Download - Anomalous mode pulling, instabilities, and chaos in a single-mode, standing-wave 3.39-μm He-Ne laser

Transcript
Page 1: Anomalous mode pulling, instabilities, and chaos in a single-mode, standing-wave 3.39-μm He-Ne laser

PHYSICAL REVIEW A VOLUME 29, NUMBER 3 MARCH 1984

Anomalous mode pulling, instabilities, and chaos in a single-mode, standing-wave3.39-)Mm He-Ne laser

R. S. Gioggia* and N. B.AbrahamDepartment ofPhysics, Bryn Mawr College, Bryn Mawr, Pennsylvania 19010

(Received 15 July 1983)

A high-gain, single-mode He-Ne 3.39-pm laser shows extreme mode pulling, as the laser operat-

ing frequency varies up to three times more slowly than the corresponding variation of the empty-

cavity frequency. When the laser is operated at low pressures, new features appear in the vicinity

of, and in conjunction with, the Lamb dip. Dispersive effects from the overlapping of the holes dis-

torts the more nearly linear mode pulling observed for high pressures near line center. When thedispersive effects are strong enough instabilities are observed including self-pulsing, quasiperiodic,and chaotic output. At sufficiently high gain, the dispersion is distorted to give a multivalued re-

gion for the laser operating frequency for a single cavity frequency as predicted by Bennett.

I. INTRODUCTION

High-gain lasers have been of particular interest becauseof the associated anomalous dispersion which enhancesmany otherwise small corrections to simple laser analyses.%"hen the laser medium is also inhomogeneouslybroadened, the dispersive effects of the holes burned in thegain profile can lead to single-mode self-pulsing, modesplitting and anomalous tuning of the laser frequency.Several of these effects are studied here using the 3.39-pmtransition in neon in a low-g, Fabry-Perot laser.

II. BACK(GROUND

The analysis of single-mode inhomogeneouslybroadened lasers was provided very early in the develop-ment of laser theory by the pioneering contributions ofBennett' and Lamb. Among the classic predictionsvFere the phenomenon of hole burning by selective satura-tion of narrow velocity classes of atoms within theDoppler-broadened transition and the appearance of theLamb dip, a reduced power output when the laser wastuned within a homogeneous linewidth of the resonanceline center. General treatments of the dispersive effects ofholes burned in the population inversion also lead to anunderstanding of many forms of mode competition(through "hole repulsion" ) and other effects variouslytermed "mode pushing" or "mode pulling. " The effectsof pressure broadening and interatomic collisions havealso been investigated in detail.

One of the often neglected effects predicted by BennettI'see Ref. 3, p. 128] was that as a single-mode laser wastuned through the resonant frequency of the transition (inthe vicinity of the Lamb dip), the mode would interactwith the dispersive effects of the "mirror-image" hole(burned because of bidirectional propagation of the beam)causing a distortion of the simple mode-pulling effectsthat would be observed, for example, in a single-mode ringlaser. Bennett was able to confirm the existence of thiseffect by finding that mode pulling could be distorted a

few tens of kilohertz out of a few tens of megahertz in aHe-Ne 1.15-pm laser. More extreme effects were predict-ed for higher-gain and lower-Q cavities, but until thepresent work, no studies have been made to observe thepredicted region of multiple frequencies.

Most mode-pulling studies have been made in 1aser sys-tems where the effects have been, at most, a few per-cent. However, higher-gain systems, such as the 3.51-pm line in xenon in particular, have shown very largeanomalous dispersion leading to mode-pulling factors offrom 2 to 50. ' These studies were sparked in part bythe prediction of special effects in such cases. ' More re-cently, renewed interest in the theory of inhomogeneouslybroadened lasers has also focused on dispersive effects inorder to understand the appearance of spontaneous pulsa-tions in dc-excited systems. '" Casperson' hasdeveloped a simple intuitive picture which helps todemonstrate how dispersive effects may lead to thedevelopment of sidebands to the laser operating frequencymanifesting themselves as pulsations in the intensity ofthe laser.

These oscillations have been observed in high-gain gaslaser systems, including standing-wave xenon lasers at3.51 pm, "' ' ' ' unidirectional ring lasers at 3.51pm, and in standing-wave helium-neon lasers at 3.39pm. The theories, however, have concentrated almostentirely on predictions for unidirectional ring lasers forsimplicity. Reported here is an experiment using aFabry-Perot helium-neon laser which displays both phe-nomena, self-pulsing instabilities and anomalous (andmultivalued) mode pulling.

III. EXPERIMENTAL SETUP

The experiments reported here used a laser tube andcavity specially designed for high-gain and single-modeoperation. A schematic is shown in Fig. 1. This laser hasbeen used previously in the study of single-mode instabili-ties in xenon lasers. "' ' ' The 909-MHz free spectral

1304 1984 The American Physical Society

Page 2: Anomalous mode pulling, instabilities, and chaos in a single-mode, standing-wave 3.39-μm He-Ne laser

ANOMfALOUS MODE PULLING, INSTABILITIES, AND CHAOS. . . 1305

C-

I

pzT

99/o

riJ-'6.$cI~.

10cm

f-mm-i. d.capillary

Q IVI

50/

FIG. 1. Schematic diagram of high-gain, single-mode laser.

A, anode; C, cold cathode; R, nonexcited return path; Q, quartzwindows at Brewster's angle; M, mirrors (20-cm focal length

and wedge); PZT, piezoelectric mirror drive.

range helps to ensure only single longitudinal mode opera-tion, while the 1-mm-diam capillary tube enhances thepopulation inversion (presumably by improving depletionof the lower level through wall collisions). The relativelylow-Q cavity (using one 50% refiecting mirror) is neces-sary for observations of both the single-mode instabilitiesand the unusually high degree of mode pulling. Consider-ing mirror and diffractive losses as well as absorption inthe quartz windows, we calculate a cavity lifetime of 1

nsec, or equivalently, a decay rate for the electric field of5)&10 sec

The laser was filled with a mixture of single-isotopeneon gas (99.9% Ne) and research-grade helium at pres-sures of 150 and 500 mTorr, respectively. The naturallinewidth of the neon 3.39-pm transition caused by thespontaneous decay of the upper and lower levels is 20+2MHz. The pressure broadening coefficients havebeen measured to be 12.4+4.9 MHz/Torr of neon and25.6+0.6 MHz/Torr of helium. ' These numbers are inrather considerable dispute from a survey of the litera-ture, however, our results indicate that the numbersquoted by Ohi are much more consistent with the datathan the generally higher numbers cited elsewhere. Thegenerally accepted value for the Doppler broadening ofthe transition in a dc discharge is 290+10 MHz.

From our operating conditions we calculate a homo-geneously broadened linewidth of 35+3 MHz giving aninhomogeneous to homogeneous linewidth ratio of 9:1 anda polarization decay rate of 10 sec '. These operatingconditions are the same as those used earlier in thediscovery of single-mode instabilities in the operation of a3.39-pm laser using a natural isotopic mixture of neon(91% Ne, 9% Ne). Note that the cavity linewidth isgreater than the polarization decay rate, a condition neces-sary for the observation of single-mode instabilities. '

The laser cavity could be tuned through variations inthe voltage applied to the piezoelectric crystal on whichthe spherical mirror was mounted. Nonlinearities in thecrystal expansion versus voltage were calibrated using anexternal He-Ne 0.6328-pm laser and the experimentallaser cavity as a passive cavity interferometer.

The laser was operated in a cold-cathode mode with dcexcitation using a 1 —3 kV power supply and a 100-kQseries ballast resistor. The power supplies used were mea-sured to have less than 0.2% ripple for the maximumpower output used in these experiments.

In order to measure the laser operating frequency, a

FIG. 2. Schematic of optical bench setup for detection ofsingle-mode laser output and providing of a reference laser fre-

quency: reference lasers L 1, L2; laser under study, L„' externaloptical amplifier, AMP, quartz lens, Q; InAs detector, D;reverse-biased InAs detector, HD; piezoelectric mirror drives, P;high-voltage optical amplifier, HV; ramp generator, R; oscilla-

tor, 0; lock-in amplifier, LA; spectrum analyzer, SA; fast oscil-

loscope, FO; integral preamplifier, A; ir line filter, I'; plotter, X-

P; chopper, C.

single-mode reference laser system was designed as shownin Fig. 2. A first single-mode laser was feedback stabi-lized to the peak of the its output profile after amplifica-tion by an external gain cell. Modulating the piezoelectriccrystal voltage at a 5-KHz rate with an amplitude that re-sulted in an optimal frequency modulation of +5 MHz,an error signal was generated to keep the laser locked towithin 0.5 MHz of the peak. As this modulated laser wasunsuitable for service as the desired reference laser, asecond single-mode laser was added and heterodyned withthe first. The resulting modulated beat note generated anerror signal at a second lock-in amplifier after detectionthrough an rf tuned filter. The second laser could thus belocked to a frequency offset from the first, within 0.5MHz of the center of the tuned filter. The tuned filter(actually an rf spectrum analyzer used in a tuned filtermode) could be adjusted to set this stabilized, unmodulat-ed reference laser to any chosen value within the central100 MHz of its tuning range. A portion of the output ofthe stabilized (but un modulated) reference laser washeterodyned with the laser under study with the resultingbeat note providing the desired information on the laseroperating frequency (or frequencies). The same detectorwas used to observe the rf power spectrum of the pulsa-tions of the laser when it went unstable.

IV. EXPERIMENTAL RESULTS

Using a programmable high-voltage power supply, ascan of power output discharge current was obtained forfixed cavity length. The laser was allowed to come tothermal equilibrium at an operating current of about 8mA before the scan was run so that the sudden heating bythe increased current would not cause thermal drifts in thecavity length. The results are shown in Fig. 3. The

Page 3: Anomalous mode pulling, instabilities, and chaos in a single-mode, standing-wave 3.39-μm He-Ne laser

R. S. GIOGGIA AND N. B. ABRAHAM 29

I e

0

CL

QI

CL

IcoCO

0—0 5 10 15

Discharge Current (mA)

/1 /

/I

II

III. /

-300t

300

10

I1/3CI$

1Q

IIIII

I/

/IIII

201

- 10\1

300

II

-3000P1

0)

Q 2

Q.

I

0

3FIP21 I I I P'1 1

2F P2

I JI

II

II

III

III

300 0 300

0e

0CN

O

nIJQ

Cl

0'C

KtIt

V+ -1OO-CQ0'S

100

(a)

-300

(c)

300

FIG. 3. Average laser power output versus discharge current

using a programmable high-voltage power supply.

PiP1 1P2 2 P2

/ I

/

threshold discharge current is 5.2 0.1 mA. The initiallylinear behavior of the curve in the 5.5—8-mA range is tobe expected for a mixed broadening case and the rolloff athigher currents can be explained as evidence of dischargeexcitation of the lower lasing level. However, on the as-sumption that the linear relationship between gain andpower output continues at high currents, we can use thepower output to scale the threshold conditions of the lasereven in the nonlinear region of Fig. 3.

Figure 4(a) shows a laser power output scan versus cavi-ty detuning. Using the measurements of the actual laseroperating frequency the full width at half maximum(FWHM) of the Lamb dip is found to be 52+6 MHz.This is in reasonable agreement with the predicted homo-geneously broadened linewidth of 35+3 MHz and theoriesof high intensity layers.

In addition, the value of the pulsation frequency (21MHz) at the instability threshold (laser threshold parame-ter —1.7) should provide a sensitive prediction of thehomogeneous linewidth (cf. Refs. 16—20). While formu-las are not provided for the particular case of He-we, esti-mates of the homogeneous linewidth [using, for example,

Eq. (5.1) of Ref. 19 and similar results in Refs. 16—20]yield values in the range of 21 —50 MHz. The agreementis quite good but more general theory and more precisespectroscopic data on the 3.39 transition are required be-fore precise comparisons can be made.

Figures 4(a)—4(d) show combinations of laser poweroutput versus cavity length, peaks in the intensity powerspectrum (cf. Fig. 5) versus cavity length, and laseroperating frequency versus cavity detuning. A single-mode instability is observed at the center of the Lamb dipas seen previously using natural isotopic neon.

