Slow- and fast-light: fundamental limitations

Similar documents
Slow, Fast, and Backwards Light Propagation in Erbium-Doped Optical Fibers. Zhimin Shi

Slow, Fast, and Backwards Light: Fundamentals and Applications Robert W. Boyd

Slowing Down the Speed of Light Applications of "Slow" and "Fast" Light

Robert W. Boyd. Institute of Optics and Department of Physics and Astronomy University of Rochester

Performance Limits of Delay Lines Based on "Slow" Light. Robert W. Boyd

Slow and Fast Light in Room-Temperature Solids: Fundamental and Applications. Robert W. Boyd

Brief Research Update

Limits on the Time Delay Induced by Slow-Light Propagation

Superluminal Light Pulses, Subluminal Information Transmission

All-Optical Delay with Large Dynamic Range Using Atomic Dispersion

arxiv:quant-ph/ v1 2 Oct 2003

The speed of light in a vacuum

Applications of Slow Light. Robert W. Boyd. Institute of Optics and Department of Physics and Astronomy University of Rochester

Light transmission through and its complete stoppage in an ultra slow wave optical medium

Slow and fast light: fundamentals and applications Robert W. Boyd a a

Optical delay with spectral hole burning in Doppler broadened cesium vapor

Slow Light in Crystals

Ultra-Slow Light Propagation in Room Temperature Solids. Robert W. Boyd

Maximizing the opening of eye diagrams for slow-light systems

arxiv:physics/ v1 9 Jan 2006

9 Atomic Coherence in Three-Level Atoms

Absorption-Amplification Response with or Without Spontaneously Generated Coherence in a Coherent Four-Level Atomic Medium

Quantum Information Storage with Slow and Stopped Light

Slow light with a swept-frequency source

A Review of Slow Light Physics and Its Applications

Distortion management in slow-light pulse delay

Stopped Light With Storage Times Greater than 1 second using Electromagnetically Induced Transparency in a Solid

Fast Light, Slow Light

SUB-NATURAL-WIDTH N-RESONANCES OBSERVED IN LARGE FREQUENCY INTERVAL

Slow and stored light using Rydberg atoms

THE KRAMERS-KRONIG RELATIONS FOR LIGHT MOVING BACKWARDS IN TIME. Evangelos Chaliasos. 365 Thebes Street. GR Aegaleo.

VIC Effect and Phase-Dependent Optical Properties of Five-Level K-Type Atoms Interacting with Coherent Laser Fields

Slow light in various media: a tutorial

Metamaterial-Induced

Roles of Atomic Injection Rate and External Magnetic Field on Optical Properties of Elliptical Polarized Probe Light

Enhanced Nonlinear Optical Response from Nano-Scale Composite Materials

Breakup of Ring Beams Carrying Orbital Angular Momentum in Sodium Vapor

A Stern-Gerlach experiment for slow light

Spatial evolution of laser beam profiles in an SBS amplifier

Experimental constraints of using slow-light in sodium vapor for light-drag enhanced relative rotation sensing

Ultra-Slow (and Superluminal) Light Propagation in Room Temperature Solids. Robert W. Boyd

Photon Physics. Week 5 5/03/2013

Plasmonics: elementary excitation of a plasma (gas of free charges) nano-scale optics done with plasmons at metal interfaces

Elements of Quantum Optics

EIT and diffusion of atomic coherence

Asymptotic description of wave propagation in an active Lorentzian medium

Electromagnetically Induced Transparency (EIT) via Spin Coherences in Semiconductor

Negative refractive index in a four-level atomic system

Atomic Coherent Trapping and Properties of Trapped Atom

arxiv:quant-ph/ v1 24 Jun 2005

B 2 P 2, which implies that g B should be

Analytical Investigation of Slow Light Systems with Strained Quantum Wells Structure under Applied Magnetic and Electric Fields Based on V-type EIT

Transparent anomalous dispersion and superluminal light-pulse propagation at a negative group velocity

Gain without inversion in a V-type system with low coherence decay rate for the upper levels. arxiv:physics/ v1 [physics.atom-ph] 15 May 2004

Chapter 24 Photonics Question 1 Question 2 Question 3 Question 4 Question 5

Lecture 14 Dispersion engineering part 1 - Introduction. EECS Winter 2006 Nanophotonics and Nano-scale Fabrication P.C.Ku

Effects of self-steepening and self-frequency shifting on short-pulse splitting in dispersive nonlinear media

Zeno logic gates using micro-cavities

Nonlinear effects and pulse propagation in PCFs

Spectroscopy in frequency and time domains

MODERN OPTICS. P47 Optics: Unit 9

Slow and Fast Light. September 24, 2001

Free-Electron Lasers

Information Theoretic Framework for the Analysis of a Slow-Light Delay Device Mark A. Neifeld, 1,2 and Myungjun Lee 1, * 1 Department of Electrical Co

Optical Multi-wave Mixing Process Based on Electromagnetically Induced Transparency

Superluminal optical pulse propagation in nonlinear coherent media

Slowdown of group velocity of light in PPLN by employing electro-optic effect

Enhanced Nonlinear Optical Response from Nano-Scale Composite Materials

Do we need quantum light to test quantum memory? M. Lobino, C. Kupchak, E. Figueroa, J. Appel, B. C. Sanders, Alex Lvovsky

THEORETICAL PROBLEM 2 DOPPLER LASER COOLING AND OPTICAL MOLASSES

arxiv: v2 [physics.optics] 11 Jan 2015

ELECTROMAGNETICALLY INDUCED TRANSPARENCY IN RUBIDIUM 85. Amrozia Shaheen

Stored light and EIT at high optical depths

Dmitriy Churin. Designing high power single frequency fiber lasers

Interference Between Distinguishable States. Thomas Alexander Meyer

Nonlinear Effects in Optical Fiber. Dr. Mohammad Faisal Assistant Professor Dept. of EEE, BUET

FIBER OPTICS. Prof. R.K. Shevgaonkar. Department of Electrical Engineering. Indian Institute of Technology, Bombay. Lecture: 15. Optical Sources-LASER

Quantum Electronics Laser Physics. Chapter 5. The Laser Amplifier

Transit time broadening contribution to the linear evanescent susceptibility

External (differential) quantum efficiency Number of additional photons emitted / number of additional electrons injected

Spectral Broadening Mechanisms

Analysis of second-harmonic generation microscopy under refractive index mismatch

Cavity decay rate in presence of a Slow-Light medium

The physics of the perfect lens

Linear and nonlinear spectroscopy

An optical rotation sensor based on dispersive slow-light medium

F. Elohim Becerra Chavez

arxiv: v1 [physics.optics] 26 Sep 2017

Quantum Electronics/Laser Physics Chapter 4 Line Shapes and Line Widths

Observation of the nonlinear phase shift due to single post-selected photons

PRINCIPLES OF NONLINEAR OPTICAL SPECTROSCOPY

What is spectroscopy?

Tailorable stimulated Brillouin scattering in nanoscale silicon waveguides.

arxiv:quant-ph/ v3 17 Nov 2003

Supplementary Information: Three-dimensional quantum photonic elements based on single nitrogen vacancy-centres in laser-written microstructures

NON LINEAR PULSE EVOLUTION IN SEEDED AND CASCADED FELS

Introduction to Nonlinear Optics

Nonlinear Electrodynamics and Optics of Graphene

Time Domain Modeling of All-Optical Switch based on PT-Symmetric Bragg Grating

Absorption and scattering

Transcription:

Journal of Modern Optics Vol. 54, Nos. 16 17, 10 20 November 2007, 2403 2411 Slow- and fast-light: fundamental limitations ROBERT W. BOYD*y and PAUL NARUMz ythe Institute of Optics, University of Rochester, Rochester, NY 14627, USA znorwegian Defence Research Establishment, Norway (Received 19 March 2007; in final form 14 August 2007) We present an analysis of the propagation of light pulses through materials possessing extreme values of the group velocity. We begin with an analysis of the behaviour that occurs upon propagation through materials possessing simple Lorentzian gain or absorption lines or materials possessing sharp dips in gain or absorption features. We also describe how more complicated lineshapes can be used to tailor the dispersion of a slow-light system. We furthermore present an analysis of fundamental limitations to how many pulse widths a pulse of light can be delayed or advanced. We show how these mechanisms lead to different limits for slow and fast propagation media. 1. Introduction In recent years, many researchers have been intrigued by the recently acquired ability to study light propagation under highly exotic conditions [1, 2]. Equally exciting is the possibility that these techniques will lead to useful applications in the fields of telecommunications and electronics [3, 4]. This paper is devoted to the examination of some fundamental aspects of slow- and fast-light propagation. In particular, we make a careful examination of the role of atomic resonances in achieving extreme values of the light propagation velocity, and we analyse some of the physical processes that can pose a limitation on how much a pulse of light can be delayed or advanced upon propagation through a resonant medium. 2. Methods for controlling the speed of light We begin by making some general observations regarding several of the mechanisms that have been used to control the velocity of light. We first note that the most dramatic modifications of the velocity of light are often those based on the exploitation of atomic resonances. The reason for this *Corresponding author. Email: boyd@optics.rochester.edu Journal of Modern Optics ISSN 0950 0340 print/issn 1362 3044 online # 2007 Taylor & Francis http://www.tandf.co.uk/journals DOI: 10.1080/09500340701639649

2404 R. W. Boyd and P. Narum behaviour is the following. The velocity with which pulses of light propagate through a material system is conventionally described by the group velocity where the group index is given by v g ¼ c n g, n g ¼ n þ! dn d! : ð2þ In these equations, c is the speed of light in vacuum,! is the angular frequency of the optical radiation, and n is the usual refractive index. In essentially all slow-, fast- and backwards-light situations, the second term in the expression for n g dominates. Thus, highly dispersive media are best suited for displaying extreme values of the group velocity, and sharp spectral features such as those encountered in atomic vapours are well suited for producing these effects. Note that dn=d! can be either positive or negative. When the group index is large and positive, one speaks of slow light [5 12]; when the group index is either negative or positive and less than unity one speaks of fast light [13 18]. The special situation in which the group index is negative is sometimes known as backwards light. Backwards light is considered to be a form of fast light because, even though the peak of the pulse moves backwards within the material, the peak emerges from the material earlier than it would for propagation through an equal length of vacuum. Also intriguing is the observation that certain material systems can display either slow- or fast-light depending on the details of the experimental conditions [19, 20]. The simplest procedure for obtaining extreme values of the group velocity is to make use of the linear optical response of an isolated absorption or gain resonance. This situation is illustrated in figure 1. The case of an absorption resonance is ð1þ (a) Absorption resonance (b) g Gain resonance n 0 n 0 n g n g Figure 1. Slow- and fast-light based on isolated atomic resonances: (a) absorption resonance; (b) gain resonance.

Slow- and fast-light: fundamental limitations 2405 illustrated on the left of the figure and that of a gain resonance on the right of the figure. Associated with the absorption or gain resonance is a rapid frequency variation of the refractive index, as required by Kramers Kronig relations, which is illustrated in the central panel on each side. Note that the sense of the frequency variation is exactly opposite in the two cases. The bottom panel on each side shows the resulting variation of the group index, as predicted by equation (2). Note that, for an isolated absorption resonance, fast-light occurs on line centre and slow-light occurs in the wings of the line; the situation is just the opposite for the case of a gain resonance. Because of the large attenuation that light experiences at the centre of a strong absorption resonance, it is difficult (but not impossible [13]) to observe fast light under these circumstances. In most practical uses of slow- and fast-light it is necessary to be able to actively control the value of the group velocity. Nonlinear optical methods are often used to achieve this control of the group velocity. For example, electromagnetically induced transparency [9], coherent population oscillations [10], and saturated absorption [11] have been used to create narrow transparency windows in absorbing materials. The variation in refractive index associated with these absorption dips can be used to create slow- and fast-light. This situation is illustrated on the left-hand side of figure 2, with the analogous situation for a spectral hole in a gain medium illustrated on the right. As in figure 1, the central panels show the associated variation in refractive index and the lower panels show the variation of the group index. Again, in both cases either slow- or fast-light can be obtained depending on the detuning from line centre. In certain applications of slow- and fast-light, it is desirable to tailor the dispersion properties to minimize the effect of pulse distortion. The procedure of Stenner et al. [12] is shown in figure 3. These researchers make use of a double gain line to flatten the gain profile of their material system. As a result, the refractive index profile has the form shown in the lower panel. Note that the region of nearly linear (a) (b) g n Dip in gain feature 0 0 n Dip in absorption feature n g n g Figure 2. Slow- and fast-light based on a dip in (a) a gain or (b) absorption feature.

