions of point source and its image
588: across a 11a 2D PC slab at frequency $\omega=0.42(2\pi c/a)$ for S
589: wave. The point source is placed at 2a distance from the left
590: surface of the slab. Schematic picture depicting the lensing of a
591: point source by a PC slab to an image are shown on top of the
592: figure.
593:
594: Fig.8, The intensity distributions of point source and its image
595: across a 11a 2D PC slab at frequency $\omega=0.42(2\pi c/a)$ for S
596: wave. The point source is placed at 9a distance from the left
597: surface of the slab. Schematic picture depicting the lensing of a
598: point source by a PC slab to an image are shown on top of the
599: figure.
600:
601: Fig.9, (a) The intensity distribution of point source and its
602: image across a 7a 2D PC slab at frequency $\omega=0.43(2\pi c/a)$
603: for P wave. (b) The corresponding case for a slab with 11a
604: thickness. The point source is placed at a distance with half
605: thickness of the sample from the left surface of the slab.
606:
607: Fig.10, The intensity distribution of point source and its image
608: across a 11a 2D PC slab at frequency $\omega=0.43(2\pi c/a)$ for P
609: wave. (a) and (b) represent the cases with different source
610: positions, 2a and 9a distances from the left surface of the slab,
611: respectively.
612:
613: Fig.11, Intensity distribution along the transverse (y) direction
614: at the image plane for S wave. (a) The case with absorption (B)
615: and that without absorption (A). (b) The case with absorption (B)
616: and that with absorption and gain (C). The crystal and parameters
617: are identical to those in Fig.6.
618:
619: Fig.12, Intensity distribution along the transverse (y) direction
620: at the image plane for P wave. (a) The case with absorption (B)
621: and that without absorption (A). (b) The case with absorption (B)
622: and that with absorption and gain (C). The crystal and parameters
623: are identical to those in Fig.9.
624:
625:
626:
627: \end{document}
628: