13.07.2015 Views

Reeta Vyas - Physics - University of Arkansas

Reeta Vyas - Physics - University of Arkansas

Reeta Vyas - Physics - University of Arkansas

SHOW MORE
SHOW LESS
  • No tags were found...

You also want an ePaper? Increase the reach of your titles

YUMPU automatically turns print PDFs into web optimized ePapers that Google loves.

PHYSICAL REVIEW A 76, 052317 (2007)Generation and evolution <strong>of</strong> entanglement in coupled quantum dots interactingwith a quantized cavity fieldAmab Mitra and <strong>Reeta</strong> <strong>Vyas</strong>Department <strong>of</strong> <strong>Physics</strong>, <strong>University</strong> <strong>of</strong> <strong>Arkansas</strong>, Fayetteville, Arkarlsas 72701. USADaniel ErensoDepartment <strong>of</strong> <strong>Physics</strong> & Astmnomv, Middle Tennessee State <strong>University</strong>, Murfeesboro, Tennessee 37132, USA(Received 14 August 2007; published 20 November 2007)The generation <strong>of</strong> entanglement between two identical, interacting quantum dots-initially in groundstates-by a coherent field and the subsequent time evolution <strong>of</strong> the entanglement are studied by calculatingthe concurrence between the two dots. The results predict that while it is possible to generate entanglement (orentanglement <strong>of</strong> formation, as defined for a mixed state) between the two dots: at no time do the dots becomcfully entangled to each other or is a maximally entangled Bell state ever achieved. We also observe that thedegree <strong>of</strong> entanglement increases with an increase in the photon number inside the cavity and a decrease in thedot-photon coupling. The behavior <strong>of</strong> the two-dot system, initially prepared in an entangled state and interactingwith thermal light, is also studied.DOI: 10.llO3/PhysRevA.76.0523 17 PACS number(s): 03.67.Mn, 78.67.Hc, 03.65.Ud. 42.55.SaI. INTRODUCTIONThe coherent evolution <strong>of</strong> two quantum bits (qbits) in anentangled state <strong>of</strong> the Bell type is at the heart <strong>of</strong> the study <strong>of</strong>quantum entanglement and is fundamental to both quantumcryptography [l] and quantum teleportation [2]. Differentsystems and methods for the preparation and measurement <strong>of</strong>maximally entangled states are being investigated intensively.Many <strong>of</strong> these investigations have centered on atomicand quantum-optical systems [3-51. However, generation <strong>of</strong>entangled states in solid-state systems such as semiconductorquantum dots (QDs) has also received a lot <strong>of</strong> attention[6-91. Recent experiments have shown that QDs can showmany interesting quantum features such as entanglement[lo], photon antibunching [I 11, Rabi oscillations [12], andmodification <strong>of</strong> spontaneous emission decay rate [13].In our present work, we study the generation and evolution<strong>of</strong> entanglement between two coupled, identical quantumdots interacting with a quantized cavity field. Similarstudies have been reported by several previous authors.Quiroga and Johnson [6,7] have shown that two andthree equally coupled QDs, interacting with a suitably chosenfield, can generate maximally entangled Bell andGreenberger-Home-Zeilinger (GHZ) states respectively.However, they used a classical field instead <strong>of</strong> a quantizedone and measured the degree <strong>of</strong> entanglement between thedots by calculating the overlap <strong>of</strong> the state vector with acompletely entangled state. Yi et al. [8] extended this workfor two coupled QDs interacting with a quantized coherentfield; however, they do not give any explicit expression forthe entanglement between the dots as well, rather take theratio <strong>of</strong> the probabilities <strong>of</strong> the QDs population being foundin two fully entangled Bell states as an indirect measure <strong>of</strong>entanglement. Wang et al. [9] have considered a fully quantizedfield, and they have also given an exact mathematicalexpression for the entanglement between the two dots, butthey did not take into consideration any interaction betweenthe dots. In our work we consider the QDs to be coupled toeach other, taking into account the possibility <strong>of</strong> their mutualinteraction, and we measure the degree <strong>of</strong> entanglement betweenthe two dots by calculating the concurrence betweenthe two dots, after taking partial trace over the field states.We assume the dots to be initially in the ground state andstudy the creation <strong>of</strong> entanglement as a result <strong>of</strong> the interactionwith the coherent cavity field and between thcn~selves.Several interesting results come out which diflcs from thqpredictions where the field is treated as classical [h,7] or thedot-dot interaction is neglected [9] or entanglemen( is calculatedonly qualitatively by finding the probability <strong>of</strong> the systemstate to be in a Bell state [6-81. Wc ti:.: t1:it the dctscould never be made fully entangled, even when we negl:i:tall kinds <strong>of</strong> losses, and in general the entanglemen! functionis not periodic in time. We also study the effect <strong>of</strong> the fieidstrength and the relative strengths <strong>of</strong> the coupling parameterson entanglement. Next we consider the interaction with ther- ,ma1 light and obtain an interesting observatio~i that if anentangled state is left to interact with the thermal light; itreduces the amount <strong>of</strong> entanglement, but does no1 completelydestroy it.The paper is organized as follows. Section I1 sutnmarizesthe model Hamiltonian describing N coupled QDs interactingwith a quantized electromagnetic field. In Sec. Ill wc use thisHarniltonian for two QDs to determine the state vector <strong>of</strong> thecoupled QD-field system. We then study the rvolution o'fentanglement for the cavity field, initially in coherent ari'dthermal states. Section IV summarizes the main ~.csults cii'this paper.U. TWO COUPLED QDs INTERACTING ,.WITH A QUANTIZED CAVITY kit,%. 2We consider N identical semiconductor quantym dots tha:are equally coupled to each other via Coulombic interaction.The QDs interact with a quantized field (dipole interactiod)in a high-Q cavity. Then the coupled QD-field system is .described by the Hamiltonian [6]1050-2947/2007n6(5)/05231 7(7) 052317- 1 02007 The American Physical socle,t;


Conditional homodyne detection <strong>of</strong> light with squeezed quadrature fluctuationsJustin Vines, Recta <strong>Vyas</strong>, and Surcndra Singhllepartrne~it o/'P/ysirs, Universiiy <strong>of</strong> Arkanscls, Fayettevrlle, Arkclnsa,~ 72701, USA(Received 16 June 2006; published 24 August 2006)We discuss the detection <strong>of</strong> field quadrature fluctuations in conditional homodyne detection experiments andpossible sources <strong>of</strong> error in such an experiment. We also present modifications to these expeiiments to helpeliminate such errors and extend their range <strong>of</strong> applicability.I. INTRODUCTIONFluctuations <strong>of</strong> light provide a window on the underlyingquantum dynamical evolution <strong>of</strong> a light-emitting source.Einission <strong>of</strong> a photon by a light source signals quantumfluctuations in progress, and a measurement that is conditionedon a photodetection allows us to study the timeevolution <strong>of</strong> the fluctuations [I]. For example, measurement<strong>of</strong> light intensity conditioned on a photodetection, alsoknown as two-time intensity correlation, reveals informationregarding bunching and antibunching that is not available inunconditional intensity measurements [2-61.The conditional measurement <strong>of</strong> quadrature fluctuations(CMQF) proposed by Cannichael er ul. [7] reveals in a novelway the nonclassical nature <strong>of</strong> light from a cavity containingtwo-level atoms; this has been experimentally observed byFoster et al. [XI. Nonclassical effects in conditional intensityand squeezing in light from a degenerate parametric oscillatorhave been studied [6]. The conventioilal methods <strong>of</strong> dctectingquadrature squeezing involve unconditional measurementsthat are degraded by detection inefficiencies and donot explore the time evolution <strong>of</strong> quadrature fluctuations[1,2,9,10]. The CMQF, on the other hand, is essentially independent<strong>of</strong> detection efficiency and provides a sensitiveprobe <strong>of</strong> the fluctuations' development over time. It has beenshown that the conditional measurement can reveal remarkablenonclassical behavior <strong>of</strong> not only squeezed but also <strong>of</strong>unsqueezed quadrature fluctuations [6-81.111 Sec. 11, we briefly summarize-the theoretical conceptsunderlying a CMQF experiment [6,7]. We then consider intracavitysecond harmonic generation (ISHG) [ll-131 andpresent a theoretical analysis <strong>of</strong> the nonclassical features <strong>of</strong>ISHG quadrature fluctuations.The CMQF technique achieves a measurement <strong>of</strong> thequadrature fluctuations <strong>of</strong> a given source ficld by crosscorrelalinga photon count with a balanced homodyne detection.This technique requires the use <strong>of</strong> auxiliary coherentoscillators (coherent laser sources), and the amplitudesor intensities <strong>of</strong> the coherent oscillator fields must be setto values that depend on the properties <strong>of</strong> the source field. Itcan be shown that the accuracy <strong>of</strong> the measurement'sfinal results is extremely sensitive to the precision <strong>of</strong> theseadjustments. In some cases, a very small error in theseadjustments may give rise to incorrect conclusions about thestate <strong>of</strong> the source field's quadrature fluctuations. We demonstratethis in Sec. III by developing a theoretical model <strong>of</strong>such an crror in the CMQF mcasurement and exploring itseffccts on conclusions that might be drawn about the state <strong>of</strong>the ISHG field.Weak fields with approximately Gaussian fluctuations areideal candidates for accurate quadrature fluctuation measurementsusing the CMQF method, provided that coherent fieldadjustments can be made sufficiently precise. However, thetechnique is limited in its applicability to a source field withnonzero third-order moments [I 4,351, which may obscurethe results <strong>of</strong> the mcasurement, even with perfect control <strong>of</strong>coherent light parameters. In Sec. IV, we propose an extension<strong>of</strong> the CMQF method that can achieve a measurement<strong>of</strong> the quadrature fluctuations <strong>of</strong> a completely generic sourcefield while eliminating the effccts <strong>of</strong> third-order fluctuationmoments. Additionally, this extended CMQF measurementavoids the need for precise adjustments <strong>of</strong> coherent laserfields, thus perhaps averting the kinds <strong>of</strong> error discussed inSec. III. In Sec. V, we summarize our findings.II. CONDITIONAL MEASUREMENT OF QUADRATIJREFLUCTUATIONS FOR LSHGThe quadrature variables for an optical ficld with annihilationand creation operators 6, and 6: are defined bywhere C#J is an arbitrary phase [2]. It follows from thisquadrature definition that the variances (:(Aig)?) and(:(A?,)':) are related lo (he intensity <strong>of</strong> the field fluctuations(A~~AG,) bywhere colons denote time and normal ordering <strong>of</strong> theoperators enclosed by them. For classical fields, both quadraturevariances are always greater than or equal to zero,being equal to zero only in the classical coherent state.For quantum fields, however, the nornlally ordered variance<strong>of</strong> a quadrature Xg can become negative as long as the normallyordered variance <strong>of</strong> ?@ increases in such a way thatEq. (3) is still satisfied. In such a case, thc quadrature i4 issaid to be squeezed and the field 6 is said to be in a squeezedstate [2]. This fact and the fact that the fluctuation intensity isnonnegative lead to the inequality [6,7]1050-2947/2006/74(2)/023817(7) 023817-1 02006 The American Physical Society


Eur. Phys. J. D 29, 95-103 (2004)DOI: 10.1140/epjd/e2004-00018-2On tlie bichromatic excitation <strong>of</strong> a two-level atomwith squeezed lightAmitabh Joshia, <strong>Reeta</strong> <strong>Vyas</strong>, Surendra Singh, and Min XiaoDepartment <strong>of</strong> <strong>Physics</strong>, <strong>University</strong> <strong>of</strong> <strong>Arkansas</strong>, Fayetteville AR 72701, USAReceived 3 September 2003 / Received in final form 11 November 2003Published online 17 February 2004 - @ EDP Sciences, Societk Italiana di Fisica, Springer-Verlag 2004Abstract. Analytical results for the dynamical evolution <strong>of</strong> a single two-level atom coupled to an ordinaryheat bath and bichromatically excited with two finite bandwidth squeezed fields are presented bysolving the Heisenberg equations <strong>of</strong> motion. Photon statistics <strong>of</strong> the system in terms <strong>of</strong> the second-orderintensity-intensity correlation function are also discussed. Transient fluorescent intensity as well as intensitycorrelation function exhibit oscillatory phenomenon even in the weak field limit. The latter also showsenhanced delayed bunching effect. All these effects are sensitive to the bandwidth <strong>of</strong> squeezed light.PACS. 42.50.Ct Quantum description <strong>of</strong> interaction <strong>of</strong> light and matter; related experiments -42.50.D~ Nonclassical states <strong>of</strong> the electromagnetic field, including entangled photon states; quantum stateengineering and measurements1 IntroductionStudies related to the dynamical evolution, the photonstatistics, spectral and radiative properties <strong>of</strong> one andmanv two-level atoms embedded in a broadband saueezedbath have been the topics <strong>of</strong> keen interest in quantumoptics. Gardiner [I] studied the interaction <strong>of</strong> a two-levelatom with a broadband squeezed bath and predicted unequalpolarization quadrature-decay rates. Carmichael,Lane and Walls [2] were able to discover a significantphenomenon <strong>of</strong> sub-natural linewidth in the fluorescencespectrum <strong>of</strong> a driven two-level atom, in the presence<strong>of</strong> squeezed light. Atomic absorption spectrum has discussedby Ritsch and Zoller [3] in the presence <strong>of</strong> coloredsqueezed vacuum. Since then many interesting results inatom-squeezed field interaction have been reported whichinclude both two and three-level atoms interacting withbroad bandwidth or narrow bandwidth squeezed baths [4,51. In a recent study the interaction <strong>of</strong> a two-level atomwith the squeezed vacuum <strong>of</strong> bandwidth smaller than thenatural atomic linewidth was considered and the holeburning and the three-peaked structure in spectra <strong>of</strong> fluorescenceand transmitted field were predicted [6]. Theseresults essentially show that squeezed fields having pairwisecorrelations and anisotropic noise distribution cangive rise to interesting phenomena including novel featuresin spectral properties <strong>of</strong> atoms, formation <strong>of</strong> purestates and photon statistics. A more realistic model <strong>of</strong> finitebandwidth squeezed light interacting with a singletwo-level atom has been studied by <strong>Vyas</strong> and Singh (71and Lyublinskaya and <strong>Vyas</strong> [8] where the source <strong>of</strong> thea e-mail: ajoshiouark. edusqueezed light employed was a degenerated parametricoscillator (DPO) operating below threshold and a homodynedDPO [9]. In another work, a two-level atom insidean optical parametric oscillator has been considered andhole and dips in the fluorescence and transmitted lighthas been observed [lo]. The interest in other sources <strong>of</strong>squeezed light as well as its applications in a wide variety<strong>of</strong> areas has continued unabated [ll-131.The interaction <strong>of</strong> a single two-level atom with bichromaticdriving field has also been studied extensively boththeoretically and experimentally [14]. These studies weremotivated by the observations that the bichromatic nature<strong>of</strong> the driving field can lead to a number <strong>of</strong> novel featureswhich are different from the monochromatic case. For example,the fluorescence intensity exhibits resonances atsubharmonics <strong>of</strong> the Rabi frequency and different spec-tral characteristics when com~ared with the usual Mollowtriplet. Recently, some new calculations for resonance fluorescenceand absorption spectra <strong>of</strong> a two-level atom drivenby bichromatic field have been reported [15]. Also, reportedare the effects <strong>of</strong> broadband squeezed reservoir onthe second order intensity correlation function and squeezingin the resonance fluorescence for a bichromaticallydriven two-level atom [16]. Coherent population trappingand Sisyphus cooling under bichromatic illumination havealso been studied [17]. In another recent work the electromagneticallyinduced transparency (which normally occursin three-level atoms) has been demonstrated in a twolevelatom excited by a bichromatic field (one strong andone weak field) and possibility <strong>of</strong> squeezed-light generationhas also been discussed 1181. . .In this work, we study the interaction <strong>of</strong> a single twolevelatom with a bichromatic electromagnetic field that is


PHYSICAL REVIEW A 67, 063808 (2003)Conditional measurements as probes <strong>of</strong> quantum dynamicsShabnanl Siddiqui, Daniel Erenso, <strong>Reeta</strong> <strong>Vyas</strong>, and Surcndra SinghDepurtmet~t q/'Pl~):sics. Universii): <strong>of</strong>At.kutans, &vetfe\~ille, Arkamcus 72701, USA(Rcccivcd 5 December 2002; published 17 June 2003)Wc discuss conditional rneasuremcnts as probcs <strong>of</strong> quantum dynamics and show that thcy provide differentways to characterize quantum fluctuations. \Ve illustrate this by considering thc light from a subthresholddcgcncratc parametric oscillator. Analytic rcsults and curvcs arc prcscntcd to illustrate thc behavior.DOI: 10.1103/PhysRcvA.67.063808PACS numbcr(s): 42.50.Dv, 42.50.Ar, 42.65.K~1. INTRODUCTIONFluctuations are an integral feature <strong>of</strong> quantum dynamicalevolution. For dissipative quantum optical sources, thesefluctuations are reflected in the photoemissions fYom thesesources. In this paper, we will use the terms photoemissionsand photodetections interchangeably because photodetectionis sin~ply a matter <strong>of</strong> casting a net to catch the photons emittedby the source. A photoemission signals a fluctuation inprogress. Hence, a conditional measurement that commenceswhen a photoenlission has occurred catches a fluctuation inthe act and, in fact, allows us to observe the time evolution<strong>of</strong> the fluctuation [I]. This sort <strong>of</strong> information is not availablein unconditioned measurements. Thus, conditional measurenientsallow us to probe quantum dynamics at a deeperlevel. Perhaps the best known example <strong>of</strong> conditional measurementis the measurement <strong>of</strong> the second-order intensitycorrelation that measures fluctuations <strong>of</strong> light intensity followinga conditioning photodetcction [2]. This is the basis <strong>of</strong>the observation <strong>of</strong> photon bunching or antibunching. Recently,measurements <strong>of</strong> quadrature squeezing conditionedon a photodetection have also been proposed and reported incavity quantum electrodynamics [3,4]. In this papel; we considerconditional measurements in the context <strong>of</strong> parametricoscillators and show that such measurements provide sensitiveprobes <strong>of</strong> quantum dynamics and the language <strong>of</strong> conditionalmeasurements provides powerhl conceptual toolsfor unraveling and understanding nonclassical features <strong>of</strong>quantum dynamics [5- 81.Optical parametric oscillators (OPOs) and an~plifiers [9]based on down-conversion have played a central role in thestudies <strong>of</strong> nonclassical photon correlations and variousschemes for quantum coinmunication and computation [lo].The fitndamental process in optical parametric oscillators isconversion <strong>of</strong> a pump photon <strong>of</strong> frequency w,, into a pair <strong>of</strong>photons (signal and idler) <strong>of</strong> lower frequencies w, and wi ina nonlinear medium inside an optical cavity. The processconserves energy and momentum. Conservation <strong>of</strong> energyrequires the pump and down-converted frequencies to be relatedby wp= us+ wi and conservation <strong>of</strong> momentum, alsoknown as phase matching, requires the use <strong>of</strong> certain anisotropicmaterial media and specific states <strong>of</strong> polarization forthe pump, signal, and idler photons. If the down-convertedphotons have the same frequency (ws= wi=wd), polarization,and direction <strong>of</strong> propagation, the process is called degenerateotherwise it is called nondegenerate. In the fonnercase, we speak <strong>of</strong> a degenerate parametric oscillator (DPO)and in the latter <strong>of</strong> a nondegeneratc parametric oscillator.We begin by considering a degenerate paranlet~ic oscillatorin Sec. I1 and discuss conditional measurements <strong>of</strong> intensityand amplitude. Conditional measurements <strong>of</strong> quadraturefluctuations are discussed in Sec. 111. We restrict our considerationsto its operation below the threshold <strong>of</strong> sustainedoscillations. This allows us to obtain simple analytic cxpressionsfor various quantities <strong>of</strong> interest. The results are summarizedin Sec. IV.11. CONDITIONAL MEASUREMENTS OF A DPOThe field from the DPO is governed by the interactionHamiltonian for phase-matched down-conversion inside anoptical cavity driven by a classical injected signal [5,6]. Theequation <strong>of</strong> motion for the density operator bd <strong>of</strong> the DPOfield is thenwhere K is the mode-coupling constant, E is the dimensionlessamplitude <strong>of</strong> the classical pump field, y is the cavitylinewidth, and id and 6; are the annihilation and creationoperators for the DPO. In writing the equation <strong>of</strong> motion forthe density matrix, we have neglected pump depletion that isa reasonable approxin~ation for low down-conversion efficienciesand subthreshold operation <strong>of</strong> the DPO consideredhere. The combination KE has been chosen to be real by asuitable definition <strong>of</strong> the phases <strong>of</strong> id and 61.Using the positive-P representation for the density matrix,we can map the equation <strong>of</strong> motion for the annihilation andcreation operators for the DPO field onto a set <strong>of</strong> stochasticequations for the c-number amplitudes cud and ad, corresponding,respectively, to the annihilation and creation operators6 and Cid. These equations read [5,6]where [,it) and &(t) are two statistically independent realGaussian white-noise processes with zero means and unitintensity. Nonnally ordered averages <strong>of</strong> ri$ and rid can thenbe calculated by using the mapping1 0S0-2947/2003/67(6)/063808(9)/$20.00 67 063808-1 Q2003 The American Physical Society


PHYSICAL REVIEW A 67, 013818 (2003)Quantum well in a microcavity with injected squeezed vacuumDaniel Brenso, Rceta <strong>Vyas</strong>, and Surcndra SinghI)epczrtmenI <strong>of</strong> Ph.vsics, Universiw <strong>of</strong> Ar/can.sas, Fayetteville, <strong>Arkansas</strong> 72 701(Received 27 August 2002; published 3 1 January 2003)A quantum wcll with a singlc cxciton mode in a microcavity driven by squcczed vacuum is studicd in thelow exciton density regime. By solving the quantum Langevin equations, we study the intensity, spectrum, andintensity correlation function for the fluorcsccnt light. An cxprcssion for thc Q function <strong>of</strong> thc ficld insidc thccavity is derived from the solutions <strong>of</strong> the quantum Langevin cquations. Using the Q function, the intracavityphoton number distribution and thc quadrature fluctuations for both thc cavity and fluorcsccnt fields arcstudied. Several interesting and new el'fects due to squeezed vacuum are found.DOI: 10.1 103/PhysRevA.67.013818PACS number(s): 42.55.Sa. 78.67.De, 42.50.Dv, 42.50.L~1. INTRODUCTIONWith the development <strong>of</strong> semicond~ictor optical microcavities,there has been considerable interest in excitoncavitycoupled systems [1,2]. These systcms have revealedsome interesting phenomena that are similar to those observedin the interaction <strong>of</strong> a two-level atom with light [3-91.The exciton-cavity system gives rise to the so-called polaritons,which are the no~nlal inodes <strong>of</strong> a coupled excitonphotonsystem. The excitation spectrum <strong>of</strong> the compositeexciton-cavity system is characterized by two well-resolvedpolariton resonances (or nornlal mode resonances) when g>(ye, yc), where g is the dipole coupling between the excitonand the cavity mode, and ye and y, are the exciton andcavity mode damping rates, respectively. In this limit, anexcitation <strong>of</strong> the cavity mode can lead to a coherent oscillatoryenergy exchange (or normal mode oscillation) betweenthe exciton and the cavity due to the vacuum Rabi oscillations.The vacuum Rabi oscillations in a coupled exciton-photonsystem in sen~iconductor microcavity lasers have been observedby Weisbuch et al. [2]. Following this observation,extcnsive experimental and theoretical studics have been carriedout [lo- 171. These studies have confirmed normal modesplitting and oscillatory emission from exciton microcavities.Theoretical investigations in the linear regime, where the excitonscan be approximated as bosons, have bccn carricd outby Pa11 et 01. [I 51. Wang et al. [I 61 investigated the effects <strong>of</strong>inhomogeneous broadening <strong>of</strong> excitons on normal mode oscillationsin semiconductor microcavities using the coupledoscillator model. Their results show that inhomogen~ousbroadening can drastically alter the coherent oscillatory energyexchange process cven in regimes whcre normal modesplitting remains nearly unchanged.In this paper, we study the excitonic system in a microcavitywhere the cavity is driven by squeezed vacuum. Anoutline <strong>of</strong> the system is shown in Fig. 1. A sen~iconductorquantum well is embedded between two Bragg reflectingmirrors. One <strong>of</strong> these mirrors acts as an input port throughwhich light in a squeezed vacuum state is injected into thecavity. We includc dissipation <strong>of</strong> both thc cavity and excitonmodes. In Sec. 11, we derive the quantum Langevin equationsfor the exciton and cavity modcs. We solve these equationfor the case in which the damping constants are equal. These~.csults are used in Scc. I11 to study the cffects <strong>of</strong> initial cavityphoton number as wcll as squeezed-vacuum photon numberon the intensity, spectrum, and the second-order intensityco~relation <strong>of</strong> the fluorescent light. In Sec. IV, we obtain theQ-distribution function and use it to study the intracavityphoton number distribution and squeezing <strong>of</strong> thc cavitymode and the fluorescent light. We summarize the principalresults <strong>of</strong> the paper in Sec. V.11. QUANTUM LANGEVIN EQUATIONWe consider a semiconductor quantum well fQW) in thelinear excitation regime where the density <strong>of</strong> excitons issmall so that exciton-exciton interaction can be ignored. Theexcitons can then be approximated as a dilute boson gas [IS].In this approximation, the microscopic Hamiltonian in theinteraction picture describing the exciton-cavity system isgiven by [17,19]The Hamiltonian <strong>of</strong> Eq. (1) is written in the rotating-waveapproximation and in the dipole approximation. Here and6 are the annihilation operators for the cavity and excitonmodes, respcctivcly, in a frame rotating at frequencyw, , f, (I? ,) is the reservoir operator responsible for cavityfield (exciton) damping, g is the coupling constant characterizingthe strength <strong>of</strong> interaction between the exciton and thecavity field, and detuning A w = (we- w,), where w, and w,.are the frequencies <strong>of</strong> the exciton and cavity modes, respectively.Normally, the exciton and cavity inodcs are coupled toa continuum <strong>of</strong> thennal reservoir modes. This leads to theirMirro@fl-MirrorpmpFIG. 1. An outline <strong>of</strong> the physical system.1050-2947/2003/67(1)/0138 18(10)/$20.00 67 013818-1 Q2003 The American Physical Society


PHYSICAL REVIEW A, VOLUME 65, 063808Two-level atom coupled to a squeezed vacuum inside a coherently driven cavityDaniel Erenso and <strong>Reeta</strong> <strong>Vyas</strong>Department <strong>of</strong> Phj~sics, Univer:~ir), <strong>of</strong> <strong>Arkansas</strong>. F'czyetteville, Arkaizsns 72701(Rcceivcd 8 Septembcr 2001 ; published 5 June 2002)A single two-lcvcl atom in a cohcrcntly driven cavity and damped by a broadband squcczcd vacuumcentered about the atomic transition liequency is studied. A second-order Fokker-Planck equation for thissystem is obtained without using system size cxpansion. Effects <strong>of</strong> dctunings and cavity decays arc alsoincorporated in the Fokker-Planck equation. This equation is used lo study atomic inversion, fluorescentspectrum, and the intcnsity corrclations <strong>of</strong> the transmitted and fluorcsccnt photons in thc bad-cavity limit.Several interesting effects in the atomic inversion, spectrum, and intensity correlations due to the squeezedvacuum arc presented. Thesc results are also compared with an atom that is dampcd by a thermal rcscrvoir.DOI: 10.1103/PhysKcvA.65.063808PACS nu~nbcr(s): 42.50.Lc, 42.50.Ct, 42.50.Dv, 42.50.ArI. INTRODUCTIONWith the gcncration and dctcction <strong>of</strong> sqiieczed light [I-41increasing attention is being given to the study <strong>of</strong> interaction<strong>of</strong> squcezed light with a two-level atoin [5-131. Gardinerstudied a single two-level atom embedded in a broadbandsqueezed vacuum and showcd that tlie two quadratures <strong>of</strong> thcatomic polarization decay at two distinct decay rates that aresensitive to the phase correlations <strong>of</strong> the squcezed vacuum[IS]. Carmichael rt al. [6] studied the fluorescent spectrum <strong>of</strong>an aton1 immersed in a broadband squeezed vacuum whenthe atom is driven by a coherent field. They predicted that forwcak driving fields the incoherent spectrum would narrow asthe amount <strong>of</strong> squeezing is increased. In this linlit the spectminis insensitive to the relative phase between tlie drivingfield and the squeezed vacuum. For strong driving fields, onthe other hand, the central pcak <strong>of</strong> the Mollow spectrum canbroaden or narrow, depending on the relative phase betweenthe squcezed vacuum and the driving field. The photon numberdistribution for this system has been calculated by Jagatapand Lawande [7]. <strong>Vyas</strong> and Singh considered resonancefluorescence in the weak-field limit when the atoin isdriven by squeezed liglit from an optical parametric oscillator[8], Lyublinskaya and <strong>Vyas</strong> considered when the atom isdriven by nonclassical light from intracavity second harmonicgeneration and a homodyne degenerate parametric oscillator[9].Parkins and Gardiner [lo] considered a single two-levelatom in a cavity whcn thc squeezed light is incidcnt uponone <strong>of</strong> the output mirrors. Rice and Pedrotti [I I.] placed atwo-level atom coupled to an ordinary vacuum inside a coherentlydriven optical cavity coupled to a broadbandsqueezcd reservoir through the output mirror. They foundthat it was possible to overcoine the cavity enhancement part<strong>of</strong> tlie linewidth. This work was extended by Rice and Baird[12] to calculate the second order intensity co~~elation functiong(2)(7-) and spectra <strong>of</strong> the fluorescent light.In this paper we study a single two-level atom placedinside a coherently driven cavity where the atom is directlycoupled to a squeezed vacuum but the cavity mode decaysinto an ordinary vacuum. The squeezed vacuum spectrum isconsidcrcd to bc broadband ccntcred at the atomic transitionfrequency. The model proposed here can be implemented byassuming a short cavity, which subtends a largc solid anglc atthe atom. Squeezed modes <strong>of</strong> the short cavity are directlycoupled to the atom. Another cavity with its axis perpendicularto the short cavity is driven by a coherent driving field.We derive an exact Fokker-Planck equation following theapproach <strong>of</strong> Wang and <strong>Vyas</strong> [14-161. An appealing feature<strong>of</strong> the Fokker-Planck equation approach is that it allowsquantum-operator averages to be calculated as classical-likeaverages. Thus analogies between classical and quantumfluctuations can be drawn that help in developing an intuitivefeeling for quantum fluctuations. The Folcker-~lanck equationcan be converted into a set <strong>of</strong> stochastic differentialequations which can be solved numerically and in manycases analytically. We study the effects <strong>of</strong> a squeezedvacuum on atomic inversion, the fluorescent spectrum, andthe second-order intensity col~elation function <strong>of</strong> the transmittedand fluorescent light in the bad-cavity limit. We findthat in the presence <strong>of</strong> squeezing the threshold value <strong>of</strong> thecooperativity parameter for seeing vacuum Rabi splitting canbe lowered. This bchavior is phase sensitivc and cannot beseen if squeezed light is replaced by thermal light. We findthat the transmitted light can show antibunching even for alarge cooperativity parameter. We explain the behavior <strong>of</strong>antibunching in ternls <strong>of</strong> self-hoinodying <strong>of</strong> coherent andincoherent components. We also find that, for large values <strong>of</strong>squeezing, antibunching results due to a reduction in the intensityfluctuations <strong>of</strong> the incoherent component. This differsfrom the case for small squeezing. For sinall squeezing antibunchingresults from an interference <strong>of</strong> the coherent componentwith the incoherent component.Squeezed VacuumFIG. 1. Physical scheme.1050-2947/2002/65(6)/063808( 1 1)/$20.00 65 063808-1 02002 The American Physical Society


Erenso et al. Vol. 19, No. 6IJune 2002lJ. Opt. Soc. Am. B 1471Higher-order sub-Poissonian photon statisticsin terms <strong>of</strong> factorial momentsDaniel Erenso, <strong>Reeta</strong> <strong>Vyas</strong>, and Surendra SinghDepclrtment <strong>of</strong> <strong>Physics</strong>, <strong>University</strong> <strong>of</strong> <strong>Arkansas</strong>, Fayetteville, <strong>Arkansas</strong> 72701Received September 10, 2001We introduce the concept <strong>of</strong> higher-order super-Poissonian and sub-Poissonian statistics and show that higherordersub-Poissonian statistics is a signature <strong>of</strong> a nonclassical field. Fields generated in intracavity secondharmonicgeneration and single-atom resonance fluorescence are shown to exhibit higher-order sub-Poissonianstatistics. 0 2002 Optical Society <strong>of</strong>AmericaOCIS codes: 270.0270, 270.5290, 190.4970, 270.2500.1. INTRODUCTIONNonclassical properties <strong>of</strong> electromagnetic fields receive agreat deal <strong>of</strong> attention, as these properties provide a testingground for the predictions <strong>of</strong> quantum electrodynamics.'~'Sub-Poissonian Photon statistics based onsecond-order intensity correlations provide one way tocharacterize the nonclassical nature <strong>of</strong> a light beam.3.4Nonclassical effects as they relate to higher-order momentshave also been discussed in the literat~re.~-'Aganval and Tara discussed higher-order nonclassical effectsfor single-mode fields in terms <strong>of</strong> normally orderedmoments."erina and co-workers studied noilclassicalbehavior in optical parametric processes, Raman andBrillouin scattering6 and nonlinear optical couplers7 interms <strong>of</strong> moments <strong>of</strong> integrated intensity. Lee consideredsingle-mode fields and defined higher-order noi~classicaleffects in terms <strong>of</strong> factorial moments <strong>of</strong> the photon distributionby using majorization theory.8 <strong>Vyas</strong> and Singhconsidered higher-order nonclassical effects in terms <strong>of</strong>factorial moments <strong>of</strong> the photocount distribution.' Inthis paper we introduce criteria for evaluating higherordersub-Poissonian statistics in terms <strong>of</strong> factorial momentsand show that higher-order sub-Poissonian statisticsare indicative <strong>of</strong> nonclassical fields. The criteria thatwe introduce are independent <strong>of</strong> the efficiency <strong>of</strong> detection.We then show that the light from intracavitysecond-harmonic generation (ISHG) and light fromsingle-atom resonance fluorescence exhibit higher-ordersub-Poissonian statistics.2. HIGHER-ORDER SUB-POISSONIANSTATISTICSFor second-order sub-Poissonian statistics variance((Am)') = (m2)- (m)' <strong>of</strong> the photon-counting distributionis less than the meail <strong>of</strong> the distribution, (m). Bynoting that second-order factorial moment (m(")= (m(m - 1)) - (m') - (m), we can write the criterionfor the second order sub-Poisson statistics asNote that for a Poissonian distribution (m(") = (m)', soinequality (1) holds as an equality. The departures fromPoisson statistics are then characterized in terms <strong>of</strong> theFano factor (F = ((~m)~)l(m)) or the Q parameter {Q= [(m'") - (m)2]l(m)}.3~4 For a sub-Poissonian distributionthe Q parameter is negative. We now extend thiscriterion to higher-order factorial moments. The lth (1 isa positive integer) order factorial moment <strong>of</strong> the photocountdistribution is defined by(m(')) = m(m - 1) ...( m - 1 + l)p(m, T), (2)m=lwhere p(m, T) is the probability <strong>of</strong> detecting m photoi~sin the counting interval [0-TI. Here we have suppressedthe time argument in the factorial moments. Toextend the criteria for sub-Poissonian statistics to higherordermoments we introduce a parameter S1 :It is easily proved that, for a Poisson distribution, S1= 0 for all 1. Parameter Sl for 1 3 2 provides a measure<strong>of</strong> the deviation <strong>of</strong> the I th factorial moment from that fora Poisson distribution with the same mean. Sl > 0 definesa super-Poissonian distribution, and S1 < 0 definesa sub-Poissonian distribution. Note that parameter Sz isnot equal to the Q parameter but is related to it byS, = Ql(m). As the values <strong>of</strong> higher-order factorial momentscan be large, it is convenient to use normalized factorialmoments rather than the analogs <strong>of</strong> Q to extend theconcept <strong>of</strong> sub- and super-Poissonian statistics to higherordermoments. Another advantage <strong>of</strong> using the SL parametersis that they are independent <strong>of</strong> the efficiency <strong>of</strong>detection.We now show that negative values oF parameter S1 forI > 2 (higher-order sub-Poissonian statistics) indicate thenonclassical nature <strong>of</strong> light. To establish this we notethat for a classical field the factorial moments must satisfythe inequalitys3'0740-3224/2002/061471-05$15.00 O 2002 Optical Society <strong>of</strong> America


PHYSICAL REVIEW A, VOLUME 64, 043806Nonclassical effects in photon statistics <strong>of</strong> atomic optical bistabilityDaniel Erenso, Ron Adan~s, Hua Deng, Rceta <strong>Vyas</strong>, and Surendra Sing11Department <strong>of</strong> <strong>Physics</strong>, <strong>University</strong> oJArkunsuv, F~ryetfeville, <strong>Arkansas</strong> 72701(Received 30 Miiy 2001; published 13 September 2001)Homodync statistics <strong>of</strong> light generated by an atomic system cxhibiting optical bistability arc analyzed. Usingthe dynamical equations <strong>of</strong> lnotion for a single atom in a coherently driven cavity in the good cavity limit, weshow that thc homodync ficld can bc dcscribcd in terms <strong>of</strong> two iltdepcndcnt rcal Gaussian stochastic processcsand a coherent component. By making a Karhuneil-Loeve expansion <strong>of</strong> thc field variables we derive thegcl~crating function for the photoclcctron statistics. From this gc~~crating function photoclcctron-counting distribution,factorial moments, and waiting-time distribution are obtained analytically. These quantities are directlymeasurablc in photon-counting expcrimcnts. We show that thc homodync ficld cxhibits many interestingnonclassical features including nonclassical etrects in higher-order factorial moments.DOI: 10.1 103iPhysRevA.64.043806PACS number(s): 42.50.Ar, 42.50.Dv, 42.50.CtTnteraction <strong>of</strong> a single two-level atom with a quantumficld insidc a coherently drivcn cavity in the good cavitylimit, is known to show optical bistability [1,2]. We will referto this system as single atom optical bistability (SAOB).Similarly, N two-level atoms placed inside a coherentlydriven cavity also exhibit optical bistability that we shallrefer to as multiatom optical bistability (MAOB) [3,4]. Thesesystems are also known to show antibunching, although thesize <strong>of</strong> antibunching is small. In order to enhance antibunchingand other nonclassical effects, several schcmcs based oninterference [5], passive filter cavities [6], or homodyne detection[7] have been proposed.Homodyning a field with a coherent local oscillator (LO)provides one way <strong>of</strong> cnhancing nonclassical effects. The homodynefield can exhibit strong nonclassical featurcs, whichare not shown by the original ficld. The homodyile statisticsare sensitive to the phase difference betwecn the signal andthe LO. An example <strong>of</strong> this behavior is provided by the lightfrom the degenerate parametric oscillator, which is highlybunched and super-Poissonian. When this field is homodynedwith a LO, the homodync field shows a variety <strong>of</strong>nonclassical cffccts such as antibunching, sub-Poissonianstatistics, and violation <strong>of</strong> other classical inequalities [8-101.In this paper we consider homodyning <strong>of</strong> the light from asystem that exhibits SAOB with the light beam from a LO ata losslcss beam splitter as shown in Fig. 1. A detector <strong>of</strong>efficicncy 77 placed at one <strong>of</strong> the output ports <strong>of</strong> the beamsplitter detects the homodyne field and generates photoelectricpulses, which are measured by suitable electronics. Westudy photoelectron statistics measured by the detector. InSec. I1 we start from the equations <strong>of</strong> motion derived byWang and <strong>Vyas</strong> for a single two-level atom in the good cavitylimit [2] and show that the ficld from the SAOB can beexpressed in terms <strong>of</strong> two Gaussian random variables. Wethen derive the equations that govern the dynamics <strong>of</strong> thehomodyne field. Using these equations and applying theKarhunen-Lokve expansion for the field variables, we calculatethe moment generating function for the photocount distribution.We also show that a system exhibiting MAOB canalso be described by similar expressions with an appropriatechange in parameters. In Scc. TI1 we present an analytic expressionfor thc moment gencrating function. Photon statistics<strong>of</strong> the homodync field are then analyzed with the help <strong>of</strong>the moment generating function. The photocount distribution,its moments, and the waiting-time distribution for thehomodyne field are calculated. Finally, in Sec. VI, a summary<strong>of</strong> the main results <strong>of</strong> the paper is presented.11. DYNAMICS OF THE HOMODYNE FIELD ANDTHE GENERATING FUXCTIONIn this section we derive equations <strong>of</strong> motion describingthe dynamics <strong>of</strong> the homodync field when the signal is fromthe SAOB. We will see that similar equations are obtainedwhen the signal is from thc MAOB.Consider a single damped two-level atom with transitionfiequcncy w,, interacting with a single mode <strong>of</strong> a cavitywith resonance frequency w,. The cavity is driven by a coherentexternal field <strong>of</strong> frequency w, and amplitude E. In theelectric dipole and rotating-wave approximation, the Hamiltonianfor this system call be written asHereand dt are the annihilation and creation operators forthe cavity mode, a+, 6.- , and 6, are the Pauli spin matricesdescribing the two-level atom, g is the atom-field couplingconstant, and H,,,, describes atomic losses due to spontaneousdecay and field losses at the cavity mirrors.Two-level atom+FIG. I. System for homodyning the SAOB field with the LOfield. BS denotes the beam splitter and D denotes a detector.1050-2947/2001/64(4)/043806(7)1$20.00 64 043806-1 02001 The American Physical Society


VOLUME 86, NUMBER 13 PHYSICAL REVIEW LETTERS 26 MARCH 2001Entanglement, Interference, and Measurementin a Degenerate Parametric OscillatorHua Deng, Daniel Erenso, <strong>Reeta</strong> <strong>Vyas</strong>, and Surendra SinghDepartment <strong>of</strong> <strong>Physics</strong>, <strong>University</strong> <strong>of</strong> <strong>Arkansas</strong>, Fayetteville, <strong>Arkansas</strong> 72701(Received 25 August 2000)Quantum dynamical equations <strong>of</strong> motion for homodyne detection <strong>of</strong> the degenerate optical parametricoscillator are solved exactly. Nonclassical photon statistics are shown to be a consequence <strong>of</strong> interference<strong>of</strong> probability amplitudes, entanglement <strong>of</strong> photon pairs from such an oscillator, and the role <strong>of</strong>measurement in quantum evolution.DOI: 10.1 103/PhysRevLett.86.2770PACS numbers: 42.50.Dv, 42.50.A~ 42.65.K~Fluctuations <strong>of</strong> photon beams reflect the quantum dynamics<strong>of</strong> photoemissive sources. In quantum mechanics,probabilities for observed events are derived from an underlyingwave function that can interfere and collapse as itevolves. A consequence <strong>of</strong> this is that quantum mechanicscan lead to correlations between observed events whicha classical stochastic theory may not. Examples <strong>of</strong> thesenonclassical correlations include squeezing, antibunching,and violations <strong>of</strong> Bell's inequalities [1,2].The subthreshold degenerate parametric oscillator(DPO) has played a central role in the study <strong>of</strong> nonclassicalphoton correlations, particularly, squeezing [1,2]. TheDPO radiates a highly bunched light beam that exhibits alarge degree <strong>of</strong> squeezing. Interestingly, the squeezed andhighly bunched light from the DPO when combined with acoherent light field, as in homodyne detection, is expectedto display a rich variety <strong>of</strong> nonclassical photon correlationsincluding antibunching [3]. It is intriguing that a highlybunched entangled photon beam from the DPO whenmixed with a coherent field will exhibit correlations similarto those exhibited by a single-atom resonance fluorescencein free space or in cavity quantum electrodynamics (QED)[4,5]. Antibunching <strong>of</strong> light emitted by a single two-levelatomic system can be eventually traced to the atomic deadtime that a two-level atom cannot emit a second photonimmediately after the emission <strong>of</strong> a first photon. The situationis not so simple for homodyne detection <strong>of</strong> the lightfrom the DPO because there is no obvious mechanismfor a dead time. By solving the equations <strong>of</strong> motion forhomodyne detection exactly, we show that nonclassicalphoton correlations in homodyne detection <strong>of</strong> the DPO area consequence <strong>of</strong> the interference <strong>of</strong> probability amplitudes,entangled nature <strong>of</strong> photon pairs generated by theDPO, and measurement. These are the features that mostdistinguish quantum mechanics from classical mechanics.An outline <strong>of</strong> the experimental setup for homodyne detection<strong>of</strong> the DPO light is shown in Fig. 1. The DPO andlocal oscillator (LO) fields are combined by a beam splitterto produce the source field at the detector. The field fromthe DPO is governed by the interaction Hamiltonian for aphase matched DPO driven by a classical injected signal<strong>of</strong> amplitude E [6]:ihK&H = (&J2 - 6;) + Aloss . (1)LHere K is the mode-coupling constant and and adtarethe annihilation and creation operators, respectively, forthe DPO. H~,,, describes the loss suffered by the DPOfield. The combination KE can be chosen to be real by asuitable definition <strong>of</strong> phases.The equation <strong>of</strong> motion for the density matrix bd <strong>of</strong> theDPO field is thent ..t.. t+ ~(2adbdad - adadbd - bdadad)? (2)where 2 y is the cavity decay rate. The steady-state solutionto this equation in positive-P representation is given bywhere -a < x, y < are both real variables and In) is acoherent state <strong>of</strong> ad with adlx) = xlx). From this expressionfor the density matrix, the steady-state expectationvalue <strong>of</strong> an operator 0 can be calculated as (o),, =T~(o$,,). This leads to the following expectation valuesFrom DPOFrom LODetectorFIG. 1. Schematic experimental setup for the homodyne detection<strong>of</strong> the light from a degenerate parametric oscillator.2770 0031 -9007/01/ 86(13)/2770(4)$15.00 O 2001 The American Physical Society


PHYSICAL REVIEW A, VOLUME 62, 033803Higher-order nonclassical effects in a parametric oscillator<strong>Reeta</strong> <strong>Vyas</strong> and Surendra Singh<strong>Physics</strong> Deportment, Lr/ziversity <strong>of</strong> Arkcmsas, Fayetleville, <strong>Arkansas</strong> 72701(Received 10 January 2000; published 8 Aubwst 2000)Factorial ~noments for thc light from a degencratc paramctlic oscillator mixed with a cohcrcnt local oscillatorare calculated and shown to reveal novel nonclassical features <strong>of</strong> light. We have round novel regimeswhcrc the nonclassical charactcr <strong>of</strong> thc ficld is rcflccted not in thc violations <strong>of</strong> incqualitics bascd on thcsecond-order moments but in those based on higher ordcr moments. These violations can be observed mphotoclcctric mcasurclnents <strong>of</strong> light.PACS numbcr(s): 42.50.Dv, 42.50.Ar, 42.65.K~Nonclassical properties <strong>of</strong> light such as squeezing [I],sub-Poissonian statistics [2], and antibunching [3] have beendiscussed in terms <strong>of</strong> quadratic field or intensity correlations.With the developn~ent <strong>of</strong> techniques for making higher ordercorrelation measurements. the interest in nonclassical effectsnaturally extends to the higher order correlations. Of particularinterest are physical systems where nonclassical features<strong>of</strong> light are revealed in higher order correlations but not inthe lower order correlations. Higher order squeezing was introducedby Hong and Mandel [4]. Aganval and Tara introducednonclassical behavior in terms <strong>of</strong> normally ordercdmoments <strong>of</strong> a distribution [5]. Lee [6] considered a singlemode field and defined higher order nonclassical effects interms <strong>of</strong> factorial moments <strong>of</strong> the photon distribution by usingmajorization theory.Jn this paper we present a simple criterion for higher ordernonclassical behavior in tcrms <strong>of</strong> factorial moments <strong>of</strong> thephotoelectron counting distribution. We then show that thelight fi.om a subthreshold degenerate parametric oscillator(DPO), when superposed on the light from a coherent localoscillator (LO), satisfies the criteria for highcr order nonclassicalbehavior in terms <strong>of</strong> factorial moments. These violationscan be measured in photon counting experiments.Let us first recall that the probability <strong>of</strong> detecting m photoelectronsin the time interval [O - T] is [7]where, for simplicity, we have suppresscd the dependence <strong>of</strong>the moments on the counting interval T. The second factorialmoment (rn(l)) is related to the variancc ((dm)') <strong>of</strong> photonnumber distribution byFor a coherent field (Poissonian distribution) ((Am12)= (m) is independent <strong>of</strong> the value <strong>of</strong> T. Expressing p(m, 71in Eq. (3) in terms <strong>of</strong> 0 and using thc Glauber-Sudarshan Prepresentation for the density matrix <strong>of</strong> the field, the factorialmoments can be written aswhere U is a positive real number and for a classical state theprobability distribution P(U) must be a positive and nonsingularfunction (no more singular than a 6 function). Nowdefine a positive fbnction in tams <strong>of</strong> two positive real variablesU and Waswhere the colons : . . . : denote time ordering and normal ordering<strong>of</strong> the operator product between the colons. In wi-itingEq. (I), we have assumed stationary light fields. The operatorfi is given byOn expanding and simplifying this, we find thatTherefore for a classical probability density P(U,W)[= P(U)P(W)], which is positive and no more singular thana 6 function,where i(t) is the photon flux operator (number <strong>of</strong> photonsper second) and r] is the detection efficiency. The angularbrackets denote averaging with respect to the state <strong>of</strong> thefield.The /-th order factorial moment (/ is a positive integer)<strong>of</strong> the photoelectron counting distribution is defined byThus the factorial inoments <strong>of</strong> photon counting distributionfor a classical field must satisfy the inequalities1050-2947/2000/62(3)1033803(5)15 15.00 62 033803-1 Q2000 The American Physical Society


JOURNAL OF APPLIED PHYSICS VOLUME 84, NUMBER 7 1 OCTOBER 1998Thermal diffusivity measurement <strong>of</strong> solid materials by the pulsedphotothermal displacenient techniqueG. L. Bennis, R. <strong>Vyas</strong>, and R. GuptaDepartment <strong>of</strong> <strong>Physics</strong>, <strong>University</strong> <strong>of</strong> <strong>Arkansas</strong>, FayetteviNe, <strong>Arkansas</strong> 72701S. Ang and W. D. BrownDepartment <strong>of</strong> Electrical Engineering, <strong>University</strong> <strong>of</strong> <strong>Arkansas</strong>, FayefteviNe, <strong>Arkansas</strong> 72701(Received 27 February 1998; accepted for publication 2 July 1998)A simple, noncontact technique for the measurement <strong>of</strong> thermal diffusivity <strong>of</strong> solids isexperimentally demonstrated. The technique is based on the photothermal displacement effect.Excellent agreement between the quasistatic theory <strong>of</strong> photothermal displacement and theexperiment has been obtained. The technique has been demonstrated by measuring the thermaldiffusivities <strong>of</strong> GaAs and InGaAsIAlGaAs multiple quantum wells. O 1998 American Institute <strong>of</strong><strong>Physics</strong>. [SOO21-8979(98)04519-81I. INTRODUCTIONThermal diffusivity <strong>of</strong> a material is an important property,a knowledge <strong>of</strong> which is required for a variety <strong>of</strong> applications.Of the many methods available for making diffusivitymeasurements, noncontact methods are generallypreferable over the more traditional methods in which probesmust be inserted into the sample. One <strong>of</strong> the popular noncontactmethods is the photothermal technique.'!2 In this technique,local heating <strong>of</strong> the sample is created by a laser beam(pump beam), and the effects <strong>of</strong> the heating are observed,generally by another, weaker laser beam (probe beam). Theexperiments can be performed by using either a cw modulatedpump beam or a pulsed pump beam. The former areknown as the frequency-domain experiments, whereas thelatter are known as the time-domain experiments. The followingvariations <strong>of</strong> the photothermal technique have beenused to measure thermal diffusivity in solids: (i) Phototherma1radiometry. In this method, the heating is produced by apulsed laser and the time evolution <strong>of</strong> the heating is measuredby observing the infrared emission from the samplesurface. Thermal diffusivity is obtained from the timeevolution<strong>of</strong> the radiometric Chen and ande el is'have used a variation <strong>of</strong> this technique, called rate-windowspectrometry, to enhance the signal-to-noise ratio. (ii) Photothermalrefraction: The heating <strong>of</strong> the sample is observedby a probe laser passing through the sample, which is transparentto the probe beam. The heating produces refractiveindex gradients in the sample, which cause the probe beam todeflect. The thermal diffusivity is obtained by measurements<strong>of</strong> this deflection either with cw-modulated excitation6 orwith pulsed e~citation.~ (iii) Photothermally induced mirageeffect: In this technique, the measurement is made in a fluidmedium surrounding the sample. The heat from the solidsample is conducted to the fluid, which produces refractiveindex gradients in the fluid. Deflection <strong>of</strong> a probe beam passingthrough the fluid is measured. Generally, a cw-modulatedpump is used and deflection is measured as a function <strong>of</strong>pump-probe distance. Thermal diffusivity is obtained eitherby an analysis <strong>of</strong> the phase information contained in thesignal839 or by a multiparameter fit to the data.'' (iv) Probebeam reflection: The temperature change <strong>of</strong> the sample canalso be monitored by observing the change in the reflectivity<strong>of</strong> the sample with(v) Photothermal displacement:In 1983, Olmstead et al.I3 showed that the heating<strong>of</strong> the sample by the pump beam produces a thermoelasticdeformation <strong>of</strong> the sample which can be detected bydeflection <strong>of</strong> the probe beam reflected from the sample surface,interferometric methods, or an attenuated total reflectionscheme. Analysis <strong>of</strong> the signal, either in the frequencyor time domain, in principle, would yield the thermal diffusivity.However, to the best <strong>of</strong> our knowledge, this techniquehas not been fully exploited and is the subject <strong>of</strong> this article.(vi) Transient thermal grating: In this technique, two coherentbeams are made to interfere on the surface <strong>of</strong> the sample.This produces a thermal grating. In one <strong>of</strong> the schemes, thisgrating is detected by observing the temperature inducedchanges in reflectivity.14 The thermal grating also produces adeformation grating on the sample surface by thermoelasticdeformation, which can be observed by the diffraction <strong>of</strong> aprobe beam,15 or by the deflection <strong>of</strong> a well-focused probebeam from one undulation <strong>of</strong> the deformation grating.'6 Jaureguiand welsch17 have reviewed the theory <strong>of</strong> thermal gratingsin both the frequency and time domains. The above list<strong>of</strong> various photothermal techniques for thermal diffusivitymeasurements is by no means exhaustive, as the literature onthis subject is vast. A good recent review <strong>of</strong> the field is givenby Park et a/.In this article we demonstrate that the thermal diffusivitycan be measured very simply by the temporal evolution <strong>of</strong>the thermoelastic deformation following a short laser pulse.As mentioned in (v) above, to the best <strong>of</strong> our knowledge, thissimple technique has not been fully exploited previously.The feasibility <strong>of</strong> thermal diffusivity measurement by thismethod was reported by Karner et a/." and by Vintsents and~andornirskii.~' However, no analyses <strong>of</strong> the data were performedto obtain actual thermal diffusivity values. We findthat our data fit extremely well to the relatively simple theory<strong>of</strong> thermoelastic deformation given by ~i;' and we obtain0021 -8979/98/84(7)/3602/9/$15.00 3602 O 1998 American Institute <strong>of</strong> <strong>Physics</strong>Downloaded 26 Feb 2008 to 130.284.237.6, Redistribution subject to AIP license or capyrlght; see http:l1jap.aip.orgljaplcopyright.jy1


Photothermal lensingdetection: theory and experimentQifang He, <strong>Reeta</strong> <strong>Vyas</strong>, and Rajendra GuptaPhotothermal lensing signal shapes are experimentally investigated and compared with those predictedtheoretically in our earlier paper. The investigation included flowing and stationary media and pulsedand cw excitations. Good qualitative agreement between theory and experiment is found. Since thelensing signal is almost always accompanied by a deflection signal, the influence <strong>of</strong> the deflection signalon the detection <strong>of</strong> lensing signal is investigated. For a perfectly aligned detection geometry theinfluence <strong>of</strong> the deflection signal on the lensing signal is negligible, but in the presence <strong>of</strong> misalignmentsa significant amount <strong>of</strong> deflection signal could be superimposed on the lensing signal. The effect <strong>of</strong>lensing on the deflection signal is also been considered. The effect <strong>of</strong> the finite size <strong>of</strong> the probe beam onthe lensing signals is also investigated. O 1997 Optical Society <strong>of</strong> AmericaKey words: Photothermal lensing, photothermal deflection, thermal blooming.1. IntroductionThere is currently an intense interest in the technique<strong>of</strong> photothermal spectroscopy1 because thetechnique has found many applications in diversefields.2 Sometime ago we gave the complete theory<strong>of</strong> photothermal lensing spectroscopy (PTLS) in aflowing fluid medium.3 A general treatment wasgiven that consisted <strong>of</strong> pulsed (<strong>of</strong> arbitrary pulselength) or cw (modulated or unmodulated) excitation,flowing or stationary media, and transverse or collineargeometry. To our knowledge no systematicexperimental investigation <strong>of</strong> the signal shapes presentedin our previous paper has yet been published.In the present paper we present the experimentalinvestigation <strong>of</strong> the lensing signals and their comparisonwith the theoretical results presented in Ref. 3,which we will hereafter refer to as paper I. Duringthis investigation we realized that the photothermaldeflection spectroscopy (PTDS) signal, which almostalways accompanies the PTLS signal, has the potentialfor complicating the shapes <strong>of</strong> the PTLS signals.Therefore in this paper we also present both the the-When this research was performed all the authors were with theDepartment <strong>of</strong> <strong>Physics</strong>, <strong>University</strong> <strong>of</strong> <strong>Arkansas</strong>, Fayetteville, <strong>Arkansas</strong>72701-1201. Q. He is now with the <strong>Arkansas</strong> State <strong>University</strong>,Beebe Branch, Beebe, <strong>Arkansas</strong> 72012.Received 24 October 1996; revised manuscript received 26 February1997.0003-6935/97/277046-13$10.00/0O 1997 Optical Society <strong>of</strong> Americaoretical and the experimental investigation <strong>of</strong> the effect<strong>of</strong> PTDS on the shapes <strong>of</strong> the PTLS signals. Forcompleteness, in Appendix A we also discuss the effect<strong>of</strong> PTLS on the shapes <strong>of</strong> PTDS signals. Theeffect <strong>of</strong> the finite size <strong>of</strong> the probe beam on the detection<strong>of</strong> the lensing signal has also been investigated.The basic idea underlying dual-beam PTLS isshown in Fig. 1. A laser beam (pump beam) propagatesthrough a medium, and it is tuned to one <strong>of</strong> theabsorption frequencies <strong>of</strong> the medium. The mediumabsorbs some <strong>of</strong> the optical energy from the laserbeam. If the collision rate in the medium is sufficientlyhigh compared with the radiative decay rates,most <strong>of</strong> the energy appears in the translationalrotationalmodes <strong>of</strong> the medium within a short period<strong>of</strong> time. In other words, the laser-irradiated regionbecomes slightly heated. The refractive index <strong>of</strong> themedium is thus modified. The refractive-indexchange can be monitored in several different ways.4In this paper we are concerned with a technique thatrelies on the lensing effect <strong>of</strong> the medium to monitorthe refractive-index change. A weak probe beampasses through the pump-irradiated region, as shownin Fig. 1. Owing to the curvature <strong>of</strong> the refractiveindex, the probe beam diverges, which can be detectedas a change in the intensity <strong>of</strong> the probe beampassing through a pinhole. In other words, underthe influence <strong>of</strong> the pump beam the medium acts as adiverging lens. In certain circumstances the mediumacts as a converging lens also. If a pulsedpump laser is used, a transient lens is formed; the7046 APPLIED OPTICS / Vol. 36, No. 27 / 20 September 1997


Theory <strong>of</strong> photothermalspectroscopy in an optically dense fluidQifang He, <strong>Reeta</strong> <strong>Vyas</strong>, and R. GuptaThe theory <strong>of</strong> photothermal spectroscopy in an optically dense fluid is presented. The general case isconsidered in which the fluid may be flowing or stationary, and the excitation could be cw (modulated orunmodulated) or pulsed (arbitrary pulse length). All three detection schemes, deflection, phase shift,and lensing, are considered. This is the most complete theory <strong>of</strong> photothermal spectroscopy in fluids todate. O 1997 Optical Society <strong>of</strong> AmericaKey words: Photothermal deflection, photothermal phase shift, photothermal lensing, photothermalinterference, thermal blooming.1. IntroductionThere is currently a great interest in the technique <strong>of</strong>photothermal spectroscopy. This is because thetechnique has numerous applications in many diversefields.1 For example, the technique can beused for the measurement <strong>of</strong> species concentration,thermal diffusivity, laser spatial pr<strong>of</strong>ile, relaxationtime, local temperature, and flow velocity. Thesemeasurements are nonintrusive and can be performedwith high degrees <strong>of</strong> spatial and temporalresolution and are particularly useful in hostile environmentssuch as combustion.2The principle <strong>of</strong> the technique is simple. A laserbeam (pump beam) passes through the medium <strong>of</strong>interest. The laser is tuned to an absorption line <strong>of</strong>the medium and the optical energy is absorbed by themedium. In a fluid medium, if the collisionalquenching rate is much higher than the radiativerate, most <strong>of</strong> this energy appears in the rotationaltranslational(thermal) modes <strong>of</strong> the molecules. Theheating <strong>of</strong> the medium modifies its refractive index.The change in the refractive index <strong>of</strong> the medium isdetected by a second (generally a He-Ne) laser beam(probe beam).Three distinct methods <strong>of</strong> monitoring the change inthe refractive index are in use. 1. Photothermalphase-shift spectroscopy (PTPS). In this techniqueThe authors are with the Department <strong>of</strong> <strong>Physics</strong>, <strong>University</strong> <strong>of</strong><strong>Arkansas</strong>, Fayetteville, <strong>Arkansas</strong> 72701.Received 26 March 1996; revised manuscript received 29 July1996.0003-6935/97/091841-19$10.00/0O 1997 Optical Society <strong>of</strong> Americaone can measure the refractive-index change directlyby placing the sample inside a Fabry-Perot cavity orin one arm <strong>of</strong> an interferometer. The refractiveindexchange produces a change in the optical pathlength that is detected as a fringe shift.3j4 2. Photothermaldeflection spectroscopy (PTDS). Therefractive-index change produced by the absorption <strong>of</strong>the pump beam is, in general, nonuniform. The nonuniformrefractive index causes the probe beam todeflect; this is detected by a position-sensitive opticaldetector.5 The PTDS signal thus is proportional tothe gradient <strong>of</strong> the refractive index anlax. 3. Photothermallensing spectroscopy (PTLS). The nonuniformrefractive index also causes a focusing ordefocusing <strong>of</strong> the probe beam, which is detected as anintensity change <strong>of</strong> the probe beam as it passesthrough a pinhole.6 The PTLS signal is proportionalto the curvature <strong>of</strong> the refractive index a2n/dx2.In a series <strong>of</strong> papers Gupta, <strong>Vyas</strong>, and collaboratorshave provided a coherent theoretical treatment <strong>of</strong> thethree techniques in fluid media.7-10 Both pulsedand cw optical excitation were considered. For thecw case the theory allowed for both modulated andunmodulated excitation. For the pulsed case bothshort and long excitation pulses compared with thediffusion and forced convection times were considered.Both the stationary and flowing media weretreated. In other words, that theoretical developmentis valid for general conditions and has the furtheradvantage that all three techniques (PTPS,PTDS, and PTLS) and the various conditions (cw orpulsed excitation, stationary or flowing medium, etc.)are treated in a coherent fashion.Here we generalize the theory further to includeoptically thick media. Therefore the theory pre-20 March 1997 / Vol. 36, No. 9 / APPLIED OPTICS 1841


PHYSICAL REVIEW A VOLUME 55, NUMBER 1 JANUARY 1997Cavity-modified Maxwell-Bloch equations for the vacuum Rabi splittingChangxin Wang and <strong>Reeta</strong> <strong>Vyas</strong><strong>Physics</strong> Department, <strong>University</strong> <strong>of</strong> <strong>Arkansas</strong>, Fayetteville, <strong>Arkansas</strong> 72701(Received 5 March 1996; revised manuscript received 29 August 1996)We derive a set <strong>of</strong> cavity-modified Maxwell-Bloch equations for a two-level atom in a cavity driven by acoherent field from the single-atom Fokker-Planck equation in the bad-cavity limit. These equations have thesame form as the Maxwell-Bloch equations for a two-level atom in free space interacting with a coherent field.The presence <strong>of</strong> a cavity is reflected in three cavity-modified parameters: decay rate ye//, Rabi frequencya, and detuning A,//. We show that the cavity-modified Maxwell-Bloch equations provide an easy way tostudy cavity-modified spontaneous emission, cavity-induced radiative energy level shifts, and vacuum Rabisplitting. [S1050-2947(96)075 12-91PACS number(s): 42.50.Ct, 32.80.-t, 42.50.L~The radiative properties <strong>of</strong> atoms in a cavity have beeninvestigated by many authors in the framework <strong>of</strong> cavityquantum electrodynamics [I]. Fields inside a cavity are subjectto boundary conditions and when the dimensions <strong>of</strong> thecavity are comparable to the wavelength <strong>of</strong> the radiation, thespectrum and spatial distribution <strong>of</strong> the electromagnetic fieldmodes are modified. As a consequence. atoms inside such acavity exhibit many different radiative properties from thosein free space, such as cavity-modified spontaneous emission[2], cavity-induced radiative frequency shifts [3], andvacuum Rabi splitting [4]. In the laboratory, several experimentshave verified the enhancement or inhibition <strong>of</strong> thespontaneous emission rate <strong>of</strong> an atom in a cavity [S]. Thephenomena <strong>of</strong> vacuum Rabi splitting in the strong couplingregime have been observed experimentally, for cavities witha large number <strong>of</strong> atoms, as well as for a small number <strong>of</strong>atoms [6,7]. In particular, the case <strong>of</strong> a single atom on theaverage inside an optical cavity has been studied experimentally[7].To describe a single atom inside a cavity we derived ageneralized second-order Fokker-Planck equation for asingle two-level atom interacting with a cavity mode, withoutusing the system size expansion and any truncation [8].In this paper, we show that this exact single-atom Fokker-Planck equation in the bad-cavity limit leads to a set <strong>of</strong>cavity-modified Maxwell-Bloch equations. We use theseequations to study the cavity-modified spontaneous emission,radiative level shifts, and vacuum Rabi splitting.Our quantum dissipative system consists <strong>of</strong> a singledamped two-level atom with transition frequency w, interactingwith a damped cavity mode <strong>of</strong> resonance frequencyw,. The cavity is driven by a coherent external field <strong>of</strong> frequencyw0 and amplitude E . In a frame rotating at the frequencyw0 <strong>of</strong> the external field, the behavior <strong>of</strong> the combinedatom-field system is governed by the master equation for thedensity operator p(t) [8],Here 6 and it are the field annihilation and creation operators;&+, &- , and &, are the Pauli spin operators describingthe two-level atom; y is the atomic spontaneous emissionrate; and 2 K is the rate at which the cavity is losing photons.Our model incorporates both the atomic detuningA,(= w,- wo) and the cavity detuning A,(= w,- wo).The master equation (1) can be transformed into a secondorderFokker-Planck equation without using system size expansionor any truncation in the bad-cavity limit [8]. Byeliminating the field variables adiabatically, we obtained asecond-order Fokker-Planck equation containing only theatomic variables. The drift terms <strong>of</strong> the Fokker-Planck equationlead to the differential equations for the mean values <strong>of</strong>the atomic operators [see Eqs. (28) in Ref. [8])], which canbe expressed in the following form:where cavity-modified parameters aredb-= -iAc[6t6,p]-iAll[&, ,p]+g[6t&--6&+ ,p]dtHere c = ~ ~ / is K the ~ single-atom version <strong>of</strong> the cooperat-ivity parameter, Y = 2 fig el^ is the dimensionless drivingYfield amplitude, 6, = 2 A, 1 y is the atomic detuning param-+~[i~-a^,bl+ -[2&-6&+-&+&-p-p&+&-]2eter, and aC= A, IK is the cavity detuning parameter. Equations(2) have the same form as the well-known Maxwell-+ ~[26p6~-ii~iip-p6~;]. (1) Bloch equations for a single two-level atom interacting with1050-2947/97/55(1)/823(4)/$10.00 - 55 823 O 1997 The American Physical Society


Modern <strong>Physics</strong> Letters B, Vol. 10, No. 14 (1996) 661-669@World Scientific Publishing CompanyPARTICLE TRAPPING BY OSCILLATING FIELDS:CONNECTING SQUEEZING WITH COOLINGB. BASEIA'Instituto de Fisica, Universidade de Sio Paulo, Caiza Postal 66318,05389-970, Slo Paulo (SP), BmzilREETA VYAS<strong>Physics</strong> Department, <strong>University</strong> <strong>of</strong> <strong>Arkansas</strong>, Fayetteville, AR 72701, USAV. S. BAGNATOInstifufo de Fisica e ~u:mica de Sa'o Carlos, Uniuersidode de Sio Paulo,CP 369, 13560-970, Sio Carlo$ (Sf), BmtilReceived 6 November 1995Revised 15 June 1996The first observation <strong>of</strong> the squeezing effect outside the optical domain was reported vervrecently for trapped atoms [D. M. Meekh<strong>of</strong> et al., Phys. Rev. Lett. 76, 1796 (1996)l.Putsuing this line and a sequel <strong>of</strong> previous works <strong>of</strong> ours we employ a model by Glauberand a method <strong>of</strong> invariants by Lewis and Riesenfeld to establish a connection betweenthe squeezing and the cooling effects in the system. Rom this connection the occumnce<strong>of</strong> squeezing could be detected through the measurement <strong>of</strong> the dispersion in the variableP (or i) and <strong>of</strong> the temperature.Laser cooling and trapping techniques have played an important role in newadvances <strong>of</strong> atomic physics. hiany techniques <strong>of</strong> cooling atoms or ions down totheir zero point vibrational energy have emerged from the successful application <strong>of</strong>radiation pressure and the Paul and Penning traps.'?* These techniques are appliedto reduce relative motion <strong>of</strong> the trapped particle and such systems, with extremelylow center <strong>of</strong> mass energy, must be treated quantum me&anically. Squeezed states<strong>of</strong> electromagnetic field are weII known in Quantum O~tics.~ For squeezed statesfluctuations in one <strong>of</strong> the quadrature is less than the mcuurn fluctuations. Recentlythe idea <strong>of</strong> squeezed states has been extended to mechanical motion <strong>of</strong>a trapped particle and schemes for preparation <strong>of</strong> squeezed state <strong>of</strong> motion hasbeen pr~~osed.~*~ Heizen and Wineland4 proposed generation <strong>of</strong> squeezed state <strong>of</strong>'Email: hasilioOif.usp.br; On leave from Departamento de Fisica, Universidade Federal daPara'ba (PB), Brazil.PACS Nus.: 32.80.Pj; 32.80.P~; 42.50.D~661


Physica A 232 (1996) 273-303PHYSIGA BMultiphoton interaction <strong>of</strong> a phased atom with asingle mode fieldA.N. Chabaay*, B. Baseiaa, Changxin wangb, <strong>Reeta</strong> vyasba Unicersidade Federal da Paraiba. Departamento de Fisica, CCEN, Caixa Postal 5008, CEP58.051-970. Joao Pessoa, Paraiba, BrazilPhysfcs Department, Unicersiry <strong>of</strong> <strong>Arkansas</strong>, Fayettecille, AAR 72701, USA-AbstractReceived 29 November 1995Multiphoton interaction <strong>of</strong> a phased atom prepared in a coherent superposition <strong>of</strong> its ground andexcited states with a general single mode optical field is studied. Calculations are done by usinga state vector <strong>of</strong> the atom-field composite system. Dynamic behavior <strong>of</strong> atomic inversion, fieldstatistics, and squeezing are studied for the multiphoton interaction <strong>of</strong> a phased atom includingthe effects <strong>of</strong> detuning. Results are discussed for various special cases <strong>of</strong> initial superposed atomicstate and field states such as number state, coherent state and squeezed states.1. IntroductionIn this paper, we discuss the dynamics <strong>of</strong> a two-level atom initially in a state whichis a coherent superposition <strong>of</strong> its ground and excited states (called phased atom), witha field initialIy in a pure state in a lossless cavity. We have used a generalization<strong>of</strong> the Jaynes-Cummings model [I] (JCM), proposed by Buck and S~k~mar [2], thatinvolves a multiphoton interaction between the atom and the field. The Hamiltonianfor this system in the dipole and the rotating wave approximation isWe shall refer to this as p-photon JCM. The standard model, which corresponds toP = I, will be called simply the JCM. Here w is the mode frequency <strong>of</strong> the field andA is the detuning, A = wo - po, coo being the transition frequency <strong>of</strong> the atom, and9 is the coupling strength for the ndiatiowatom interaction. The Pauli matrices c+,O-, and 03 represent the mising, lowering, and the inversion operators for the atom' Corresponding author."--.. ..-.,. . . #-.-- ,a . tnnr clr-.n.nr Crimr- R v AII riht~ resenled


PHYSICAL REVIEW A VOLUME 54, NUMBER 5 NOVEMBER 1996Fokker-Planck equation in the good-cavity limit and single-atom optical bistabilityChangxin Wang and <strong>Reeta</strong> <strong>Vyas</strong><strong>Physics</strong> Department, <strong>University</strong> <strong>of</strong> <strong>Arkansas</strong>, Fayetteville. <strong>Arkansas</strong> 72701(Received 23 January 1996; revised manuscript received 11 July 1996)A generalized Fokker-Planck equation is used to study the problem <strong>of</strong> single-atom optical bistability in thegood-cavity limit. The effects <strong>of</strong> quantum fluctuations are investigated by linearizing fluctuations around thesteady-state value. It is shown that in the good-cavity limit, quantum fluctuations are relatively small, andhence there exist both absorptive and dispersive single-atom bistability. By comparing our single-atom Fokker-Planck equation with the N-atom Fokker-Planck equation, we find that in the good-cavity limit, these twodifferent Fokker-Planck equations reduce to identical equations. We therefore come to an important conclusionthat the results obtained from the many-atom Fokker-Planck equation in the good-cavity limit are also valid fora single atom. [S1050-2947(96)00411-81PACS number(s): 42.50.Lc, 42.50.Ct, 42.50.Dv, 42.50.ArI. INTRODUCTIONThe simplest quantum mechanical model for radiationmatterinteraction is the Jaynes-Cummings model [I]. It consists<strong>of</strong> a single two-level atom interacting with a singlequantized electromagnetic field mode. The Jaynes-Cummings model is exactly solvable under a variety <strong>of</strong> conditions.These analytical solutions provide important insightsinto the quantum characteristics <strong>of</strong> the electromagnetic field.As a result, this model has become a cornerstone in the modemquantum optics [2]. On the other hand, the simplicity <strong>of</strong>the Jaynes-Cummings model brings some limitations. Onelimitation results from the neglect <strong>of</strong> dissipation. At the opticalfrequencies dissipation is present both as spontaneousemission and as cavity loss through finite reflectivity mirrors.Consequently, an extension <strong>of</strong> the Jaynes-Cummings modelthat includes dissipation and an external driving field hasrecently drawn considerable interest. This driven Jaynes-Cummings model with dissipation exhibits many novel effects,including photon antibunching [3], cavity-enhancedspontaneous emission [4], and nonclassical steady-stateatomic inversion [5].Of particular interest is whether or not optical bistabilityexists for this single-atom system. McCall and Gibbs [6]proved that the semiclassical criteria for absorptive bistabilitycould be met for a single atom in a cavity. However, theysuggested that quantum fluctuations would strongly affectbistability and no reasonable hysteresis could be preservedfor the single-atom system.In the bad-cavity limit, it has been shown that large quantumfluctuations do indeed destroy bistability predicted bythe semiclassical theory. Hence there is no single-atom bistabilityin the bad-cavity limit [3,7]. In the good-cavity limit,Sarkar and Satchel1 [8] have used the master equation approachto study optical bistability with small numbers <strong>of</strong>atoms. However, they failed to find any evidence for singleatombistability because they chose a very small saturationphoton number n, in their numerical calculations. Using anumerical approach, Savage and Carmichael [9] have shownthat single-atom optical absorptive bistability does exist for aparameter regime at the interface between the quantum limitand the classical limit.In Ref. [7] we have derived a generalized Fokker-Planckequation for a single-atom in an optical cavity and studied itsbehavior in the bad-cavity limit. In this paper we study thegeneralized Fokker-Planck equation in the good-cavity limitby adiabatically eliminating atomic variables. We derive ananalytical solution to the problem <strong>of</strong> single-atom optical bistability.We use linearized theory to investigate the effects<strong>of</strong> quantum fluctuations on single-atom absorptive and dispersivebistability. We show that for a given cooperativityparameter as the saturation photon number n, increasesquantum fluctuations decrease. In this limit the condition forthe good-cavity is better satisfied. In the good-cavity limit,therefore, quantum fluctuations are too small to destroysingle-atom bistability. We compare our single-atom Fokker-Planck equation with an N-atom Fokker-Planck equation(FPE). The N-atom FPE was thought to be valid only for alarge number <strong>of</strong> atoms (N* 1) [lo-141. In the good-cavitylimit, we find that our single-atom Fokker-Planck equation isidentical to the N-atom Fokker-Planck equation if onechooses N= 1. This leads to an important conclusion that inthe good-cavity limit, all the results obtained from theFokker-Planck equation for many-atoms are also valid forthe single-atom system.11. SINGLE-ATOM FOKKER-PLANCK EQUATIONIN THE POSITIVE-P REPRESENTATIONThe extension <strong>of</strong> the Jaynes-Cummings model, as describedabove, consists <strong>of</strong> a single damped two-level atomwith transition frequency o, interacting with a singledamped cavity mode with resonance frequency o, . The cavityis driven by a coherent external field <strong>of</strong> frequency oo andamplitude E. In a frame rotating at the frequency oo <strong>of</strong> theexternal field the behavior <strong>of</strong> the combined atom-field systemis governed by the master equation for the density operator&t),4453 O 1996 The American Physical Society


PHYSICAL REVIEW A VOLUME 54, NUMBER 3 SEPTEMBER 1996Homodyne detection for the enhancement <strong>of</strong> antibunching<strong>Reeta</strong> <strong>Vyas</strong>, Changxin Wang, and Surendra Singh<strong>Physics</strong> Department, <strong>University</strong> <strong>of</strong> <strong>Arkansas</strong>, Fayetteville, AR 72701(Received 15 February 1996; revised manuscript received 22 April 1996)We propose a scheme based on homodyne detection for enhancing antibunching in second-harmonic generationand multiatom optical bistability. We show that depending on the reflectivity <strong>of</strong> the beam splitter,relative field strengths, and relative phase it is possible to achieve perfect antibunching in the superposed field.We also discuss other nonclassical effects exhibited by the superposed field and present curves to illustrate thebehavior. [S 1050-2947(96)09008-71PACS number(s): 42.50.Dv, 42.50.Ar, 42.65.K~I. INTRODUCTIONSqueezing [I], antibunching, and sub-Possonian statistics[2,3] are nonclassical features <strong>of</strong> the electromagnetic field.These nonclassical features have been <strong>of</strong> considerable interestas they provide testing grounds for the prediction <strong>of</strong>quantum electrodynamics. Squeezing is related to the wavelikecharacter <strong>of</strong> the electromagnetic field. It is measured ininterference experiments. Antibunching and sub-Poissonianstatistics, however, reflect the particlelike behavior <strong>of</strong> thefield and are measured in photon counting experiments. Asdiscussed in Ref. [4] squeezing, antibunching, and sub-Poissonian statistics are, in general, distinct nonclassical effectsin the sense that an electromagnetic field may exhibitone but not the other.The antibunching effect has been predicted in intracavitysecond-harmonic generation (ISHG) [5,6] and multiatom opticalbistability (MAOB) [7,8]. However, the predicted size<strong>of</strong> antibunching is small and would be difficult to detectexperimentally, as it occurs against a large coherent background.The predicted antibunching in these systems is inverselyproportional to the saturation photon number n,,which is <strong>of</strong> the order <strong>of</strong> lo6- 10' for the ISHG, andlo3 - lo4 for the MAOB. Several schemes based on interference[9] or passive filter cavities [lo-121 have been proposedto enhance the antibunching effect.We propose a scheme based on homodyne detection[13-151 for enhancing antibunching in these systems. Homodynedetection experiments have been used for measuringphase-sensitive properties <strong>of</strong> squeezed light [I]. It has beenshown that the light from a degenerate parametric oscillator,which is highly bunched and super-Poissonian [16,17], canexhibit many nonclassical effects using a similar detectionscheme [13]. In the homodyne detection experiment we considerthe interference <strong>of</strong> the signal beam from the ISHG orthe MAOB with a coherent local oscillator (LO) at a losslessbeam splitter as shown in Fig. 1. A detector <strong>of</strong> efficiency 7 isplaced at one <strong>of</strong> the output ports <strong>of</strong> the beam splitter. Thestatistics measured at the detector is sensitive to the relativephase between the signal and the LO. Thus particlelike properties(photon statistics) are intimately connected to wavelike(phase) property <strong>of</strong> the field. Because <strong>of</strong> this phase dependence,the homodyne field can exhibit enhanced antibunchingand violation <strong>of</strong> various classical inequalities. Since inthis scheme one can readily adjust various parameters suchas the strength <strong>of</strong> the local oscillator, transmittance, and relativephase, this scheme may provide a better way <strong>of</strong> enhancingantibunching.In Sec. I1 we briefly describe the homodyne detectionscheme. In Sec. 111 we apply this technique to the ISHG. InSec. IV we discuss the enhancement <strong>of</strong> antibunching for theMAOB. Finally, a summary and main conclusions <strong>of</strong> thepaper are presented in Sec. V.11. HOMODYNE DETECTIONFigure 1 shows a schematic diagram for the homodynedetection experiment. For the ISHG, a nonlinear crystal isplaced inside the cavity, whereas for the MAOB, N two-levelatoms are placed inside the cavity. The light from the ISHGor the MAOB is superimposed with the light from a LO at alossless beam splitter. The annihilation operators b, and b2at the output ports are related to those at the input ports by[13,141withNLC or N atomsIIIIFIG. 1. System for homodyning the ISHG or MAOB field withthe LO field. For the ISHG a nonlinear crystal (NLC) is placedinside the cavity and for the MAOB N two-level atoms are placedinside the cavity. BS denotes a beam splitter and D denotes a detector.1050-2947/96/54(3)/2391(6)/$10.00 3 2391 O 1996 The American Physical Society


VOLUME 74, NUMBER 12 PHYSICAL REVIEW LETTERS 20 MARCH 1995Exact Quantum Distribution for Parametric Oscillators<strong>Reeta</strong> <strong>Vyas</strong> and Surendra SinghDepartment <strong>of</strong> <strong>Physics</strong>, <strong>University</strong> <strong>of</strong> <strong>Arkansas</strong>. Fayetteville, <strong>Arkansas</strong> 72701(Received 15 August 1994)An exact quantum distribution for the nondegenerate parametric oscillators is presented and used todiscuss their coherence properties. It is found that while each mode individually approaches a classicalstate, many quantum features exhibited by their combination survive even in the semiclassical limit.PACS numbers: 42.50.Dv, 42.50.Ar. 42.65.K~Optical parametric oscillators (OPOs) are quantummechanical devices with a definite threshold for selfsustainedoscillations [ 1 -41. They have played a centralrole in squeezing and twin-beam noise reduction experiments[5,6]. Theoretical understanding <strong>of</strong> these properties<strong>of</strong> the OPOs has been based mostly on linearized treatmentsabove and below the threshold <strong>of</strong> oscillation [7-91.In the region <strong>of</strong> threshold where linearization fails, thecomplex-P distribution has been used [lo]. The complex-P distribution, unfortunately, does not have the character<strong>of</strong> a probability density and is <strong>of</strong> limited use for gaininginsights into the coherence properties <strong>of</strong> the OPOs.Another distribution, closely related to the complex P,is the positive-P distribution which is a true probabilitydensity [ll]. In this paper we present an analyticsolution for the positive-P function for the optical parametricoscillators. With the help <strong>of</strong> analytic solutionstremendous insights into the coherence properties <strong>of</strong> otheroscillators have been gained [12- 141. Our analytic treatmentis based on the observation that the quantum dynamics<strong>of</strong> OPOs is naturally confined to a bounded regionin an eight-dimensional phase space. The solution presentedhere provides us with an elegant picture <strong>of</strong> howthe coherence properties <strong>of</strong> the OPOs are transformed inthe threshold region. It also allows us to discuss quantumfeatures that survive even as the field amplitudes grow upto macroscopic values above threshold.We model the parametric oscillator by two quantizedfield modes <strong>of</strong> frequencies w, and w2 interacting witha third mode <strong>of</strong> frequency w3 = w, + w2 inside anoptical cavity via a X(2) nonlinearity. Modes o, and0 2 experience linear losses characterized by the decayrates yl = y2 = y, and mode w3 suffers linear lossescharacterized by the decay y3. The cavity is excited by aclassical pump at frequency w3. In the interaction picturethe microscopic Hamiltonian takes the form(1)where aj and 6,tare the annihilation and creation operatorsfor the modes, K is the mode coupling constant, e isthe pump field amplitude, and ffl,,, describes mode losses.This nonlinear quantum mechanical problem can bemapped into a classical stochastic process by using thepositive-P representation [Ill. Eliminating the pumpmode adiabatically (y3 >> y) we obtain the following set<strong>of</strong> Ito stochastic differential equations:where vi are white noise Gaussian random processeswith zero mean and correlation functions given by{q,(t)v,(t')) = Sij8(t - t'). Here time is measured inunits <strong>of</strong> ye', no = 2-yy3/~' is parameter that sets thescale for the number <strong>of</strong> photons necessary to explore thenonlinearity <strong>of</strong> interaction, and u = 2y3e/~is a dimensionlessmeasure <strong>of</strong> the pump field amplitude scaled togive a = no as the threshold condition. In the absence<strong>of</strong> mode losses adiabatic elimination <strong>of</strong> the pump mode isnot justified [15]. Note that the adiabatic approximationdoes not amount to a neglect <strong>of</strong> the entanglement <strong>of</strong>pump and down-converted modes. Complex variablestai and ai, correspond to 2; and ai, respectively. In thepositive-P representation, ai and ai, are not complexconjugates <strong>of</strong> each other.Equations (2)-(5) describe trajectories in an eightdimensionalphase space. An examination <strong>of</strong> theseequations reveals that the eight-dimensional phase spaceis naturally divided into two subspaces. If we considerthe four-dimensional subspace a2 = (cYI)*, a2* = (a~*)*,and < a/2, Ia2*J2 < u/2, we notice that thetrajectories starting in this subspace initially remainconfined to this subspace. In other words, the conditiona2 = (a,)f, aZr = (alr)* is preserved for all times ifinitially we start out in this subspace. The initial state,2208 0031 -9007/95/74(12)/2208(4)$06.00 O 1995 The American Physical Society


PHYSICAL REVIEW A VOLUME 52, NUMBER 2 AUGUST 1995'Statistical properties <strong>of</strong> a charged oscillator in the presence <strong>of</strong> a time-dependentelectromagnetic fieldA. L. de Brito, A. N. Chaba, and B. BaseiaDepartamento de Fisica-CCEN, Universidade Federal da Paraiba, Caixa Postal 5008, 58.051-970 Joao Pessoa, Paraiba, Brazil..<strong>Reeta</strong> <strong>Vyas</strong>Department <strong>of</strong> <strong>Physics</strong>, <strong>University</strong> <strong>of</strong> <strong>Arkansas</strong>, Fayetteville, <strong>Arkansas</strong> 72701(Received 29 August 1994; revised manuscript received 13 January 1995)The statistical properties <strong>of</strong> a charged oscillator in the presence <strong>of</strong> a uniform magnetic field are investigatedfor the case <strong>of</strong> a time-dependent electromagnetic field. Quite general results are obtained whenthe time dependence <strong>of</strong> the external field is not specified. As an example, an oscillating external field isconsidered. It is found that, depending upon the parameters, this system can show many nonclassicalfeatures, such as sub-Poissonian statistics and squeezing.PACS number(s): 42.50.Dv, 84.30.Ng, 03.65. - wI. INTRODUCTION 11. CHARGED PARTICLEIN A TIME-DEPENDENT ELECTROMAGNETICIn both quantum mechanics and electromagneticFIELD AND TIME-DEPENDENTtheory one can find extensive literature on charged parti-HARMONIC OSCILLATORcles and charged oscillators in the presence <strong>of</strong> a constant[I] or time-dependent fields [2]. People mainly concen-Let us consider a particle, which may be an isotopic ostratetheir attention on calculating propagators, invaricillator,<strong>of</strong> charge e and mass M moving in gn axiallyants, eigenvalues and eigenvectors, coherent andsymmetric magnetic field defined by the vector potentialsqueezed states, etc. There is also considerable interest in(cf., e.g., Lewis and Riesenfeld [2])the production <strong>of</strong> quantum states <strong>of</strong> light, such assqueezed states [3], sub-Poissonian states [4], andphoton-number states [5]. Agarwal and Arun Kumar [6] and a scalar potentialstudied statistical properties <strong>of</strong> a one-dimensional oscillatorwith a time-dependent frequency and showed that anonadiabatic change <strong>of</strong> frequency may exhibit nonclassicalbehavior such as squeezing and sub-Poissonian statis- where (2.1) is valid if B (t) is uniform. We assum: that'tics. Abdalla [l] investigated statistical properties <strong>of</strong> a the magnetic field B is switched on at time t =O. k is acharged oscillator in the presence <strong>of</strong> a spatially constant unit vector along the symmetry axis; r2=x2+y *, where xbut time-independent magnetic field. He considered sta- and y are two Cartesian components perpendicular to thetistical aspects that, in the realm <strong>of</strong> quantum optics, re- symmetry axis. B (t) and q( t) are arbitrary piecewise',late photon-number fluctuations, bunching and anti- continuous functions <strong>of</strong> time and c is the speed <strong>of</strong> light.bunching, quasiprobability distribution function, etc. The Hamiltonian for the present system is given bySince a charged oscillator in a magnetic field is not thesame as a mode <strong>of</strong> a radiation field in a magnetic field, wewill prefer the term "excitation" instead <strong>of</strong> "photon" tobe consistent with a realm that is not that <strong>of</strong> quantum optics.whereIn the present paper we study statistical properties <strong>of</strong>an excitation for a time-dependent (TD) electromagneticfield. Among the motivations one can cite the problem <strong>of</strong>particle trapping by oscillating fields (cf. Glauber [2]) andothers involving TD Hamiltonians [7].The present paper is arranged as follows In Sec. I1 wedefine the system Hamiltonian. In Sec. I11 we quantizethe system and obiain the Heisenberg equations <strong>of</strong>motion. In Sec. IV we solve the Heisenberg equations for where LZ is the z component <strong>of</strong> the angular momentumour TD system and calculate the variances <strong>of</strong> the quadra- L,=(rXp),=xp,,-ypx andture phase amplitudes. In Sec. V we consider some statisticalproperties for our system and present results for atime-dependent frequency.I1050-2947/95/52(2)/15 18(7)/$06.00 - 52 1518 a1995 The American Physical Society ',


PHYSICAL REVIEW A VOLUME 5 1, NUMBER 3 MARCH 1995Fokker-Planck equation for a single two-level atom:Applications in the bad-cavity limitChangxin Wang and <strong>Reeta</strong> <strong>Vyas</strong><strong>Physics</strong> Department, <strong>University</strong> <strong>of</strong> <strong>Arkansas</strong>, Fa~etteville, <strong>Arkansas</strong> 72701(Received 14 October 1994)A generalized second-order Fokker-Planck equation is derived for a single two-level atom in acavity driven by an external classical field without using system size expansion or any truncation.Effects <strong>of</strong> detuning, as well as atomic and cavity decays, are incorporated in this Fokker-Planckequation. This equation is used to study atomic inversion and the second-order intensity correlationfunction <strong>of</strong> light in the bad-cavity limit by adiabatically eliminating the field variables. Several novelfeatures in atomic inversion and intensity correlations that arise due to atomic and cavity detuningare discussed. Curves are presented to illustrate the behavior.PACS number(s): 42.50.Lc, 42.50.Ct, 42.50.Dv, 42.50.ArI. INTRODUCTIONThe Fokker-Planck equation has played an importantrole in quantum optics. For the problem <strong>of</strong> N two-levelatoms interacting with a cavity field, the Fokker-Planckequation approach is well developed and it has beenwidely used to study optical bistability [I-71, photon antibunching[I-31, and squeezing [8,9]. Haken [lo] used theSudarshan-Glauber P representation [11,12] based on thenormally ordered characteristic function to map an operatormaster equation into a c-number Fokker-Planckequation. This equation, when applied to study opticalbistability, results in a nonpositive diffusion matrix. Thescaling argument, used by Haken to drop the <strong>of</strong>fendingterms for the laser near threshold, does not work for opticalbistability, photon antibunching, and squeezing. Animportant step towards resolving this problem was takenby Drummond and Gardiner [13] who introduced theso called positive-P representation. The Fokker-Planckequation based on the positive-P representation has beenused to study quantum effects in an atomic system consisting<strong>of</strong> N two-level atoms in the good-cavity limit [I]and the bad-cavity limit [2,3]. Another approach basedon the symmetrically ordered Wigner representation hasalso been used for optical bistability, which for opticalbistability leads to a positive diffusion matrix [4,5].The Fokker-Planck equation approach <strong>of</strong>fers some appealingfeatures [6,7]. First, the analogies between classicaland quantum fluctuations can be drawn. This helpsin develodne - intuition for auantum fluctuations. Secuond, all quantum-operator averages can be calculated asclassical-like averages by integrating the correspondingc-number variables multiplied by the distribution function.Third, the Fokker-Planck equation has been studiedextensively in classical statistical physics and mathematicaltechniques developed for analyzing the Fokker-Planck equations for classical systems can be applied tosolve quantum-mechanical problems. Finally, a Fokker-Planck equation can be converted into an equivalent set<strong>of</strong> first-order stochastic differential equations, which canbe solved analytically or numerically.The Fokker-Planck equations derived for N two-levelatomic systems are, unfortunately, not applicable to asingle-atom system. Various representations, such as theSudarshan-Glauber-Haken, Wigner, or positive-P representations,used for mapping a density-operator masterequation into a c-number differential equation alllead to a generalized Fokker-Planck equation which containsderiktives <strong>of</strong> all orders. To obtain a second-orderFokker-Planck equation the system-size expansion is usedto justify the truncation <strong>of</strong> series and retain only terms upto the second-order derivatives. The system-size expansion,however, breaks down when the number <strong>of</strong> atomsN becomes small. That is why most <strong>of</strong> the studies onquantum fluctuations by using the Fokker-Planck equationassume the system size (here, the number <strong>of</strong> atoms)to be large. On the other hand, small systems like a singletwo-level atom are <strong>of</strong> particular interest because in thesmall systems quantum effects are expected to be mostdramatic. So far, to study the quantum effects <strong>of</strong> a singletwo-level atom inside a cavity, the density-operator masterequations have been used by most authors [14-181.In this paper we derive an exact second-order Fokker-Planck equation for a single two-level atom interactingwith a quantized cavity mode. The cavity is drivenby an external classical field. This Fokker-Planck equationincludes the effects <strong>of</strong> atomic detuning, cavity detuning,atomic decay, and cavity decay. Since no partialderivatives beyond second-order exist in our Fokker-Planck equation no system-size expansion or truncationis needed. Our work may <strong>of</strong>fer an alternative for studyingcoupled single-atom field. systems.Section I1 describes our model and the phase space representationused to arrive at the Fokker-Planck equation.In Sec. 111, a generalized second-order Fokker-Planckequation for both field and atomic variables is derived.In Sec. IV, field variables are adiabatically eliminatedfrom the generalized Fokker-Planck equation by usingthe bad-cavity limit. A generalized set <strong>of</strong> Bloch equationsdescribing the dynamics <strong>of</strong> a single atom coupledto the field inside the cavity are derived. These equationsare used to study the steady-state inversion. In Sec.105@2947/95/5 1(3)/25 16( 14)/$06.00 - 51 2516 01995 The American Physical Society


3 1 October 1994<strong>Physics</strong> Letters A 194 (1994) 153-158PHYSICS LETTERS AI!i. .i:' !/1Scattering <strong>of</strong> atoms by light:probing a quantum state and the variance <strong>of</strong> the phase operatorB. Baseia "*I, <strong>Reeta</strong> <strong>Vyas</strong> b2, Cklia M.A. Dantas c3, V.S. Bagnato c*3 Ia Departamento de Fisica, CCEN, Universidade Federal ah Parah. 58059-970 Joio fissoa (PB), BrazilDepamnt <strong>of</strong> <strong>Physics</strong>, <strong>University</strong> <strong>of</strong> <strong>Arkansas</strong>. Fayettcville, AR 72701, USAlnstituto de Fisica de SCo Carlos, Universidade de Sdo Poulo, Cx. Postal 369, 13560-970 Sdo Carlos (SP), BrazilReceived 22 June 1994; revised manuscript received 6 September 194; accepted for publication 12 September 1994Communicated by F?R. Hollandi:f;,FGrI*:-7';AbstractAn experiment is proposed to study the quantum state and the variance <strong>of</strong> the phase operator by measuring the momentumdistribution <strong>of</strong> the atoms resulting from the interaction <strong>of</strong> an atomic beam with a single mode electromagnetic field via atwo-photon interaction. A comparison <strong>of</strong> the results with those obtained for a one-photon interaction is made. Adwt~gesin using the two-photon interaction are pointed outIn the last decade there has been great interest in the deflection <strong>of</strong> atoms by light fields [I]. The momentumtransferred to an atomic beam by a standing wave has been shown to be an important tool in the determination<strong>of</strong> the light field properties [Z-41.Recently Freyberger and Herkornmer [4] proposed a simple and interesting setup to obtain complete informationabout the quantum state <strong>of</strong> a single mode <strong>of</strong> the radiation field. The proposed method uses a barn<strong>of</strong> two-level atoms as a meter for the quantized mode radiation field. Atoms are prepared in a supemsitionState (Jg) + eie Ie)), where Ig) and le) denote ground and excited states <strong>of</strong> the atom respectively, and 6r 18 therelative phase between them. The prepared atomic beam passes through a region near the node <strong>of</strong> a standingelectromagnetic wave <strong>of</strong> wavelength A in a cavity. Atoms interact with the field via a one-phoron interactiondffcribed by the resonant Jaynes-Cumming model. It is assumed that the time <strong>of</strong> interaction is very short sothat spontaneous decay <strong>of</strong> the atom is negligible. The atomic diffraction pattern resulting from the interactionshows asymmetries providing information about the state <strong>of</strong> the field and also about the expectation value <strong>of</strong>the phase operator exp(-$) [4].The proposed one-photon interaction method (OPIM) is interesting, however, for some special fields thecannot provide complete information about the field. As-in the OPIM, we consider the standing lightwave to be a quantum state IP(x, 0)) = C Cmlrn) with the coefficients C,,, <strong>of</strong> a number state (m). For the case' E-mail: basilio@ifqsc.sc.usp.br.2Email: w@24533uafsysb.uark.edu.3E-mail: vandcr@ifqsc.sc.usp.br.0375-9601/94/$f)7.00 @ 1994 Elsevier Science B.V. All rights nservcd0375-9601 (94)00755-1


PHYSICAL REVIEW A VOLUME 47, NUMBER 4 APRIL 1993Homodyne photon statistics <strong>of</strong> the subthreshold degenerate parametric oscillatorA. B. Dodson and <strong>Reeta</strong> <strong>Vyas</strong><strong>Physics</strong> Department, <strong>University</strong> <strong>of</strong> <strong>Arkansas</strong>, Fayetteville, <strong>Arkansas</strong> 72701(Received 17 August 1992)Homodyne statistics when the light from a degenerate parametric oscillator (DPO) is mixed withcoherent light from a local oscillator are discussed. By using dynamical models <strong>of</strong> these beamswe discuss the photon sequences underlying the superposed beam in terms <strong>of</strong> photoelectric pulsesequences recorded by a detector. The field produced by the parametric oscillator is expressed interms <strong>of</strong> two independent real Gaussian random variables. Using this property <strong>of</strong> the field variables,we derive the generating function for the photoelectron statistics analytically. From this generatingfunction, expreseions for the photoelectron-counting distribution, factorial moments, and thewaiting-time distribution are derived. These quantities are directly measurable in photon-countingexperiments. The results reported here are applicable to arbitrary strengths <strong>of</strong> the signal and thelocal-oscillator fields. We also show that, depending on the relative strength and relative phasebetween coherent light from the local oscillator and squeezed light from the DPO, the superposedlight beam may exhibit many interesting nonclassical features. These and other effects are describedand curves are presented to illustrate the behavior.PACS nurnber(s): 42.50.Dv, 42.50.Ar, 42.65.K~I. INTRODUCTIONInterest in the optical degenerate parametric oscillator(DPO) as a vehicle for studying the photon statistics<strong>of</strong> squeezed light [I, 2) has motivated the publication <strong>of</strong>several papers describing the statistics <strong>of</strong> the DPO [3-61.<strong>Vyas</strong> and Singh obtained exact analytical expressions forthe statistics <strong>of</strong> the DPO [3, 41. In this paper we applythe same approach to study the statistics <strong>of</strong> the fieldproduced by superposing the light from the DPO withthe light from a coherent local oscillator (LO) in homedyne detection experiments. In homodyne detection experimentsthe LO frequency is the same as that <strong>of</strong> thesignal. Such experiments have been used for measuringphase-sensitive properties <strong>of</strong> squeezed light [I, 2, 61.Detailed discussions <strong>of</strong> homodyne detection and beamsplittermodeling may be found in the literature 16-17].The light from the DPO produces a large amount <strong>of</strong>squeezing 121. Inside the DPO cavity a pump photonis down converted into two photons at the subharmonicfrequency. These pairs <strong>of</strong> photons are highly correlated,and it is this correlation that results in various interestingnonclassical features that we describe in this paper.In the homodyne detection experiments usually astrong LO is used. This assumption is also implicit in thetheoretical treatments 16-17]. In this limit many effectsdiscussed in this paper are either too small or the technique<strong>of</strong> photoelectron counting cannot be used as thephoton fluxes involved are too high. In our approach wedo not make this strong-local-oscillator approximation.This allows us to treat the case <strong>of</strong> even small photonnumbers and interestingly we find that many quantumeffects that are otherwise small are more pronounced inthis regime. Since low photon numbers are involved photoelectroncounting experiments can be used to measurethese effects.The theoretical investigations <strong>of</strong> the fluctuat~on prop-;erties <strong>of</strong> the light from the DPO operating belaw thresholdshow that this light may exhibit pronounced photonbunching and super-Poissonian statistics 13-81. Thesefeatures by themselves do not reflect the unusual nature<strong>of</strong> the light from the DPO. In this paper we show thatwhen the output light from the DPO is homodyned witha coherent LO field, the superposed field exhibits a vanety<strong>of</strong> interesting statistical features, ranging froq super-Poissonian and highly bunched photon sequences to sub-Poissonian and antibunched photon sequon--. Betweenthese limits the superposed field can show photo11 bun, 11-",ing accompanied by sub-Poissonian statistics, and antibunchingaccompanied by super-Poissonian statistics, indicatingclearly that sub-Poissonian statistics and photonantibunching are distinct effects [8,18-211. Other interestingfeatures are reflected in the waiting-time distribution[22]. The waiting-time distribution for the photonsshows both a minimum and a maximum. The minimum,is found at the time <strong>of</strong> the order <strong>of</strong> the cavity lifetime.This minimum corresponds to a reduced probability <strong>of</strong>detecting two consecutive photons separated by a time<strong>of</strong> the order <strong>of</strong> the cavity lifetime.In addition to the effects described in the precedingparagraph the second-order intensity correlation functiog'shows new violations <strong>of</strong> Schwartz's ineaualities, indicat-.ing the quantum nature <strong>of</strong> the light from the DPO. ~11-.;<strong>of</strong> these features depend on the relative strengths <strong>of</strong> thetwo fields and their phase differences. The superpose$Ifield thus <strong>of</strong>fers a system that allows us to generate a va- ., ,riety <strong>of</strong> nonclassical features simply by turning a virtual"knob" <strong>of</strong> parameters.It is interesting to note that thermd L;;b'r is aisubunched and super-Poissonian, but homodyning the ~1.e::-ma1 field with a local oscillator does not show these non:.classical features [14]. This clearly indicates that the n6n- .01993 The American Physical Society


PHYSICAL REVIEW A VOLUME 48, NUMBER 5 NOVEMBER 1993Single-atom fluorescence with nonclassical lightIrina E. Lyublinskaya and <strong>Reeta</strong> <strong>Vyas</strong><strong>Physics</strong> Department, <strong>University</strong> <strong>of</strong> <strong>Arkansas</strong>, Fayetteville, <strong>Arkansas</strong> 72701(Received 20 May 1993)The statistical properties <strong>of</strong> fluorescent light from a single two-level atom driven by nonclassicallight beams that exhibit antibunching and squeezing are discussed in the weak-field limit. The fieldsproduced in intracavity second-harmonic generation and those produced by superposing the lightfrom a degenerate parametric oscillator with a coherent field are considered to be the models <strong>of</strong>nonclassical light beams. These fields can be described in terms <strong>of</strong> real Gaussian random processeswith nonzero mean coherent components. Analytical expressions for the fluorescent light intensity,the second-order intensity correlation function, and the variance <strong>of</strong> the photon number in the steadystate are obtained. The effects <strong>of</strong> driving-field correlations, detuning, and other parameters characterizingthe atom-field interaction on the statistics <strong>of</strong> scattered light are investigated. It is foundthat for a certain range <strong>of</strong> parameters the second-order intensity correlation function may vanishat nonzero times. Under certain other circumstances the antibunching <strong>of</strong> the fluorescent photons isenhanced. These and other interesting features that appear in the scattered light are discussed.PACS number(s): 42.50.Lc, 32.80.-t, 42.50.Dv, 32.50.+dI. INTRODUCTIONThe phenomenon <strong>of</strong> resonance fluorescence, when asingle two-level atom is illuminated by a light beamwhose frequency is nearly resonant with the transitionfrequency <strong>of</strong> the atom, has been treated by a number<strong>of</strong> workers [I-101. Most treatments assume the incidentlight beam to be in a coherent state. This leads to mathematicallysolvable expressions for photon statistics. Acomplete solution to photon statistics in resonance fluorescenceunder coherent illumination has been given inRef. [5]. Recently, two-level atoms driven by nonclassicalstates <strong>of</strong> light have also been considered [7-101.Carmichael, Lane, and Walls [9] investigated the phenomena<strong>of</strong> subnatural linewidth in the presence <strong>of</strong> broadbandsqueezed light. The interaction <strong>of</strong> a single two-levelatom with finite-bandwidth squeezed light was studied by<strong>Vyas</strong> and Singh [lo]. They considered a degenerate parametricoscillator (DPO) operating below threshold as thesource <strong>of</strong> squeezed light.In this paper we study the interaction <strong>of</strong> a single twolevelatom with an electromagnetic field that is producedby superposing the light from the DPO with the lightfrom a coherent local oscillator (LO) at a beam splitterin a homodyne-detection-type setup. We will referto this light source as HMDPO. Another source <strong>of</strong> lightwe consider is the intracavity second-harmonic generation(SHG). The fundamental beam derived from thissource is used for illuminating the atom. We also studythe effects <strong>of</strong> detuning on the statistical properties <strong>of</strong> thefluorescent photons.The fundamental process underlying the DPO is theconversion <strong>of</strong> a pump photon into two photons at thesubharmonic frequency inside an optical cavity resonantat the subharmonic frequency. Due to the strong correlationsbetween the pairs <strong>of</strong> photons produced in the process<strong>of</strong> down-conversion the light from the DPO is highlybunched and super-Poissonian [ll]. When the light fromthe DPO is superposed with the light from a coherent LOat a beam splitter, the superposed field exhibits a variety<strong>of</strong> quantum features [12], ranging from super-Poissonianand highly bunched photon sequences to sub-Poissonianand antibunched photon sequences. Thus the HMDPOfield allows us to control the statistical properties <strong>of</strong> thedriving field simply by changing parameters such as relativephase and the relative intensities <strong>of</strong> the two superposedbeams.The SHG produces antibunched and sub-Poissonianphoton sequences at the fundamental wavelength. Thenonclassical effect <strong>of</strong> photon antibunching in the secondharmonicgeneration was predicted theoretically [13-151.We assume that both the DPO and the SHG are operatingbelow threshold and consider a fully quantummechanicaltreatment <strong>of</strong> noise in both sources. In bothcases the field can be modeled by two real Gaussian randomprocesses with different variances and correlationtimes [Ill. Because <strong>of</strong> the finite correlation time <strong>of</strong> theincident field, atomic states and field states develop correlationsduring their dynamical evolution. These correlationsrender the problem <strong>of</strong> photon statistics morecomplex for these fields, in general, than in the casefor coherent excitation. For weak driving fields, such asthose produced by the light sources under consideration,it is possible to apply perturbative approach to understandthe behavior <strong>of</strong> a two-level atom when the excitingfield has finite correlation time [lo]. This is the approachtaken in this paper.The paper is organized as follows. In Sec. I1 we summarizethe Heisenberg equations <strong>of</strong> motion governing thetime evolution <strong>of</strong> atomic and field operators for finitedetuning between the atomic transition frequency andthe mean-field frequency. Using the formal solutions <strong>of</strong>these equations we write down general expressions for thescattered light intensity, normalized second-order intensitycorrelation function, and the variance <strong>of</strong> the photoncounting distribution. In Sec. 111 these quantities are cal-1050-2947/93/48(5)/3966( 14)/$06.00 3966 01993 The American Physical Society


PHYSICAL REVIEW A VOLUME 46, NUMBER 1 1 JULY 1992Photon-counting statistics <strong>of</strong> the subthreshold nondegenerate parametric oscillator<strong>Reeta</strong> <strong>Vyas</strong>Department <strong>of</strong> <strong>Physics</strong>, Uniuersity <strong>of</strong>Arkunsos, Foyetteuille, <strong>Arkansas</strong> 72701(Received 9 September 1991)Statistics <strong>of</strong> photons emitted by a nondegenerate parametric oscillator (NDPO) operating belowthreshold are discussed. Analytical expressions for the generating function are obtained when two nondegeneratemodes <strong>of</strong> the NDPO are mixed on a photocathode or when the intensities <strong>of</strong> the two modesare added. From the generating function, photon-counting distributions and waiting-time distributionsas measured by a detector placed outside the cavity are derived. Similarities and differences between thedistributions for the NDPO, the degenerate parametric oscillator, degenerate four-wave mixing, nondegeneratefour-wave mixing, and thermal light are also discussed.PACS number(s): 42.50.Dv, 42.50.Ar, 42.65.K~I. INTRODUCTIONThere has been considerable interest in the fluctuationproperties <strong>of</strong> the nondegenerate parametric oscillator(NDPO) in recent years. In the NDPO a pump photon isdown-converted into two nondegenerate photons insidean optical cavity. These photons come out <strong>of</strong> the cavity,independently, in a lifetime <strong>of</strong> the order <strong>of</strong> the cavity lifetime.Since two nondegenerate photons are producedsimultaneously inside the cavity, the NDPO generateshighly correlated twin beams [I]. This feature <strong>of</strong> theNDPO has been used in reducing the noise in thedifference intensity I, -I, below the shot-noise limit [2].The nondegenerate modes <strong>of</strong> the parametric oscillator arepresent even when the parametric oscillator is operated inthe degenerate mode. It is <strong>of</strong> interest to investigate howfluctuation properties <strong>of</strong> the parametric oscillator changewhen nondegenerate modes are also taken into account.The degenerate parametric oscillator (DPO) is known toproduce large amounts <strong>of</strong> squeezing [3-51. Photon statisticsfor the DPO have been studied by several differentgroups [6,7] by using different techniques. For theNDPO, the mean and the variance <strong>of</strong> the photons comingout <strong>of</strong> the cavity have been calculated by Collett andLoudon [8]. Photon-counting statistics for an idealizedmodel <strong>of</strong> squeezed light have also been studied [9]. Inthis paper a realistic system, an open system with a pump- -and dissipation, for the NDPO is considered, and statistics<strong>of</strong> the photons emitted by the NDPO are discussedwhen the two nondegenerate photons have the same frequency.Here nondegeneracy refers to the direction <strong>of</strong>propagation or polarization. Two cases are considered,one in which the amplitudes <strong>of</strong> the two nondegeneratemodes are homodyne detected, and the second in whichthe intensities <strong>of</strong> the two modes are added together. Inboth cases, analytic expressions for the generating functionsfor photoelectric-counting statistics are obtained byfollowing our earlier approach to the DPO [6]. In thenext section, we show that the field for the NDPO can bedescribed in terms <strong>of</strong> four real Gaussian random processes.In Sec. 111 the generating function for the NDPO isobtained, and photon statistics in terms <strong>of</strong> photoncountingdistribution, factorial moments, and waitingtimedistribution are discussed. These results are comparedwith those for the DPO, degenerate four-wave mixing(DFWM), nondegenerate four-wave mixing (NFWM),and thermal light.11. GAUSSIAN RANDOM VARIABLESOF THE NDPO FIELDThe Hamiltonian, in the interaction picture, for theNDPO where one pump photon is down-converted intotwo nondegenerate photons can be written as [lo]Here pump depletion is assumed to be negligible, which isa good approximation below threshold. Parameter K is amode-coupling constant, E is the dimensionless amplitude<strong>of</strong> the pump beam incident on the cavitthe annihilation operators, and at and a, ?'* are the and creation areoperators for the two down-converted fields <strong>of</strong> theNDPO. a,,,, describes dissipation suffered by thesubharmonic modes due to linear absorption, scattering,and transmission.In order to obtain the equations <strong>of</strong> motion correspondingto the annihilation and creation operators, the masterequation with the interaction Hamiltonian [Eq. (111 is expressedin the positive-P representation [I I]. This leadsto a Fokker-Planck equation for probability distribution<strong>of</strong> the subharmonic mode with a positive diffusion constant.The distribution is defined over a four-dimensionalcomplex space. The four Langevin equations obtainedfrom this Fokker-Planck equation are [I21- 46 395@ 1992 The American Physical Society


,n <strong>of</strong> atomic clocks, wherenegligible [2]. In such de.pathways can couple am<strong>of</strong>microwave fieldt results this would causet frequency. Consequentl,sitic pathway would respility.optics Communications 9 1 ( 1992) 347-353~~flh-HollandTWO-photon interaction <strong>of</strong> a phased atom with squeezed vacuumC. Wang and <strong>Reeta</strong> <strong>Vyas</strong>Deparlmenf <strong>of</strong> <strong>Physics</strong>, Llniversily <strong>of</strong><strong>Arkansas</strong>, Fayeffeville, AR 72701. USAp. ~ambropoulos for:garding the effects Received 3 February 1992; revised manuscript received 17 March 1992t processes. This w;pace Corporation':h program.Two-photon interaction <strong>of</strong> a squeezed-vacuum state with a single atom prepared in a coherent superposition <strong>of</strong> its ground andexcited states is studied. Long time and short time evolution <strong>of</strong> quadrature squeezing is discussed and compared with the correspondingbehavior in the conventional Jaynes-Cummings model. It is found that the two-photon interaction <strong>of</strong> a phased atomwith field in the vacuum slate can yield upto 451 squeezing below the shot-noise limit. With a suitably prepared phased atom thevacuum state <strong>of</strong> the field can evolve periodically into a pure state that is a coherent superposition <strong>of</strong> the vacuum state ando-photon Fock state.cs <strong>of</strong> a two-level atom interacting with a single mode <strong>of</strong> the radiation field in a lossless cavity caned by the Jaynes-Cummings model (JCM) ". This model is perhaps the simplest solvable modelarchi, IEEE Trans. lnstexamples clearly indicate the essential role <strong>of</strong> quantum mechanics in the dynamical evolution <strong>of</strong>nd fluctuations because these features are absent in the semiclassical version <strong>of</strong> this model.atom with a single mode squeezed state within the framework <strong>of</strong> the JCM. The possibility <strong>of</strong> gen-sted by Meystre and Zubairy [3]. Knight and co-workers [ 10,11] have discussed the JCM with astate that is a coherent superposition <strong>of</strong> the vacuum state and the number state. They found that af 25qo squeezing can occur in the JCM if the atom is prepared initially in a coherent superpositionstate or in the excited state.is also examined as a special case and it is found that a maximum squeezing <strong>of</strong> 45Yo below the347