2406 R. W. Boyd and P. Narum (a) Gain coefficient g (ω) (b) 0 Double gain line Single gain line Extended region of linear dependence Double gain line Flattened gain profile Contributions to double gain line Single gain line 8 6 4 2 0 2 4 6 8 Frequency (ω ω 0 )/γ Figure 3. Example of disperson management in a slow-light system [12]. The use of a spectrally flattened, double gain line (a) produces a broader spectral region of nearly linear dispersion (b). (The colour version of this figure is included in the online version of the journal.) dependence of index of refraction upon detuning is larger in this case than for a single gain line. As a result, short optical pulses, which necessarily possess a broad frequency spectrum, will suffer less distortion when propagating through the doublegain-line system. 3. Fundamental limits on slow- and fast-light A key figure of merit for many slow- and fast-light systems is the number of pulse widths of delay or advancement that the system can produce. This number, which we will call the normalized delay or advancement, is equal to the number of pulses that can be contained at any time within the material and is a measure of the information storage capacity of the system. There is considerable conceptual and practical interest in determining what fundamental physical processes, if any, pose a limitation on how large this storage capacity can be [21 26]. In many situations, this storage capacity is limited by pulse distortion that will accompany large propagation distances. Pulse distortion is caused primarily by group velocity dispersion and spectral reshaping of the input pulse [23]. When working at or near line centre, the situation considered in the present paper, spectral reshaping, is the dominant mechanism. When operating far from line centre, group velocity dispersion is the dominant effect.

Slow- and fast-light: fundamental limitations 2407 As is well known, for a slow-light medium the pulse will tend to broaden in time upon propagation, whereas for a fast-light medium the pulse will tend to become narrower [22]. If the limit on pulse broadening is set to be a factor of 2 1=2 and the limit on pulse narrowing is set so that to first order the pulse becomes infinitesimally short, we find that the maximum propagation distance is set by the following inequality: L ðtþ 2 : Here is the line-centre gain or absorption coefficient, L is the length of the medium, T is the pulse length, and is the relevant linewidth, that is, the width of the isolated feature for the example given in figure 1 or of the dip in the overall profile for the example given in figure 2. The conclusion of equation (3) follows, for example, from the argument used to go from equation (13) to (14) in [23]. The maximum delay or advancement is obtained through use of this maximum value of L and is given by T T ¼ 1 2ð2 1=2 Þ ðlþ1=2 : ð4þ For the case of a slow-light system, there appears to be no fundamental limits on the extent to which one can delay a pulse of light [23], although there are very serious practical problems [24, 25]. Conversely, as we shall demonstrate below, for a fastlight system there seem to be limitations of a nearly fundamental nature that determine how much one can advance a pulse of light [22]. This limitation is often at most two pulse widths of advancement and in practice is often still more restrictive. The question then is raised as to why the two situations are so different. At first glance, the answer to this question may appear to be almost obvious. Causality requirements, for example, limit how far the information content of a pulse of light can be advanced in time, but place no limit on how much it can be delayed. However, upon reflection one concludes that this explanation cannot be adequate, because causality places a restriction on the information velocity [15] but not on the group velocity, the quantity that we are concerned with here. Moreover, we have seen in the previous section that there appears to be an almost complete symmetry between slow- and fast-light interactions, at least at the level of the atomic response. We believe that we have uncovered two key differences between slow- and fast-light situations that explain why the limitations on fast-light propagation are so much more restrictive than for slow-light media. These differences are described next. (i) Restrictions based on maximum allowable gain or loss. First, we note that for the case of an absorbing medium, if the attenuation at the signal frequency is too large, the transmitted pulse will be too weak to be detectable. We arbitrarily choose a minimum intensity transmission of T ¼ exp ð 0 LÞ¼ exp ð 32Þ at the signal frequency in our studies. The value 32 is chosen arbitrarily, but our conclusions do not depend critically upon the actual numerical value. For the case of a gain medium, the nature of the restriction is somewhat different. Here we require that the gain not be too large at any frequency, because if the gain is too large at some frequency the process of ð3þ