Green's function soluSion to the tissue bioheat equation<strong>Reeta</strong> <strong>Vyas</strong>Department <strong>of</strong> Phprics, Unimity <strong>of</strong> <strong>Arkansas</strong>, Fayetteuille, <strong>Arkansas</strong> 72701M. L. RustgiDepartment <strong>of</strong> Phpsics, StateU.'niuersi@ <strong>of</strong> New York at Buffalo, Buffalo, New York 14260[Received 18 October 1991;; accepted for publication 13 May 1992)A Green's function solutiw to the tissue bioheat equation including blood flow in cylindricalgeometry is obtained. Nlmerical results for temperature variation in the bovine muscle arereported when @he tissue ias exposed to neodymium-yttrium-aluminum garnett (Nd:YAG)hers with Gaussian prae and a comparison with recent measurements is made. A strongdependence <strong>of</strong> the tissue mperature on the beam radius and pulse time is found.I. INTRODUCTIONThe influence <strong>of</strong> lasers in biologicdamd medical research isbecoming increasingly important"-8 The neodymiumyttrium-duminumgarnett (Nd:YLLG) laser, which is capable<strong>of</strong> being operated as eithar pulsed or continuouswave (cw), with pulse widths vaqkag, with mode <strong>of</strong> operation,from :lanoseconds to picosamnds and power to theorder <strong>of</strong> 10" W has eclipsed the ruse <strong>of</strong> ruby laser. TheNd:YAG laser finds most <strong>of</strong> its -litxition in the utilization<strong>of</strong> hs 1064-nm light. The k tissue interaction ismainly determined by two param-, the interaction time<strong>of</strong> the radiation with the tissue a~ld the effective energydensity which brings about an eEcct whereby the tissuespecificabsorption must be taken $into consideration. Atlow qgy density with a long expsure the absorption <strong>of</strong>lighL a .manly leads to photochedcal processes. With decreasinginteraction time and higher energy densities is thedomain <strong>of</strong> the photothermal-indd effects. In this paper,a Green% function solution to thehue bioheat equation isobtained to describe the tempera distribution due to alaser beam with Gaussian pr ~2- k The effect <strong>of</strong> thermalconduction on the tewrd$re k y curve within theframework <strong>of</strong> the bioheat equationhas been investigated by~andhu,~ who pointed out that inbnited cases, the decaycurve may be used to measure Wood flow. Though thesolution <strong>of</strong> the bioheat equation OPH excluding blood flowhas beem obtained by a number d ;authors?-' the Green'sfunction approach used here is qifte general and includesthe blood flow in the sdution <strong>of</strong> dne bioheat equation.In Sec. I1 simple analytical expressions for Green'shnction and temperature distribtiion using a model withcylindrical symmetry are derived The heat source is assumedbo be a pulsed laser beam with a Gaussian pr<strong>of</strong>ile.The method is sufficiently general tto enable calculation <strong>of</strong>temperature distribution for other sources with cylindricalsymmetry such as cw laser usingthis Green's function. InSec. 111 results are compared with available experimentaldata.'is derived. The derivation is based on a cylindrically symmetricmodel in which the laser beam is traveling in the zdirection. The temperature distribution for this system isdescribed by a differential equation which is essentially anequation <strong>of</strong> energy conservation. The Green's function approachis followed to solve the differential equation for thetemperature distribution. Since the Green's function obtainedfor the differential equation is independent <strong>of</strong> thesource term, the same Green's function can be used tocalculate temperature distribution for various sources withdifferent spatial and temporal pr<strong>of</strong>iles. Even though thisproblem is solved for a cylindrically symmetric system, thisapproach can be extended to calculate the temperature distributionsfor systems that do not possess cylindrical symmetry.The bioheat equation was first proposed by pennesg andthe inherent assumptions in this equation are outlined inHodson et aL lo The differential equation describing thetemperature distribution can be written as9.''where the first and the second terms on the right-hand side<strong>of</strong> Eq. (1) are due to the thermal diffusion and blood flow,the temperature above the baseline temperature <strong>of</strong> the tissueis T(r,z,t), S(r,z,t) is the power deposited per unitvolume due to the absorption <strong>of</strong> the laser energy, the ther-mal diffusivity D is equal to K/~C, where K, p, and C arethermal conductivity, density, and specific heat <strong>of</strong> the tissue,respectively. The parameter b is given by opbCJC.Here o is volumetric blood flow, Cb is the specific heat <strong>of</strong>the blood, and pb is density <strong>of</strong> the blood. The boundaryconditions imposed on the temperature and its derivativesare that they vanish as r and z go to infinity. These boundaryconditions require that the source energy must vanishat infinity. It is also assumed that at time t=0, the temperatureis zero. The solution <strong>of</strong> Eq. ( 1 ) can be expressedin terms <strong>of</strong> Green's function G(r,z,t; r1,z',t') and thesource term S(r,z,t) as"11. P"F&N'S FUNCTION AND TEMPERATUREDl: t6BUTION S(r1,z',t') G( r,z,t;r1,z',t' )In this section, an expression fot the temperature distributionT(r,z,t) in a tissue that is a~osed tothe laser beam x r'dr'dz' dt'. (2)1319 Hed. Phys. 19 (5). Sep/Oct lXB2 0994-2405/91/5131406S01.20 @ 1992 Am. Assoc. Phys. Med. 1319


PHYSICAL REVIEW A VOLUME 45, NUMBER 11 I JUNE 1992Resonance fluorescence with squeezed-light excitation<strong>Reeta</strong> <strong>Vyas</strong> and Surendra SinghDepartment <strong>of</strong> <strong>Physics</strong>, <strong>University</strong> <strong>of</strong> <strong>Arkansas</strong>, Fayetteuille, <strong>Arkansas</strong> 72701(Received 30 August 199 1)Resonance fluorescence from a single two-level atom driven by a beam <strong>of</strong> squeezed light is studied inthe weak-field limit. We consider the situation where the atom is coupled to the ordinary vacuum andonly a few field modes corresponding to the driving field are squeezed. The field produced by the degenerateoptical parametric oscillator is used as the driving field. Heisenberg equations <strong>of</strong> motion are solvedin the steady-state and analytic expressions for the fluorescent-light intensity and the spectrum <strong>of</strong>fluorescent light are derived. We also consider photon statistics <strong>of</strong> fluorescent light. In particular,squeezing, antibunching, and sub-Poissonian statistics <strong>of</strong> fluorescent photons are discussed, and analyticexpressions for the quadrature variance and the two-time intensity correlation function are presented.Contrary to the case <strong>of</strong> coherent excitation, the second-order intensity correlation function does not factorize.This and other differences are discussed, and curves are presented to illustrate the behavior <strong>of</strong>various quantities. We also present results for thermal excitation <strong>of</strong> the atom.PACS number(s): 42.50.Lc, 32.80. - t, 42.50.Dv, 32.50. +dA single two-level atom interacting with an electromagneticfield is a fundamental model <strong>of</strong> quantummechanics. This simple dissipative quantum system liesat the heart <strong>of</strong> the physics <strong>of</strong> light-atom interaction. Aninteresting aspect <strong>of</strong> this problem is the phenomenon <strong>of</strong>resonance fluorescence when the atom is illuminated bv alight beam whose frequency is nearly resonant with thetransition frequency <strong>of</strong> the atom. The fluorescent lightunder these conditions displays many purely quantummechanicalfeatures [l]. These features are most clearlyreflected in the photon statistics <strong>of</strong> fluorescent light [2,3].The problem <strong>of</strong> photon statistics in resonance fluorescencehas been treated by a number <strong>of</strong> workers usingseveral different techniques [4]. Most treatments assumethe incident light beam to be in a coherent state. A completesolution to photon statistics in resonance fluorescenceunder coherent illumination has been given recently[5]. Other models <strong>of</strong> the incident light beam that takeinto account the fluctuations <strong>of</strong> incident light have alsobeen considered. These include the phase-diffusion model[6], the chaotic-field model [7], and the jump models [8]<strong>of</strong> phase and amplitude fluctuations. More recently, nonclassicalstates <strong>of</strong> the driving field have also been considered.Gardiner considered the ~roblem <strong>of</strong> radiativedecay in the presence <strong>of</strong> broadband (white-noise)squeezed light [9]. The possibility <strong>of</strong> a subnaturallinewidth in resonance fluorescence under similar conditionshas been investigated by Carmichael, Lane, andWalls [lo]. More realistic models <strong>of</strong> squeezed light,where only a few modes are saueezed (colored saueezedlight), have been considered b; Ritsch and ~olle; in thediscussion <strong>of</strong> the atomic absorption spectrum 1111. Inthreshold [12,13]. This field can be modeled by two realGaussian processes with different variances and correlationtimes [14]. Because <strong>of</strong> the finite correlation time <strong>of</strong>the incident field, atomic states and field states developcorrelations during their dynamical evolution. Thus, unlikethe case <strong>of</strong> coherent excitation, where the factorization<strong>of</strong> correlation functions [2,3,5,15] leads to asimplified description, the problem <strong>of</strong> photon statisticsbecomes complex in the present case. Nevertheless, forweak driving fields appropriate to subthreshold degenerateparametric oscillators, it is still possible to gainsome insight into the behavior <strong>of</strong> a two-level atom whenthe exciting field is squeezed.In Sec. I1 we derive the equations <strong>of</strong> motion governingthe time evolution <strong>of</strong> atomic and field operators. Solutionsto these equations are used to discuss the time evolution<strong>of</strong> fluorescent-light intensity in Sec. 111. The spectrum<strong>of</strong> scattered light is calculated in Sec. IV. Photonstatistics and the two-time intensity correlation function<strong>of</strong> scattered light are discussed in Sec. V, and Sec. VIpresents results for thermal beam excitation <strong>of</strong> the atom.Finally, the principal results <strong>of</strong> this paper are summarizedin Sec. VII.11. EQUATIONS OF MOTIONConsider a two-level atom having energy states 11) and12) separated by an energy gap fh, and interacting withan electromagnetic field via a dipole interaction. TheHamiltonian for this system, in the rotating-wave approximation,isB=tio$,+io,+-[A'-'(~,t)a-ct)-A(+)co,t)a+ct)]this paper we discuss the interaition <strong>of</strong>a single ~wo~level +& . (1)atom with a finite-bandwidth squeezed light. The model<strong>of</strong> squeezed light that we adopt corresponds to light froma degenerate parametric oscillator operating belowHere f3F represents the energy <strong>of</strong> the electromagneticfield. R,(t), b,(t), and f7,(t) are three dynamical spin45 8095 @ 1992 The American Physical Society


volume 79, number 1,2 OPTICS COMMUNICATIONSI October 1990Two-mode ring laser with a saturable absorberDepartment <strong>of</strong> <strong>Physics</strong>. <strong>University</strong> <strong>of</strong><strong>Arkansas</strong>, Faye1teville. AR 72701. USAReceived 9 November 1989Deterministic as well as fluctuation properties <strong>of</strong> the mode intensities in a two-mode bidirectional ring laser with a saturableabsorber (TRSA) are studied. Both homogeneously and inhomogeneously broadened media are considered. Equations <strong>of</strong> motionfor the field amplitudes are derived by using Lamb's semiclassical approach to laser theory. Stability <strong>of</strong> various steady states inthe multidimensional parameter space is discussed. It is found that for a certain range <strong>of</strong> operating parameters a homogeneouslybroadened laser may exhibit tetrastability and an inhomogeneously broadened laser may exhibit tristability. In the former casethe intensity probability density for mode shows a three peak structure.a saturable absorber. We will see that the addition <strong>of</strong>a saturable absorber can lead to many new phenom-We consider both the deterministic and statisticalith a absorber can exhibit many oms. In case <strong>of</strong> similar atoms the level <strong>of</strong> excitation,distinguish the two cells. The amplifying and absorbingmedia are physically separated from one anotherso that they interact only through the cavityfield. As a result the total polarization driving thelaser field can be written as the sum <strong>of</strong> the polarivityabsorbers have also been used inzations <strong>of</strong> the amplifying and absorbing media. Forilization and ~ - ~ ~ i ~ ~ h an i absorbing[31. ~ ~ medium that saturates faster than thesers with saturable absorbers have been gain medium polarization terms Up to fifth order inentally [7]. Much less attention hasultimode lasers with a saturable abafirst step in this direction we con-bility <strong>of</strong> steady states. We have followed Lamb'ssemiclassical approach [ 10,11] to the laser theory toobtain polarization terms up to fifth order. We con-


PHYSICAL REVIEW A VOLUME 42, NUMBER 1 1 JULY 1990Photon-counting statistics <strong>of</strong> the subthreshold transient degenerate parametric oscillator<strong>Reeta</strong> <strong>Vyas</strong> and A. L. DeBrito*Department <strong>of</strong> <strong>Physics</strong>, <strong>University</strong> <strong>of</strong> <strong>Arkansas</strong>, Fayetteville, <strong>Arkansas</strong> 72701(Received 13 February 1990)Transient photoelectron-counting statistics <strong>of</strong> a degenerate parametric oscillator (DPO) operatingbelow threshold are studied. A generating-function technique has been used and analytical expressionsfor photon-counting distribution, factorial moments, and waiting-time distribution have beenderived. These results are compared with the steady-state results for the DPO and thermal light.Recently there has been considerable interest in thesqueezed light generated by a degenerate parametric oscillator(DPO) below threshold.' This system is known toproduce the largest amount <strong>of</strong>The fundamentalprocess in the DPO consists <strong>of</strong> the downconversion<strong>of</strong> a pump photon <strong>of</strong> frequency 20 into twophotons at the subharmonic frequency o. Quantum statisticalproperties for an idealized squeezed state havebeen discussed by many author^.^,^ For the DPO meanzero and variances <strong>of</strong> photons escaping the cavity havebeen calculated by Collett and ~oudon~ and photoncountingstatistics for small counting times have been calculatedby Agarwal and dam.' Analytical results forphoton statistics including dissipation for the DPO belowthreshold have been calculated by <strong>Vyas</strong> and Singh.' Wolinskyand Carmichae19 have followed a numerical approachto obtain photon-counting statistics <strong>of</strong> the DPO.Photon-counting distribution for the subthreshold DPOis found to exhibit even-odd oscillations for countingtimes greater than a few cavity life times. However, it isinteresting to note that even though photons are createdin pairs inside the cavity, intracavity photon number distributionin the steady state does not show even-odd oscillation~.~~'~On the other hand, the transient intracavityphoton number distribution does exhibit even-odd oscillations,"suggesting that transient DPO can exhibitdifferent features than those exhibited by the DPO in thesteady state.In this paper we discuss quantum-statistical properties<strong>of</strong> the DPO in the transient regime, that is, during itsevolution from the vacuum state towards the steady state.We concentrate on the statistical properties <strong>of</strong> photonsemitted by the cavity. These properties can be studied bya detector placed outside the cavity.In Sec. I1 we obtain the generating function for theDPO below threshold in the transient regime. From thisgenerating function we obtain exact analytical expressionsfor the photon-counting distribution, factorial moments,and waiting-time distributions in Sec. 111. Theseresults for the DPO in the transient regime are then comparedwith those for thermal light and the DPO in thesteady state. A summary <strong>of</strong> the results <strong>of</strong> this paper ispresented in Sec. IV.11. EQUATION OF MOTIONAND GENERATING FUNCTIONIn the interaction picture the Hamiltonian for theDPO operating below threshold can be written asI2Here pump depletion is assumed to be negligible. K is themode coupling constant and E is the dimensionless amplitude<strong>of</strong> the pump beam incident on the cavity. a and a tare the annihilation and creation operators for thesubharmonic field and file,, describes losses suffered bythe subharmonic mode due to linear absorption, scattering,and transmission. It has been shown that in the positiveP representationI3 subharmonic field produced by theDPO can be described in terms <strong>of</strong> two independent, realGaussian random variables u , and u These randomvariables satisfy the following equations:ui=-hiui+~~2qi(t) , (2)where i can take values 1 or 2 and q (t) and q2( t ) are 6-correlated, real Gaussian white-noise processes withThe decay constants h, and A2 are given byHere (2y )- ' is the cavity lifetime at the subharmonic frequency.Below threshold (A > I K E ~ ) both the decay constantsare positive and the solutions <strong>of</strong> Eqs. (2) can bewritten asIn deriving this solution we have assumed that at timet =O the oscillator starts in the vacuum state with u ,(O).=O=u,(O). From Eqs. (3) and (5) we find592 01990 The American Physical Society


PHYSICAL REVIEW A VOLUME 40, NUMBER 9 NOVEMBER 1, 1989Photon-counting statistics <strong>of</strong> the degenerate optical parametric oscillator<strong>Reeta</strong> <strong>Vyas</strong> and Surendra SinghDepartment <strong>of</strong> <strong>Physics</strong>, <strong>University</strong> <strong>of</strong>drkansas, Fayetteville, <strong>Arkansas</strong> 72701(Received 22 May 1989)Nonclassical light beams generated by the degenerate optical parametric oscillator operatingbelow threshold are analyzed in terms <strong>of</strong> photoelectron-counting sequences. The positive-P representationis used to calculate the generating function for photoelectron statistics in a closed form.This generating function is used to derive expressions for the photoelectron-counting and waitingtimedistributions. The dependence <strong>of</strong> these distributions on mean photon number inside the cavityand efficiency <strong>of</strong> detection is studied. The relationship between photoelectron-counting sequenceand the photon emission sequence is used to present a simple physical picture <strong>of</strong> light beams producedby the degenerate parametric oscillator.I. INTRODUCTIONSqueezed states <strong>of</strong> light have been observed in a variety<strong>of</strong> physical systems.'-) These states do not admit a positivenonsingular diagonal representation in terms <strong>of</strong>coherent states and are, therefore, an example <strong>of</strong> nonclassicalstates <strong>of</strong> the electromagnetic field. Since squeezingonly refers to the variance <strong>of</strong> the two quadrature components<strong>of</strong> the electric field, it does not filly characterizethese states. With experimental realization <strong>of</strong> thesestates, increasing attention is being paid to theirquantum-statistical properties.435 These properties foridealized squeezed states are well kn~wn.~-~The systemsin which squeezed states have been observed experimentallyare dissipative nonlinear systems, and photon statisticalproperties <strong>of</strong> squeezed states produced by thesesystems have received much less attention.The largest amount <strong>of</strong> squeezing has been observed inan optical parametric oscillator (OPO) operating belowthre~hold.~.' This simple dissipative quantum system hasplayed an important role in recent studies <strong>of</strong> squeezing.In an OPO (Ref. 10) a strong pump beam interacts with anonlinear crystal and is frequency down-converted intotwo beams <strong>of</strong> smaller frequencies inside an optical cavity.If the two beams produced in down conversion have thesame frequency, then the oscillator is termed a degenerateparametric oscillator (DPO); otherwise it is termeda nondegenerate parametric oscillator (NDPO). Aquantum-mechanical treatment <strong>of</strong> the OPO is <strong>of</strong> courseessential since it generates light with nonclassical properties.For an oscillator a distinction must be made betweenintracavity photon statistics and the statistics <strong>of</strong> photonsemitted by the cavity. Intracavity statistics are notdirectly observable. The statistics <strong>of</strong> photons emitted bythe cavity can be measured in photon-counting experiments.The statistics <strong>of</strong> the field inside and outside thecavity are, <strong>of</strong> course, related. Many recent studies <strong>of</strong> thequantum-statistical properties <strong>of</strong> the DPO have centeredaround the calculation <strong>of</strong> the spectrum <strong>of</strong> squeezing"*'2inside and outside the cavity because <strong>of</strong> the subtleties involvedin the detection <strong>of</strong> squeezed light. Intracavityfield statistics were discussed by Drumrnond, McNeil,and ~alls" by using the complex-P representation andby Graham by using the Wigner function'). More recently,Wolinsky and Carrni~hael~*'~ have provided a completedescription <strong>of</strong> the quantum-statistical properties <strong>of</strong>the intracavity field by using the positive-P representation.For the photons escaping the cavity, the mean andvariance <strong>of</strong> photon counts have also been calculated byCollett and ~oudon.'~In this paper we discuss the quantum-statistical properties<strong>of</strong> photon beams generated by an OPO as measuredby a detector placed outside the cavity. These propertiescan be studied in photoelectric-counting and correlationexperiments with low-intensity light beams appropriatefor an OPO below threshold. From the measured photoelectronstatistics, photon statistics <strong>of</strong> the incident lightbeam can be derived. For a detector <strong>of</strong> unit efficiencyeach photodetection corresponds to an emission <strong>of</strong> a photonby the cavity. In this case, the photoelectric-countingsequence and the photon emission sequence areequivalent. We begin by expressing the photoelectroncountingstatistics in terms <strong>of</strong> a generating function inSec. 11. The statistics <strong>of</strong> the waiting time between successivephotoelectric counts can also be derived from thesame generating function. In Sec. I11 the c number equations<strong>of</strong> motion for the DPO operating below thresholdare presented. This is done by using the positive-P representation.The solutions to these c-number equations areused to obtain a closed form expression for the generatingfunction. From this generating function exact expressionsfor the photoelectron-counting distribution and thewaiting-time distribution are derived in Sec. IV. Intracavityphotonstatistics are discussed in Sec. V. We concludeby summarizing the principal results <strong>of</strong> the paper inSec. VI.11. THE GENERATING FUNCTIONConsider a photoelectric detector illuminated by a stationaryweak beam <strong>of</strong> light. The probability p(m, T) <strong>of</strong>detecting m photoelectric counts at the output <strong>of</strong> thedetector in a time interval T is given by1'5147 @ 1989 The American Physical Society


1110 OPTICS LETTERS / Vol. 14, No. 20 / October 15,1989Quantum statistics <strong>of</strong> broadband squeezed light<strong>Reeta</strong> <strong>Vyas</strong> and Surendra SinghDepartment <strong>of</strong> <strong>Physics</strong>. <strong>University</strong> <strong>of</strong> <strong>Arkansas</strong>. Fayetteville, <strong>Arkansas</strong> 72701Received May 8,1989; accepted August 2.1989Photon-emission sequences in nonclassical light beams generated by an optical parametric oscillator are studied.Exact photon-counting and waiting time distributions are derived by using a generating function technique basedon the positive-P representation. This approach is used to present a simple physical picture <strong>of</strong> photon emissionsfrom this nonclassical light source.Squeezed states <strong>of</strong> light are nonclassical states <strong>of</strong> lightthat have been experimentally realized recently in avariety <strong>of</strong> physical systems.ll2 Since squeezing refersonly to the variance <strong>of</strong> the quadrature components <strong>of</strong>the electric field, the need to characterize these statesin terms <strong>of</strong> their photon-counting distributions hasbecome urgent. These distributions for idealizedsqueezed states are well known.3 As a practical devicefor producing squeezed light, the performance <strong>of</strong> thedegenerate optical parametric oscillator (DPO) hasbeen un~urpassed.~In this Letter we present an analytic treatment forthe counting statistics <strong>of</strong> photon beams generated bythe DPO. Our approach leads to exact expressions forboth the counting and waiting time distributions.When these distributions are analyzed they lead to aclear picture <strong>of</strong> photon beams generated by the DPO.The DPO is modeled by two resonant modes <strong>of</strong> anoptical cavity, with frequencies 20 and 0, interactingwith each other through an intracavity nonlinear crystal.The pump mode 20 is excited by an injectedclassical signal <strong>of</strong> amplitude C. In the interaction picturethe Hamiltonian with perfect phase matchingreadswhere b and at and 6 and 6t are the annihilation andcreation operators for the subharmonic mode w andthe pump mode 20, respectively, K is the mode-cou-pling constant, r is the cavity linewidth at the pumpfrequency, and BIo8, describes linear absorption andscattering losses suffered by the modes inside the crystaland the cavity mirrors. By a suitable combination<strong>of</strong> phases, K and E have been chosen to be real.The Hamiltonian in Eq. (I) is used to derive anoperator master equation for the density matrix <strong>of</strong> thefield. This operator equation is converted into anequivalent set <strong>of</strong> classical Langevin equations for thecomplex field amplitudes by using the positive9repre~entation.~~~ The resulting equations for thesubharmonic mode below threshold, where pump depletionis negligible, arewhere 7 is the cavity linewidth at the eubharmonicfrequency, [l(t) and 52(t) are two real statistically independentGaussian white-noise processes with zeromean and unit intensity, and a! and a!* are two complexvariables associated with operators b and bt in thepositive-P representation. Unlike in the diagonal-Prepresentation, a! and a!* are not complex conjugates.The oscillator threshold is defined by KE = y. In thisLetter we restrict our analysis to the case KE < y correspondingto the experiments <strong>of</strong> Ref. 2. Equations (2)ensure that the variabIes a and a* are real in the steadystate because any imaginary parts to a! and a!* decayaway. It follows that the realvariables ul = (a! + a!*)/2and u2 = (a - a!*)/2 are statistically independentGaussian stochastic processes with zero mean and correlationfunctionswhere Xlnz = y r KC are positive decay constants. Thisequation completely determines the statistical properties<strong>of</strong> a! and a* and ul and uz.The generating function with parameter s for photon-countingstatistics as measured by a photodetectorplaced outside the cavity is given bywhere T is the counting interval, 0 I q I I is theefficiency <strong>of</strong> detection, f(t) is the photon number fluxoperator (number per second) for the Iight beam generatedby the oscillator, and T-: denote the time andnormal ordering <strong>of</strong> operators inside the colons. Onusing the positive-P representation the normally orderedoperator average in Eq. (4) becomes1 10-X exp 2sqy u?(t)dt 9 (5)where we have used the correspondence f(t) -. 2ya!a!*.The factorization in Eq. (5) is a consequence <strong>of</strong> thestatistical independence <strong>of</strong> the variables ul and uzexpressed by Eq. (3). Since ul and ui are real Gaussianvariables with an exponential correlation function,the averages in Eq. (5) may be evaluated by making a1)0146-9592/89/201110-03$2.00/0 O I989 Optical Society <strong>of</strong> America