2408 R. W. Boyd and P. Narum amplified spontaneous emission will occur at that frequency and deplete the gain of the material. We place a numerical restriction that the intensity gain can be no larger than G ¼ exp ð32þ, where again there is some arbitrariness in the choice of the value 32. For the case of an isolated gain or absorption resonance, of the sort described in figure 1, and for line centre operation, these conditions place a limit on the amount of advancement or retardation that can be achieved, and this limit is never more than several pulse widths. This result follows from equation (4). For L ¼ 32, we see that the normalized delay (or advance) can be no larger than approximately 2. This prediction is confirmed by the numerical results presented below. Let us next consider the case of a dip in a gain or absorption profile, the situation illustrated in figure 2. For the case of a gain medium, we find the same restriction as those given in the previous paragraph, because there is a limit to the maximum amount of gain even in the distant wings of the dip. But the situation is different for a dip in an absorption line. Here the attenuation in the wings of the line can be arbitrarily large, as long as the loss at line centre is negligible. In this case, there is no limit on how large the propagation distance can be, and hence there is no limit on how large the total delay can become. But it should be noted that this trick [23] works only for dips in absorption lines, not in gain lines. Thus, one can achieve arbitrarily large delay but not arbitrarily large advancement. (ii) Restrictions based on spectral reshaping of the pulse. Spectral reshaping of the light pulse is the dominant limiting effect in most slow- and fast-light systems. This process also behaves differently for slow- and fast-light systems, as we Input pulse T( ) Spectrally narrowed output pulse Spectrally broadened output pulse G( ) or Double-humped output pulse Figure 4. Influence of spectral reshaping of a pulse as it propagates through a slow- or fastlight system. Here T(!) is the transmission spectrum and G(!) is the gain spectrum of the material. (The colour version of this figure is included in the online version of the journal.)

Slow- and fast-light: fundamental limitations 2409 shall now see though use of figure 4. The spectrum of the input pulse is shown at the top of the figure. Under many practical situations, the spectral width of the pulse is comparable to the width of the transparency window. When this pulse propagates through a slow-light medium, the spectral wings of the pulse will experience lower transmission than the peak of the pulse. The pulse thus tends to become narrowed in frequency and hence broadened in time. This broadening is not desirable, but it is not as deleterious as what happens for the case of a fast-light medium, as illustrated near the bottom of the figure. In this case, the pulse broadens spectrally, and under certain conditions for long propagation distances the spectrum distorts so dramatically as to become multi-peaked. The pulse then shows ringing in the time domain. The onset of this severe form of pulse distortion determines the maximum possible extent of pulse advancement. 4. Results of numerical modelling We have also performed a numerical study of the propagation of optical pulses through various slow- and fast-light media. Our procedure entails decomposing the input pulse (which is assumed to be a transform-limited Gaussian pulse) into a Fourier integral and allowing each frequency component to propagate according to Eð!, zþ ¼Eð!,0Þ exp f½inð!þ!=c ð!þ=2šzg, ð5þ where nð!þ is the real part of the refractive index and ð!þ is the absorption coefficient, both evaluated at frequency!. These quantities are calculated from the Lorentz oscillator model for each of the situations shown in figures 1 and 2. These quantities thus obey causality (Kramers Kronig) requirements. Finally, the time evolution of the output pulse is determined through an inverse Fourier transform. The results of this study are presented in figure 5. The fast light results are calculated for line-centre operation with a Lorentzian absorption line with an overall transmission of exp ð 32Þ. The slow-light results are calculated for line-centre operation with a Lorentzian gain line with an overall transmission of exp ðþ32þ. The amount of advancement and delay were controlled by varying the width of the resonance. For the slow-light case, we see that the pulse tends to broaden as the amount of delay is increased. For a delay of more than four pulse widths (of the initial pulse width), the amount of pulse broadening becomes excessive. For the fast-light case, we see that serious pulse breakup occurs for an advancement of two pulse widths. 5. Summary and conclusions We have presented an analysis of the role of optical resonance in establishing slow and fast propagation of optical pulses. From the point of view of atomic response theory, there seems to be nearly complete symmetry between slow- and fast-light effects, and from this point of view it seems perhaps surprising that in practice it is

2410 R. W. Boyd and P. Narum Slow-Light Fast-Light 1 pulse-width delay Time 2 pulsewidths delay 4 pulsewidths delay 6 pulsewidths delay 1 pulse-width advance 2 pulse-widths advance 4 pulse-widths advance 6 pulse-widths advance Time vacuum Figure 5. Numercial results illustrating the properties of pulse propagation through slowand fast-light media. Note that pulse broadening occurs for the slow-light case, whereas pulse breakup occurs for the fast-light case when one attempts to advance the pulse more than approximately two pulse widths. (The colour version of this figure is included in the online version of the journal.) found to be easier to obtain large fractional delays than large fractional advances. We have identified two physical processes that lead to different behaviour of slowand fast-light media. One is that there is a maximum amount of attenuation that can be tolerated at the signal frequency for an absorbing medium, whereas there is a maximum amount of amplification that is allowable at any frequency for the case of a gain medium. The other process is spectral reshaping of the pulse, which also behaves differently for absorbing and amplifying media. We emphasize that the analysis of this paper is restricted to the situation in which the light is tuned to the centre of the resonance. Very large fractional delays have been observed in the far-detuned wings of absorption lines [11]. Acknowledgements We gratefully acknowledge valuable discussions with George Gehring, Giovanni Piredda, Aaron Schweinsberg, Katie Schwertz, Zhimin Shi, Heedeuk Shin, Joseph Vornehm, and Petros Zerom. This work was supported by the DARPA/DSO Slow Light Program and by the NSF.

Slow- and fast-light: fundamental limitations 2411 References [1] R.W. Boyd and D.J. Gauthier, in Progress in Optics, Vol. 43, edited by E. Wolf (Elsevier, Amsterdam, 2002). [2] P.W. Milonni, Fast Light, Slow Light, and Left-Handed Light (Institute of Physics Publishing, Bristol, 2005). [3] R.W. Boyd, D.J. Gauthier and A.L. Gaeta, Opt. Photon. News April 18 2006. [4] D.J. Gauthier, A.L. Gaeta and R.W. Boyd, Photon. Spectra March 45 (2006). [5] S.E. Harris, J.E. Field and A. Kasapi, Phys. Rev. A 46 R29, (1992). [6] A. Kasapi, M. Jain, G.Y. Yin, et al., Phys. Rev. Lett. 74 2447 (1995). [7] M.M. Kash, V.A. Sautenkov, A.S. Zibrov, et al., Phys. Rev. Lett. 82 5229 (1999). [8] D. Budker, D.F. Kimball, S.M. Rochester, et al., Phys. Rev. Lett. 83 1767 (1999) [9] L.V. Hau, S.E. Harris, Z. Dutton, et al., Nature 397 594 (1999). [10] M.S. Bigelow, N.N. Lepeshkin and R.W. Boyd, Phys. Rev. Lett. 90 113903 (2003). [11] R.M. Camacho, M.V. Pack and J.C. Howell, Phys. Rev. A 74 033801 (2006). [12] M.D. Stenner, M.A. Neifeld, Z. Zhu, et al., Opt. Express 13 9995 (2005). [13] S. Chu and S. Wong, Phys. Rev. Lett. 48 738 (1982). [14] L.J. Wang, A. Kuzmich and A. Dogariu, Nature 406 277 (2000). [15] M.D. Stenner, D.J. Gauthier and M.A. Neifeld, Nature 425 695 (2003). [16] B. Segard and B. Macke, Phys. Lett. A 109 213 (1985). [17] B. Macke and B. Segard, Eur. Phys. J. D 23 125 (2003). [18] G.M. Gehring, A. Schweinsberg, C. Barsi, et al., Science 312 985 (2006). [19] M.S. Bigelow, N.N. Lepeshkin and R.W. Boyd, Science 301 200 (2003). [20] E.E. Mikhailov, V.A. Sautenkov, I. Novikova, et al., Phys. Rev. A 69 063808 (2004). [21] A. Kuzmich, A. Dogariu, L.J. Wang, et al., Phys. Rev. Lett. 86 3925 (2001). [22] H. Cao, A. Dogariu and L.J. Wang, IEEE J. Select. Topics Quantum Electron. 9 52 (2003). [23] R.W. Boyd, D.J. Gauthier, A.L. Gaeta, et al., Phys. Rev. A 71 023801 (2005); ibid. 72 059903(E) (2005). [24] A. Matsko, D. Strekalov and L. Maleki, Opt. Express 13 2210 (2005). [25] J.B. Khurgin, Opt. Lett. 31 948 (2006). [26] M.S. Bigelow, N.N. Lepeshkin, H. Shin, et al., J. Phys. Condens. Matter 18 3117 (2006).