Pulsed and cw photothermal phase shift spectroscopy ina fluid medium: theoryB. Monson, <strong>Reeta</strong> <strong>Vyas</strong>, and R. GuptaA theoretical treatment <strong>of</strong> photothermal phase shift spectroscopy in a fluid medium for the most generalconditions is given. The medium is assumed to be flowing. Results for a stationary medium appear as aspecial case. Both pulsed and cw excitation are considered. For pulsed excitation, the results are valid forexcitation pulses <strong>of</strong> arbitrary length. For the cw case, modulated excitation is explicitly considered, and theresults for unmodulated excitation appear as a special case.I. IntroductionPhotothermal spectroscopy is currently an activearea <strong>of</strong> research, and a wide range <strong>of</strong> applications <strong>of</strong>this technique has been developed.' In this technique,a laser beam (pump beam) is partially or fullyabsorbed by a medium <strong>of</strong> interest. In the presence <strong>of</strong>fast quenching collisions, most <strong>of</strong> this energy appearsas heat, and the refractive index <strong>of</strong> the laser-irradiatedregion is modified. This change in the refractive indexcan be monitored by a second and weaker laser beam(probe beam) in several different ways. For example,the nonuniform refractive index produced by the absorption<strong>of</strong> the pump beam can be detected by thedeflection <strong>of</strong> the probe laser beam passing through themedium. This technique is called photothermal deflectionspectrscopy (PTDS).%3 If the refractive index<strong>of</strong> the medium has a nonzero curvature, the mediumalso acts like a lens, and a probe beam passing throughthe medium changes shape. This can be detected as achange in the intensity <strong>of</strong> the probe beam passingthrough a pinhole. This technique <strong>of</strong> monitoring thephotothermal effect is called photothermal lensingspectroscopy (PTLS).4-6 One may also detect thephotothermal effect by placing the medium in onebeam <strong>of</strong> an interferometer (or inside a Fabry-Perotcavity). The change in refractive index causes a fringeshift which can conveniently be detected as an intensitychange <strong>of</strong> the central fringe. We shall refer to thisThe authors are with <strong>University</strong> <strong>of</strong> <strong>Arkansas</strong>, <strong>Physics</strong> Department,Fayetteville, <strong>Arkansas</strong> 72701.Received 28 September 1988.0003-6935/89/132554-08$02.00/0.O 1989 Optical Society <strong>of</strong> America.technique as photothermal phase shift spectroscopy(PTPS), and it is the subject <strong>of</strong> this paper. Thistechnique was discovered by Stone7 and by Davis andPetuchow~ki,~.~ who refer to this technique as phasefluctuation optical heterodyne (PFLOH) spectroscopy.Many applications <strong>of</strong> this technique have beendiscussed in theIn this paper we give a theoretical treatment <strong>of</strong>PTPS in fluids for the most general conditions, that is,for a flowing medium. Results for a stationary mediumappear as a special case. Moreover, both pulsedand cw excitation are considered. For the pulsed case,the results are valid for excitation pulses <strong>of</strong> arbitrarylength. For the cw case, modulated excitation is consideredexplicitly, and the results for unmodulatedexcitation appear as a special case. Both collinear andtransverse geometries are considered. Davis and Petuchowskighave given a theoretical treatment <strong>of</strong> thissubject previously. However, their results are validonly for a stationary medium, collinear geometry, andin the pulsed case, for delta function excitation only.Therefore, our results have a much more general applicability.Davis and Petuchowski have considered,however, the effect <strong>of</strong> nonzero relaxation times, whereaswe have assumed that all relaxation times are negligiblecompared to the thermal diffusion and convectiontimes. In this aspect, Davis and Petuchowski'sresults are more general than ours.Expressions for the temperature distribution due tocw and pulsed excitation <strong>of</strong> a flowing medium arederived in Sec. 11. The explicit case <strong>of</strong> heating by apump laser beam is considered, which is generally theexperimental situation. However, the heating <strong>of</strong> themedium may be produced by any suitable radiation,e.g., microwaves.14 Expressions for the PTPS signalsare derived in Sec. 111. The explicit case <strong>of</strong> a MichelsonInterferometer is considered. Expressions forother types <strong>of</strong> interferometer either are identical (e.g.,2554 APPLIED OPTICS I Vol. 28, No. 13 1 1 July 1989


PHYSICAL REVIEW A VOLUME 39, NUMBER 3 FEBRUARY 1, 1989Photoelectron waiting times and atomic state reduction in resonance fluorescenceH. J. Carmichael, Surendra Singh, <strong>Reeta</strong> <strong>Vyas</strong>, and P. R. Rice*Department <strong>of</strong> <strong>Physics</strong>, Uniuersity <strong>of</strong> <strong>Arkansas</strong>, Fayetteuille, <strong>Arkansas</strong> 72701(Received 12 September 1988)Photoelectron counting sequences for single-atom resonance fluorescence are studied. The distribution<strong>of</strong> waiting times between photoelectric counts is calculated, and its dependence on drivingfieldintensity and detection efficiency is discussed. The photoelectron-counting distribution is derivedfrom the waiting-time distribution. The relationship between photoelectron counting sequencesand photon emission sequences is discussed and used to obtain an expression for the reducedstate <strong>of</strong> the atom during the waiting times between photoelectric counts. The roles <strong>of</strong> irreversibilityand the observer in atomic state reduction are delineated.I. INTRODUCTIONThe fluorescent photons emitted by a single coherentlydriven two-level atom exhibit the nonclassical property <strong>of</strong>photon antib~nchin~.'-~ The antibunching <strong>of</strong> fluorescentphotons is seen in temporal correlations betweenphotoelectric counts; the detection <strong>of</strong> one photon makesthe detection <strong>of</strong> a second, after just a short delay, improbable.Photon antibunching is traditionally defined interms <strong>of</strong> the degree <strong>of</strong> second-order temporal coherenceg'2'(r,r +T). This is the joint probability for recordingphotoelectric counts in the intervals [r,t +At) and[r +~,t +r+At), normalized by the probability for twoindependent photoelectric counts. For antibunched lightthe joint probability for recording photoelectric countsclosely spaced in time falls below the probability for statisticallyindependent counts (separated by a time longerthan the coherence time); thus, gt2'(t,t) < 1.The antibunching <strong>of</strong> fluorescent photons is alsoreflected in the sub-Poissonian character <strong>of</strong> the probabilitydensity p(n,r,t +T) for recording n photoelectriccounts in the interval [t, t + T).~ p (n,t,t + T) can be derivedfrom g'2'(t,t +T), although the detailed algebraicrelationship is quite complicated. Both g(')(t,t +T) andp (n, t, t + T) have been calculated for single-atom resonancefluorescence by a number <strong>of</strong> workers.'-lo Because<strong>of</strong> the complexity <strong>of</strong> general expressions in the timedomain, some workers only give the Laplace transformfor the photoelectron counting distribution, or give explicittime-dependent expressions only for limiting cases,such as short and long counting times.Recent theoretical work on "quantum jumps""-'2 hasdrawn attention to the distribution <strong>of</strong> waiting times betweenphoton emissions as another useful quantity forcharacterizing photon statistics-in terms <strong>of</strong> measuredquantities, the distribution <strong>of</strong> waiting times between photoelectrons.By "waiting time" we mean the time T betweena photoelectric count recorded at time t, and thenext, recorded at time t +T. If photoelectron sequencescan be described by a Markov birth process, a single conditionalprobability density w ( T I < ) specifies the distribution<strong>of</strong> waiting times between every pair <strong>of</strong> photoelectrons.We call this the photoelectron waiting-time distri-bution. Photoelectron waiting times for coherent lightare exponentially distributed.13 Antibunching impliesthat photons tend to be separated in time. The distribution<strong>of</strong> waiting times should then tend to peak around theaverage time between photoelectric counts.Photoelectron waiting times are certainly not new tothe field <strong>of</strong> photon statistics. Indeed, when a time-toamplitudeconverter is used for a delayed coincidencemeasurement, the raw data provide the distribution <strong>of</strong>waiting times between photoelectric counts. However,when the count rate is sufficiently low, this distribution isproportional (aside from dead-time corrections) tog'2'(t,t +T). l4 This relationship provides the techniqueused to measure g'2)(t,t +T) in the experiments <strong>of</strong> Kimbleef & on photon antibunching in resonance fluorescence.Thus, the waiting-time distribution and its relationshipto g'2)(t,t +T) are known. But the waiting-timedistribution has not been mentioned until re~entl~'~''"'~in the large theoretical literature on resonance fluorescence.This is a deficiency, since it provides a clearerphysical picture <strong>of</strong> photon emission sequences, correspondingphotoelectron counting sequences, and theirnonclassical properties, than g'"(t,t +TI. In this paperwe revisit the problem <strong>of</strong> single-atom resonance fluorescenceand focus attention on the waiting-time distribution.(We will discuss waiting times between photonemissions as well as between photoelectrons. When thedistinction is not important we simply refer to "the waitingtimes" or "the waiting-time distribution.")There are probably two main reasons for the lack <strong>of</strong> at-tention paid to w (TI t) in early work on resonance fluorescence.The first is that g(2)(r,r +T), not w(T(~), is thequantity accessible to measurement. It might be asked,why not use a time-to-amplitude converter to measurethe quantity it gives directly, the photoelectron waitingtimedistribution w (?It)? The problem is that photoelectricdetection is very inefficient. The average time betweenphotoelectric counts is unavoidably much longerthan the correlation time <strong>of</strong> the fluorescent light. Thenw ( ~(t) is proportional to g'2'(t, t +T); w (rlt) can be measured,but only when it effectively reduces to g'2'(t,t +TI.Actually, the proportionality between these quantitiesdoes not hold for all T, but it holds over many correlation1200 @ 1989 The American Physical Society


228 J. Opt. Soc. Am. BNol. 6, No. 21February 1989 Satyanarayana et al.Ringing revivals in the interaction <strong>of</strong> a two-level atom withsqueezed lightM. Venkata Satyanarayana, P. Rice, <strong>Reeta</strong> <strong>Vyas</strong>, and H. J. CarmichaelDepartment <strong>of</strong> <strong>Physics</strong>, <strong>University</strong> <strong>of</strong> <strong>Arkansas</strong>, Fayetteville, <strong>Arkansas</strong> 72701Received May 12,1988: accepted October 4.1988The Jaynes-Cummings interaction <strong>of</strong> a two-level atom with the radiation field is studied when the radiation isinitially in a strongly squeezed coherent state. The dynamic response <strong>of</strong> the atomic inversion shows echoes aftereach revival when the squeezed coherent state exhibits an oscillatory photon-counting distribution due to the phasespaceinterference effect. The sensitivity <strong>of</strong> the dynamic behavior to approximations used in computing the atomicinversion is discussed. Comparison is made with the intensity-dependent interaction model <strong>of</strong> Buck and Sukumar[Phys. Lett. BlA, 132 (1981)lf this model does not exhibit echoes. he mean, variance, and entropy for the photonnumberdistribution are calculated and found to show behavior similar to that <strong>of</strong> the atomic inversion.1. INTRODUCTION sum is not possible even for. an initial coherent state, althoughapproximate expressions that reproduce the generalcharacter <strong>of</strong> the revivals have been obtained.3s4 We comparethe exact numerical evaluation <strong>of</strong> Eq. (1) with variousapproximations when P(n) corresponds to a squeezed coherentstate.The plan <strong>of</strong> the paper is as follows: In Section 2 we brieflyreview the oscillatory nature <strong>of</strong> P(n) for squeezed coherentstates. We study the corresponding dynamical response <strong>of</strong>the atomic inversion in Section 3. The exact numericalevaluation <strong>of</strong> w(t) is presented, demonstrating the occurence<strong>of</strong> ringing revivals, and the origin <strong>of</strong> the ringing behavior isdiscussed. In Section 4 we obtain a closed analytical expressionfor w(t) in the harmonic approximation, in which thesquare root in the argument <strong>of</strong> the cosine in Eq. (1) is expandedto first order. The ringing <strong>of</strong> the revivals is lost inthis approximation. Expansion <strong>of</strong> the square root to secondorder recovers the ringing behavior. The summation formulathat yields an analytical result in the harmonic approximationmay also be used to derive an exact integralrepresentation for w(t) when the radiation field is initially ina squeezed coherent state. This integral representation isgiven in Appendix A. We study the photon statistics andentropy for the field in Section 5. Our results are summarizedin Section 6.The Jaynes-Cummings model <strong>of</strong> optical resonance1 describingthe interaction <strong>of</strong> a single two-level atom with a singlemode <strong>of</strong> the radiation field has predicted a number <strong>of</strong> interestingfeatures in the dynamical behavior <strong>of</strong> the atomic inversion.24Much attention has focused on the collapse andrevival <strong>of</strong> Rabi oscillations because this effect provides evidencefor the granularity <strong>of</strong> the radiation field.3.4 Experimentalrealizations <strong>of</strong> the Jaynes-Cummings model havebeen obtained by using Rydberg atoms interacting with theradiation field in a high-Q microwave cavityS6 Recentlyobservations on the collapse and revival <strong>of</strong> Rabi oscillationswere reported.6 In this paper a new feature in the dynamicalbehavior <strong>of</strong> the atomic inversion is studied, with theradiation field prepared in a strongly squeezed coherentstate whose photon-counting distribution is oscillatory.7Under these conditions, the collapse following each revivalhas an oscillatory envelope (echoes), a phenomenon that wecall ringing revivals.Milburn has studied the interaction <strong>of</strong> a two-level atomand a single mode <strong>of</strong> the radiation field with the field preparedin a squeezed coherent state.8 He showed that thecollapse time depends on the direction <strong>of</strong> the squeezing andfound that for certain squeezed states the response <strong>of</strong> theatom is similar to that for chaotic radiation. However, Milburnrestricted his study to states for which the coherentcontribution to the photon-number variance is dominant.The new behavior described in this paper is obtained withsqueezed coherent states for which the squeezed contributionto the photon-number variance is dominant.Our results for the atomic inversion are based on thenumerical evaluation <strong>of</strong> the series1Squeezed coherent states are now quite familiar in quantumoptics; we simply state their definitiong and refer the readerto two recent collections <strong>of</strong> papers1° for further details and areview <strong>of</strong> current activity regarding these states.The squeezed coherent states for a single-mode radiationfield can be obtained from the vacuum (0) aswhere P(n) is the photon-number distribution for the initialstate <strong>of</strong> the radiation field and X is the coupling constant forthe atom-field interaction. The atom is assumed to be in itsground state initially. An exact analytical evaluation <strong>of</strong> thiswhere S(t) is the squeezing operatorS(S) = exp[%(t*a2 - Sat2)1 (3)and D(a) is the displacement operatorQ 1989 Optical Society <strong>of</strong> America


Photothermal lensing spectroscopy in aflowing medium: theory<strong>Reeta</strong> <strong>Vyas</strong> and R. GuptaA complete and general theoretical deecription <strong>of</strong> dual-beam photothermal lensing spectroscopy is given.The results are valid for the moat general conditions, that is, for flowing as well as stationary media, and for cwas well as pulsed excitation. For pulsed excitation, the results are valid for arbitrary pulse length. The cwresults apply to both modulated as wellas unmodulated excitation. Both transverse and collinear geometriesare considered.I. IntroductionIn this paper we present the theory <strong>of</strong> dual-beamphotothermal lensing spectroscopy (PTLS) in a fluidmedium valid for the most general conditions, that is,for flowing as well as stationary media and for cw aswell as pulsed excitation. For pulsed excitation, thepulse length is arbitrary, and for cw excitation both themodulated and the unmodulated sources are considered.Both the transverse and the collinear geometriesare considered. A unified treatment <strong>of</strong> all cases ispresented. This is the first comprehensive treatment<strong>of</strong> this important subject.The basic idea underlying dual-beam PTLS isshown in Fig. 1. A laser beam (pump beam) propagatesthrough a medium, and it is tuned to one <strong>of</strong> theabsorption frequencies <strong>of</strong> the medium. The mediumabsorbs some <strong>of</strong> the optical energy from the laserbeam. If the collision rate in the medium is sufficientlyhigh compared to the radiative rates, most <strong>of</strong> theenergy appears in the translational-rotational modes<strong>of</strong> the medium within a short period <strong>of</strong> time. In otherwords, the laser-irradiated region gets slightly heated.The refractive index <strong>of</strong> the medium is thus modified.The refractive-index change can be monitored in severaldifferent ways.' In this paper, we are concernedwith a technique that relies on the lensing effect <strong>of</strong> themedium to monitor the .refractive-index change. Aweak probe beam passes through the pump-irradiatedThe authors are with <strong>University</strong> <strong>of</strong> <strong>Arkansas</strong>, <strong>Physics</strong> Department,Fayetteville, <strong>Arkansas</strong> 72701.Received 6 June 1988.0003-6935/88/224701-11$02.00/0.O 1988 Optical Society <strong>of</strong> America.region, as shown in Fig. 1. Due to the curvature <strong>of</strong> therefractive index, the probe beam diverges, which canbe detected as a change in the intensity <strong>of</strong> the probebeam passing through a pinhole. In other words, underthe influence <strong>of</strong> the pump beam, the medium adslike a diverging lens. In certain circumstances, themedium acts like a converging lens also. If a pulsedpump laser is used, a transient lens is formed; theprobe beam changes shape shortly after the pumpbeam is fired and returns to its original shape on thetime scale <strong>of</strong> the diffusion <strong>of</strong> heat from a probe region.If a cw laser is used, it is generally convenient to amplitudemodulate its intensity, and the PTLS signal consists<strong>of</strong> oscillations in the intensity <strong>of</strong> the probe beampassing through the pinhole. The PTLS takes onmore interesting dimensions if a flowing medium isused. The PTLS technique has been discussed extensivelyin the literature for trace detection <strong>of</strong> chemicals?and the use <strong>of</strong> PTLS in a flowing medium hasrecently been demonstrated for flow velocity measurement~.~Although the thermal lensing effect may alsobe observed by monitoring the pump beam itself, inthis paper we only consider the dual-beam techniquein which the thermal lens is created by the pump beamand monitored by the probe beam.The photothermal lensing effect was observed accidentallyby Gordon et a1.4 in 1964 when they placed acell filled with a liquid sample inside a He-Ne lasercavity. These authors correctly identified the effectand gave a theoretical description <strong>of</strong> it. In 1973, Huand Whinnery5 gave a detailed theoretical description<strong>of</strong> the effect for an extracavity sample and determinedthat the maximum signal occurred when the samplecell was placed one confocal distance away from thewaist <strong>of</strong> the beam. They also demonstrated the usefulness<strong>of</strong> this technique for measurements <strong>of</strong> smallabsorptivities. A cw laser was used, and a shutter wasemployed to effect the change necessary to observe the15 November 1988 / Vol. 27. No. 22 / APPLIED OPTICS 4701


Continuous wave photothermal deflection spectroscopy ina flowing medium<strong>Reeta</strong> <strong>Vyas</strong>, 6. Monson, Y-X. Nie, and R. GuptaA complete and rigorous theoretical treatment <strong>of</strong> the continuous wave photothermal deflection spectroscopyin a flowing medium is given. The theoretical resulta have been verified experimentally.I. IntroductionThere is presently extensive interest in the technique<strong>of</strong> photothermal deflection spectroscopy(PTDS),1-3 and this technique has found a wide range<strong>of</strong> applications. In this paper we present a comprehensiveand rigorous theoretical treatment <strong>of</strong> the cwPTDS technique for the most general conditions, i.e.,for a flowing medium. The results for a stationarymedium appear as a special case. Very interestingresults have been found, and these results will extendthe range <strong>of</strong> applications <strong>of</strong> cw PTDS in fluids. Results<strong>of</strong> an experiment used toverify the theory are alsopresented.The principle <strong>of</strong> the PTDS technique is as follows:A dye laser beam (pump beam) passes through themedium, and it is tuned to an absorption line <strong>of</strong> themolecules or atoms <strong>of</strong> interest. The molecules absorbthe laser energy, and in the presence <strong>of</strong> quenchingcollisions some <strong>of</strong> this energy appears in the rotational-translationalmodes (heating) <strong>of</strong> the molecules <strong>of</strong>the medium. If the quenching rates are high comparedwith the radiative rates, almost all the absorbedenergy appears in the heating <strong>of</strong> the laser irradiatedregion. The temperature change <strong>of</strong> the laser irradiatedregion results in a change <strong>of</strong> the refractive index <strong>of</strong>that region. This change in the refractive index can beprobed by a second, and weaker, laser (probe beam).In general, the refractive index is nonuniform, and therefractive-index gradient deflects the probe beam.The deflection <strong>of</strong> the probe beam may be detectedThe authors are with <strong>University</strong> <strong>of</strong> <strong>Arkansas</strong>, <strong>Physics</strong> Department,Fayetteville, <strong>Arkansas</strong> 72701.Received 22 December 1987.0003-6935/88/183914-07$02.00/0.63 1988 Optical Society <strong>of</strong> America.conveniently using a position sensitive optical detector.In the case <strong>of</strong> a cw pump beam, it is generallyconvenient to intensity modulate the pump beam atsome frequency f. In this way, the deflection is alsomodulated at frequency f, which can then be detectedconveniently using a phase-sensitive detector.The only previous theoretical study <strong>of</strong> the cw PTDS,to our knowledge, is that <strong>of</strong> Jackson et aL4 Theseauthors, however, consider only a stationary medium.Our study is more general (i.e., a flowing medium), andJackson et ale's results appear as a special case (flowvelocity = 0) in our solutions. Moreover, this treatment<strong>of</strong> the cw PTDS unifies the theory <strong>of</strong> cw PTDSwith that <strong>of</strong> pulsed PTDS in a flowing medium, publishedearlier by Rose et aL5The theory <strong>of</strong> cw PTDS is given in Sec. 11. Thetheoretical results are discussed in Sec. 111. The apparatusis described in Sec. IV, and the experimentalresults are presented in Sec. V.II. TheoryFigure 1 shows the basic geometry <strong>of</strong> the pump andprobe beams. Two cases are considered: TransversePTDS, in which the probe beam is perpendicular to thepump beam, and the collinear PTDS, in which thepump and probe beams are parallel. We assume thepump beam to be propagating along the z axis. Theorigin <strong>of</strong> the coordinate system lies on the axis <strong>of</strong> thepump beam. For the transverse PTDS, the probebeam propagates along they axis, and for the collinearPTDS, the probe beam propagates in the z direction.In both cases, the medium flows with velocity v, in thex direction. The pump and probe beams do not necessarilyintersect, and they may be separated by a variabledistance x in the x direction. For the collinearcase, they may also be separated by a distance y in theydirection (not shown in Fig. 1). The general case,where the pump and probe beams make an arbitraryangle 8 will not be considered here. This case has beentreated by Rose et al.5 previously in connection with3914 APPLIED OPTICS I Vol. 27, No. 18 I 15 September 1988


PHYSICAL REVIEW A VOLUME 38, NUMBER 5 SEPTEMBER 1, 1988Waiting-time distributions in the photodetection <strong>of</strong> squeezed light<strong>Reeta</strong> <strong>Vyas</strong> and Surendra SinghDepartment <strong>of</strong> <strong>Physics</strong>, <strong>University</strong> <strong>of</strong> <strong>Arkansas</strong>, Fayetteville, <strong>Arkansas</strong> 72701(Received 24 March 1988)Distribution <strong>of</strong> waiting-time intervals between the arrivals <strong>of</strong> successive photons on a photocathodeilluminated by a beam <strong>of</strong> light is discussed. Analytic expressions for the conditional andunconditional distributions for squeezed light are derived in the high degeneracy limit. Results forbinomial and thermocoherent states are also given. Curves are presented to illustrate the behavior.I. INTRODUCTIONSqueezed states <strong>of</strong> light have been observed in a number<strong>of</strong> experiments.'-4 Similar to the states exhibitingphoton antib~nchin~,~ squeezed states6 are nonclassicalstates. For these states variance in one <strong>of</strong> the quadraturecomponents <strong>of</strong> the electric field is smaller than the correspondingvariance for a coherent state. The nonclassicalnature <strong>of</strong> these states is also reflected in the fact that thecorresponding phase space density in the coherent statediagonal representation7.' does not exist as an ordinaryprobability density. Photon number distributions forsqueezed states have been discussed by a number <strong>of</strong> authors9-l4and they reveal several interesting features <strong>of</strong>these states. Depending on the parameters characterizinga squeezed state, such a state may exhibit sub-Poissonian or super-Poissonian photon statistics.I4 Since,for short times at least, sub-~oiisonian (super-Poissonian)photon statistics reflect antibunching (bunching),I5squeezed states are capable <strong>of</strong> exhibiting antibunching orbunching under suitable circumstances. The properties<strong>of</strong> photon antibunching or bunching are best visualized interms <strong>of</strong> the theory <strong>of</strong> photoelectric detection.16-" Considera photoelectric detector illuminated by a beam <strong>of</strong>light. For an ideal photodetector <strong>of</strong> unit detectionefficiency and zero dead time, the photoelectric pulses appearingat the output <strong>of</strong> the photodetector are in one-toonecorrespondence with the arrival <strong>of</strong> photons on thephotocathode. In terms <strong>of</strong> this sequence <strong>of</strong> photoelectricpulses, photon antibunching implies that the detection <strong>of</strong>a photon at a certain time t renders the detection <strong>of</strong>another photon immediately following the first less probable.The opposite is implied by photon bunching. Thismeans that for an antibunched beam <strong>of</strong> light, successivephotoelectric pulses will be separated, on the average, bylarge time intervals. We may consider this to be areflection <strong>of</strong> the tendency <strong>of</strong> the photons in the lightbeam to be separated in time. This physically appealingpicture <strong>of</strong> the behavior <strong>of</strong> photons in a light beamemerges from photodetection theory. Photons themselvesdo not lend directly to such an interpretation. Thistime-dependent behavior <strong>of</strong> photons will be reflected inthe distribution <strong>of</strong> the time interval between successiveph~todetections.''~~~ It is the purpose <strong>of</strong> this investigationto derive the time interval distribution for photons ina squeezed state.In Sec. I1 we present an outline <strong>of</strong> the photodetectiontheory leading to a general expression for the time intervaldistribution. This expression is used to discuss the behavior<strong>of</strong> photons in a squeezed beam in Sec. 111. Wealso discuss the behavior <strong>of</strong> waiting time distributions forbinomial and thermocoherent states <strong>of</strong> the field in Sec.IV.11. TIME INTERVAL DISTRIBUTION FUNCTIONIn order to discuss the time interval distribution for alight beam we need the threefold joint probability <strong>of</strong> photodetection,for the time interval T between successivephotodetections is defined in terms <strong>of</strong> three events: aphotodetection event at some time t,, no photodetectionevents in the interval [t , , t2( = t, + T )], and one photodetectionat time t2. First let us recall that the probability<strong>of</strong> detecting n photoelectric events in [t ,,t2] isI7whereand ?(t) is the photon flux operator (number <strong>of</strong> photonsper unit time) and : . . : denotes the time-ordered normalproduct <strong>of</strong> operators. The detection efficiency 77 dependson the characteristics <strong>of</strong> the detector. The angularbrackets denote averaging with respect to the state <strong>of</strong> thefield. From Eqs. (1) and (2) the probability <strong>of</strong> detectingone photon in the interval [t -At, t] is found to bewhere At is some small time interval. We also find fromEq. (1) that the probability that no photodetection occursin the interval [t t2] isEquations (3) and (4) will be used later in the paper. Thesingle-fold photoelectric counting formula (1) is wellknown.16-" The threefold photoelectric counting formula(to


Pulsed photothermal deflection spectroscopy in a flowingmedium: a quantitative investigationA. Rose, <strong>Reeta</strong> <strong>Vyas</strong>, and R. GuptaA comprehensive investigation <strong>of</strong> pulsed photothermal deflection spectroscopy in a flowing medium has beencarried out. A rigorous solution <strong>of</strong> the appropriate diffusion equation has been obtained, and experimentshave been conducted to verify the theoretical predictions. Absolute measurements <strong>of</strong> the photothermaldeflection were made and no adjustable parameters were used in the theory. Very good agreement betweenthe theory and the experiment was obtained.I. IntroductionRecently there has been extensive interest in thetechnique <strong>of</strong> photothermal deflection spectroscopy(PTDS). Since the initial work <strong>of</strong> Davis,' and <strong>of</strong> Boccara,Fournier, Amer and co-~orkers,2-~ extensive newapplications have been developed by Tam and coworker~~-~and by others.1° Gupta and co-workershave demonstrated the usefulness <strong>of</strong> this technique forcombustion diagnostic^.^^-^^ Motivated by this application,here we present results <strong>of</strong> a comprehensive andquantitative investigation <strong>of</strong> pulsed PTDS in a flowingmedium.The principle <strong>of</strong> the PTDS technique is quite simple:A dye laser beam (pump beam) passes throughthe medium <strong>of</strong> interest and is tuned to one <strong>of</strong> theabsorption lines <strong>of</strong> the molecules that are to be detected.The molecules absorb optical energy from thelaser beam and, if the pressure is sufficiently high (i.e.,if the quenching rates are sufficiently fast compared tothe radiative rates), most <strong>of</strong> the energy quickly appearsin the rotational-translational modes <strong>of</strong> the medium.The dye laser irradiated region thus gets slightly heated,leading to changes in the refractive index <strong>of</strong> themedium in that region. If the density <strong>of</strong> the absorbingmolecules is uniform over the width <strong>of</strong> the dye laserbeam, the refractive index acquires the same spatialpr<strong>of</strong>ile as the dye laser beam (e.g., a Gaussian pr<strong>of</strong>ile ifthe dye laser is working in the TEMoo mode). Now ifanother laser beam (called the probe beam) overlapsthe pump beam, it is deflected due to the changes inThe authors are with <strong>University</strong> <strong>of</strong> <strong>Arkansas</strong>, <strong>Physics</strong> Department,Fayetteville, ~rkansas 72701.Received 25 July 1986.0003-6935/86/244626-18$02.00/0.O 1986 Optical Society <strong>of</strong> America.the refractive index created by the absorption <strong>of</strong> thepump beam. This deflection is easily measured by aposition-sensitive optical detector. If a pulsed dyelaser is used, a transient deflection <strong>of</strong> the probe beam isobtained. The deflection <strong>of</strong> the probe beam is proportionalto the concentration <strong>of</strong> the absorbing molecules.Therefore the technique can be used to measure major-ity and minority species con~entrations.~l-~~ If themedium is flowing, the heat pulse produced by theabsorption <strong>of</strong> the dye laser travels downstream withthe medium. The heat pulse is, <strong>of</strong> course, accompaniedby changes in the refractive index and can bemeasured by the deflection <strong>of</strong> a suitably placed probebeam. The flow velocity <strong>of</strong> the medium can be measuredfrom a measurement <strong>of</strong> the transit time <strong>of</strong> theheat pulse between two probe beams downstream fromthe pump beam.7J5J8J9 The heat pulse broadens dueto thermal diffusion as it travels downstream. Thethermal diffusion coefficient <strong>of</strong> the medium can bemeasured from the broadening <strong>of</strong> the signal which, inturn, yields the local temperature <strong>of</strong> the medium.l4*6Therefore photothermal deflection spectroscopy is avaluable optical diagnostic technique for varied applications(combustion diagnostics is just one <strong>of</strong> the applications).Quantitative measurements <strong>of</strong> speciesconcentrations, temperature, and velocity are moredifficult, however, than one might assume from theabove discussion. The PTDS signal in a flowing mediumdepends in a complicated manner on all threeparameters (concentration, velocity, and temperature).For example, the amplitude <strong>of</strong> the signal dependsnot only on the concentration <strong>of</strong> the absorbingspecies but also on the temperature and the flow velocity.The width <strong>of</strong> the signal depends on both thetemperature and the flow velocity. The measurement<strong>of</strong> the transit time is affected by the broadening due tothermal diffusion. Therefore, a good theoretical andexperimental understanding <strong>of</strong> the size and shape <strong>of</strong>4626 APPLIED OPTICS 1 Vol. 25, No. 24 1 15 December 1986


PHYSICAL REVIEW A VOLUME 33, NUMBER 1 JANUARY 1986Laser theory without the rotating-wave approximation<strong>Reeta</strong> <strong>Vyas</strong> and Surendra SinghDepartment <strong>of</strong> <strong>Physics</strong>, <strong>University</strong> <strong>of</strong> <strong>Arkansas</strong>, Fayetteville, <strong>Arkansas</strong> 72701(Received 17 June 1985)The effect <strong>of</strong> counter-rotating terms in the interaction Hamiltonian on the photon statistics is investigatedfor a single-mode laser. The treatment is based on the Scully-Lamb model. A perturbationexpansion is used to derive an equation <strong>of</strong> motion for the field-density matrix. The counterrotatingterms are found to lead to both transient and secular effects. With some reasonable approximationsthe steady-state photon-number distribution is derived and it is shown that the counterrotatingterms tend to raise the threshold <strong>of</strong> laser action and broaden the photon-number distribution.The relative size <strong>of</strong> these effects is found to be very small and depends on the square <strong>of</strong> the ratio<strong>of</strong> the atomic linewidth and the transition frequency.I. INTRODUCTIONIn quantum optics, NMR, quantum electronics, andother branches <strong>of</strong> resonance phenomena it is customary tomake the rotating-wave approximation (RWA).' Underthis approximation certain terms which oscillate at twicethe resonance frequency are dropped from the interactionHamiltonian. The resulting Hamiltonian describes thedynamics <strong>of</strong> the system quite adequately. However, therapidly oscillating terms (also called the counter-rotatingor energy-nonconserving terms) can give rise to physicaleffects even if small. For example, in the Rabi problemthey give rise to the Bloch-Siegert shift <strong>of</strong> the resonancefrequency and small high-frequency amplitude modulation<strong>of</strong> the Rabi oscillation^.^ In laser theory the role <strong>of</strong>these counter-rotating terms has not been investigated. Inthis paper we wish to consider the effect <strong>of</strong> such terms inlaser theory. We find that the counter-rotating terms giverise to both secular and time-dependent effects. In particular,their effect is to raise the threshold <strong>of</strong> laser oscillationand broaden the photon-number distribution. However,the magnitude <strong>of</strong> these corrections is very small asexpected. In our investigations we adopt the model <strong>of</strong>Scully and Lamb3 for a single-mode laser. Unlike them,however, we use a perturbative approach.3*4 In Sec. I1 wedescribe the details <strong>of</strong> the model and derive a masterequation for the density matrix <strong>of</strong> the laser field for atraveling-wave mode. In Sec. 111, this master equation issolved in the steady state and changes produced by thepresence <strong>of</strong> counter-rotating terms are discussed. We alsoconsider the standing-wave case and discuss the differencesthat arise in the equations <strong>of</strong> motion.11. EQUATIONS OF MOTIONOF THE DENSITY MATRIXWe consider a single-mode electromagnetic field <strong>of</strong> frequencyo interacting with a group <strong>of</strong> N identical twolevelatoms. The energy separation between the upperatomic level ( a ) and the lower atomic level / b ) is fbo,where oo is equal or close to the field mode frequency.The atoms may decay nonradiatively out <strong>of</strong> the two levels/ a ) and ( b ) to various other levels at a rate y. FollowingScully and Lamb3 we suppose that the laser gain isprovided by atoms introduced uniformly at an averagerate NP throughout the cavity. These excited atoms interactwith the field inside the cavity and make their contributionsto the field one at a time. The total rate <strong>of</strong>change <strong>of</strong> the optical field may be calculated as a coarsegrainedderivative by multiplying the change produced byone atom by the rate NP at which atoms are introducedin the excited state. Similarly. laser loss is simulated byconsidering another set <strong>of</strong> fictitious two-level atoms withbroad levels introduced into the cavity in their lower stateat a certain uniform rate.3 These atoms absorb radiationand model the loss suffered by the laser field which is ultimatelydue to photons escaping from the cavity eitherbecause <strong>of</strong> the finite transmittivity <strong>of</strong> the mirrors or because<strong>of</strong> scattering from various elements inside the cavity.The Hamiltonian for an interacting atom and asingle-mode electromagnetic field is given bywhere s,,R^,,R^, are the three Pauli spin-f operators, pis the transition dipole moment between the levels / a )and / b) which we take to be real, and r is the position <strong>of</strong>the atom. The single-mode electric field operator @(r,t) isgiven by (in mks units)where E is a unit polarization vector which we take to bereal corresponding to linear polarization, a^ and ii ' are theannihilation and creation operators for the field, u (r) isthe cavity mode function, and V is the cavity volume(quantization volume). The third term in Eq. (1)represents the interaction Hamiltonian d. This term inthe interaction picture can be written aswhere01986 The American Physical Society


-==--d WOSRb d ~ i ,J <strong>of</strong> the Optical Society <strong>of</strong> Amer ,ica A - Optics Image Science and Vision, v 2, Issue 13, p25-p25, 1985me llw Intensities <strong>of</strong> the two modes. thadiierenca. and the autocarelatlon and crossarrelationhulctlons were calculated. The kind <strong>of</strong>behavla exhibited by the laser Is found to dependoitlcally on the phases Of the backscatlcmd fiilds.Sams values lead to posltive crou conelatlms InUm rqim where the laser ordinarily exhlblts zema negatlve correlations, and othen to nsgathremRNOOn 1-modes are not locked togalhr they undergo randcmjumps In Integer multiples <strong>of</strong> 2r. the resub<strong>of</strong> the urlcuktions are mmpared with expalmen-I results from whlm some conclusions about themopmaa, / backscattering can be &awn. (12 mln)IlUW4 Paper rllhbraunNWS likasuremenl <strong>of</strong> flnt-pasasgttlm dleblkrllarrfor a Os~nchsd lawMARVIN R. YOUNG and SURENORA SINW. U.<strong>Arkansas</strong>. <strong>Physics</strong> Departmnt. Fayetbvllls, AR72701.I The dlstributlons <strong>of</strong> first-passage tlm for Ihoi bulld up 01 hser radlatlon hom zero IntemHy toSome predetermined value are measured for a CTJwiidmd her. Experiments were performed on ai sing~ernade He-Ne laser operating nearthnshold.aswltching was performed by applying a voItsge,Pulse to an intracavlly awustwptlc modulator.Rwu)ts are compared to thaoretld calculatlms.(12 mln)%€fA VYAS and !SURENCRA SINGH. U. <strong>Arkansas</strong>.<strong>Physics</strong> Depamnent. Fayeneville. AR 72701.W. R. M. G PAYNE. and W. R. GARRm,Oak Ridge National Laboratory. Heam and SafelyResearch Dlvkion. Oak Ridge. TN 37831.A preWoadenlng study <strong>of</strong> t h r mand me-phG+gn resonances was perf<strong>of</strong>meti mingcounterpropagatlng, linearly polarized laserbeam. Self-bmadening. & well .as Wm collislO+Induced etfects due to Uw presenur Ol a sscondmble gas. was investigated fa tJm 6s. 6s'. Sdand7s<strong>of</strong>xenonandn Sd<strong>of</strong> -ton. A muniphotonionlzatlon scheme In a calibratd proportionalcounter is the basls <strong>of</strong> a new tectmlque by whichmllislonal processes can be lnvestlgated tor h epresscae regimes whlch have previously been inaaxssibleto hvestlgation, psrtlcularly for selfbroaddlwhere the mean (ree path <strong>of</strong> the @Ommb<strong>of</strong>lheOrkn<strong>of</strong>10'3cmfapeosures~kwas 1 Ton. Using ihis tschnlqus. we have rsmdedwlllsiorAnduced wldmP and shifts <strong>of</strong> the re-~laser~ando(herbnvldres0lQuenNnelectronics certain t m oscilhtlng at twh tharesonance freq~~cyarsdopped from me lntaac- wnce for pr~ssues up b 1000 Tm. In the hmfmHamlltonh. he ettm <strong>of</strong> mese rmntenotatdreda <strong>of</strong> tur range. asymnebic. sell-broadsniwIO tanu on the operauon <strong>of</strong> a single-made laser ISllne da%c8s were o b d wlh full wldms andWnsidereduslng therodel<strong>of</strong> ScuIfyandLamb. &isshim vrhlch were greater than five times Iho laserfaad that these t m lead to a shin <strong>of</strong> the laser bandwklth [0.002-0.005 M1). (12mm)thesholdandabmadening<strong>of</strong>~photonnunb~6sbbutlm.(12mSn)A. S. MARATHAY. 8. A. CAPRON. and M. SAR-GRJr Ifl. U. Arlmrw OpUal S c i i Cents. T*~n. AZ 85721.Refweme 1 &rive the twO+hOton15 October 1985suption co<strong>of</strong>liclent for a hanogeneourly bmadmedimsubjecbd toJEmRSONan arbmarib Intsn?w,sabra@ wave. We have gsneralked this mWVto allow fw bppk broedeninq with a ccuntnlngprobs and r a m waves. We have cardeda~ Integral3 over s Lorentrian Doppler dlsbibutlcnadytlcally and-have specialized m a number <strong>of</strong>mX1 Absoldw two-phalon abwrptlon and Ilmlts. such as emar.m Doppler W i n g . In%hr.epMon knlratlon crcm sedans la atdc this limit, we llnd the simple wok, ebsapti~norlmcoetRclent a, 2ndyDIh - Iw.Tt). wlwre ao Isthe rnsatuatedabsorptirn coeMdentbWOLAS J. BAMFORD. LEONARD E. &lSIHSKI, r,Istheseh.amr~.ykmh*0photonancl wUlAfw4 K. BISG-EL. SRI I-. Chm coherence decay rate comtanS us k the St* 3N'ftleal mi- Laboratay. Menlo PYk. CA 04025. parmeter. T, ia the popdatran difference decayBefas r photon dstectian Ichsma?i for at- tkne, and a, L I/[? + r(o + WA - PZ- ~111.~ e r r d m o l ~ c a n b s p w e d m a Fa ~ van- Stark shifts (0. 0). thb is the~ooli~aoss~la~armnl~bwi- rtandsdtwophotarFmbs~-~tiom must-be know. We hllua rma~crad bu,forhw,mecom~lexLaMltLianf)tbb~e~mcm~sectiaata~40a".~~wrlam speck b, Reme dlam*. T bnlgus <strong>of</strong> tw-~QI OXCI~~~ thaesoaoolaedtDmefwre~heams~tame2p~h~;,,$q1985 Annual Meeting OpUcal Society <strong>of</strong> America P25 ! 'il3~?~,,.0 trarultlon at 226 nm. Excltatlon was 1. M. Sargent Ill. S. Ovadii. and M c 111 Phys.tXmltored by Ob~(*~lng fluorescence at 845 nrn Rev A (1985) to be published.from the 3p SPzt,o + 3s3S, transition. The fluerescence collection system was callbraled absnlutely us[yl spa- scanelm hm TUX4 Saluratlon MKb In mukiphalon lonlzamalecularhydrogen. The spatial and toqua1 tbnpmfiles <strong>of</strong> the exclting dye laser pulses were carefullymeaswedln agently focused excltatlon peom JOHN A YEAZEU. MlCW S. MALCUIT. CAReby(peak IntensHy -10 MWl& allwlng the LOS R. STROLD. Js., end ROBERT W. BOYD, U.t w w cross section to be detmlntd. The Rochastw. Institute <strong>of</strong> OpUcr. Rochester, NYaos section fa absorption <strong>of</strong> a third, identical14627.photon (2 + 1 photoionization) was Umn beter- We have sMled two-photon rescfwl Wv*mined by rneasurlng the absolute nun- <strong>of</strong> Ions photon lonlzatlon <strong>of</strong> atomic rodium vapor. A meproduced per laser pulse. The meanred cmss laser tuned to elther the 46 or 5s twmtw alseclloluarea= 2.5 X 10-~cm4/Wandopl= 6.1 bwed transnlon was used to excite the atun ud aX 10-lS d. Wng a square pulse In spaw and second m e Intense her (A = 7000 A) yaa usedtime, we calculate that - 1 % IoniaUon Is poorlble tor lonizatlon. Tha Ion yleld was measwed 8s aforal~s~n~at226nm<strong>of</strong> 1 J/c~Inal(kri hnction <strong>of</strong> remtlaser lntwity aml detwlng.long pulsa.(12 mh) the nonrsMMnt Laser IntsmiRy, and the temporaloverlap <strong>of</strong> the exciting ard lonlzlng 1- pldses.We have found' Umt Bt high laser Intenaltlm theNX2 IWrdmt) prrrwrre bmadmhg <strong>of</strong> throe- photolmlzdtiar yield decreases as he nwyescphotonresonances fn gascrWR kzw mlemity Is increased. When tuwd tome 5s level a value <strong>of</strong> --0.6 was obtained !u thsexponentla1 index avan with the resonant lasatuned to rnaxlrnize the ion ykld. A IheoretlCdlmodel lncludlng the ettw <strong>of</strong> Stark shln QI IMgumdstale, power broadening 01 me n~-~~-"elelevel due to 6atwation <strong>of</strong> the WWhotm WaWlthand by rapid trw~tiOnrr to the Icnizatbl cerltlwasused to predict the experlmsnbl resub.Good agreement between lf=W and emirequiredthe use 01 a photolonii WOW Kcamfor me 5s bvel which was -20 times Weator rtranme acwptad value. Redts for photo~onbtbban the 4dlevel are also pressnted. (12 mEn)1 L AIM, R. W. Boyd. J KrasWl. M. S. Malcun.and c R. s~aud. Jr.. mys. Rev Len. 51. 309(1985).TUX5 Canp.tltlar b.trrwn cohnn( nnd Lnc*Mlmtmnllmu~poccrt..DAMEL J GAUMmL M I W S MA1 CUP. andROBERT W. BOYD. U Rocbsler, 1-1- <strong>of</strong> Dptlu. Rorhsstn. NY 11827Canp3tMn ha. besn obrsrred beeen qlCfledemtulon (ASQ and fw-wavemlxhg (FWH) Revlously, -Ion betweenan k&~.m$ and coharent proceu was o~xmedm ths rum<strong>of</strong> suppasion <strong>of</strong> muniphoton bnizatkmby m~ m i c ganaatta, I,# xr exoerCmcrn.un~npnbsef~bnedprecac)yto~?- jd- allowed transmon h amlc sodC~~pagnjUm&uia,wa,WredInb~the fprward backward d~rectiom. In CheWcmMom. arther FWM a ASE can cUX Igm-,ng the p~~~ibillty <strong>of</strong> canpetltlon betyscn~opmcs~~,checakulaWpaln fuASE~smuchman mat (a FWM Ho*erer. we have ch--d expmimentally that FWM and not ASE oc--. TO ~~nnure mat H IS the pewncs d emFW w- AS€, we have exmsdthe atanlc vapor lmtsad With WWKW'Wpvnp waves <strong>of</strong> muem tregoency t.ra case.FWM. be~ng a pr-9 ~~ cG-ands~ngASEnobserved Themtlw<strong>of</strong>*ccmpet~lan is erpbned by 60hnnQfoupledh&xwell and demy mablx equam 'mwsy~yshmr nisfandmthesuppr-b&JO to me desrmctive nderterence Wweenpetmws d e x ~ ~ CA ~ (he o n 3a WAwtuchresult.l~n~Lq<strong>of</strong>me~'="~~qand me. (12 m)

Hooray! Your file is uploaded and ready to be published.

Saved successfully!

Ooh no, something went wrong!