Figure 5 shows sample rf power spectra of the laser in-

tensity from the regions of instabilities shown in Fig. 4.With small changes in detuning as shown on the expandedscales of Figs. (4c), and 4(d); the pulsation characterchanges from high-Q pulsations to two-frequency opera-tion, or gives evidence of a "subharmonic bifurcation"(period doubling) with other cases [such as the transition

I 1300

Cavity

I, . I

-300

Frequency (MHz)300

from 5(b) to 5(c)] showing broadening of the peaks and theaddition of a broadband spectral component characteristicof deterministic chaos. Details of the transition of single-mode laser systems from periodic to chaotic instabilitiesare the subject of recent theoretical'7'~ and experimen-tal'z'2 reports. (See Ref. 41 for general discussions ofdynamical instabilities and the various routes to chaoticbehavior in diverse nonlinear systems. ) An unusualfeature is the three-frequency spectrum shown in Fig. 5(e).Three-frequency operation is of considerable interest be-cause of theoretical disputes over its possibility whichseem to be only recently resolved.

Two indirect methods were used to estimate the laserthreshold conditions at line-center operation for differentcurrents. The threshold parameter is defined to be the ra-tio of the unsaturated gain at the lasing threshold. Fromthe power output versus current in Fig. 3 an initiallylinear dependence can be noted and a clear threshold valueis evident. Assuming that nonlinear effects may be due toexcitation of the lower atomic level and that the poweroutput scales linearly with threshold parameter the values

FIG. 4. Average laser power output (dashed line) versus cavi-

ty detuning with frequencies of principal peaks in the rf powerspectrum of the observed single-mode instabilities as indicated

(left). Laser operating frequency or frequencies as determined

from heterodyned beat notes with the reference laser versus cavi-

ty frequency (right). Discharge currents are (a) 8.0, (b) 10.8, (c)11.8, and (d) 16.5 mA. Expanded scale regions in (c) and (d)

show limits of different dynamical behavior such as simple

periodic pulsations (P1) as in Fig. 5(a), the appearance of asubharmonic or "period-2" oscillation (P2) as in Figs. 5(b) and

5(c), the appearance of an addition subharmonic or "period-4"(P4), simultaneous operation of two relatively incommensuratefrequencies (2F) as in Figs. 5(d) and 5(f), and the relatively

anomalous three-frequency region [5(e)]. Expanded scale forFig. 4(c), —41 MHz to + 15 MHz and for Fig. 4(d), —31 MHzto + 31 MHz.

Page 4: Anomalous mode pulling, instabilities, and chaos in a single-mode, standing-wave 3.39-μm He-Ne laser

ANOMALOUS MODE PULLING, INSTABILITIES, AND CHAOS. . . 1307

-10- -10 ' e10 s

-30

E-SO- s

0OlCL

CO \a

I-10-0

EL

t 4

20 40 60

-30 ~

~10 ~

20k e

40

-30

60 0 20 40 60

-10

a30 a

-S0 ~

0t

20 40

-30

~l~ Jl ~~0 20 40

Frequency (MHz)

-30

-SO

0I

20

FKr. 5. Sample rf power spectra from the regions of instability indicated in Fig. 4. Samples show typical high-Q period doublingwith a broadband chaotic background (b), period doubling without chaos (c), and the appearance of two distinct frequencies from aninitial near degeneracy in the sequence (d-fl from quasiperiodic (d) aud multiply periodic (e) behavior to stable two-frequency opera-tion (d).

TABLE I. Calculated laser threshold parameter values.

Discharge current(mA)

Threshold parameter valuesfrom power output from detuning range

8.010.811.816.5

1.49+0.051.90+0.052.1 +0.12.6 +0.2

1.57+0.041.73+0.041.83+0.021.94+0.02

in the first column of Table I could be estimated. How-ever, for conditions of mixed homogeneous and inhomo-geneous broadening no analytic scaling law is available,nor does one exist for operation in the presence of instabil-ities. The line-center threshold parameter values tabulatedin the second column were calculated by taking the mea-sured tuning range of the laser operation and the assump-tion of an approximately Gaussian gain profile with a292-MHz FWHM {[(hvD)z+(hvr, ) ]'i ). As the laserwas stable at the extremes of its tuning range this calcula-tion is unaffected by instabilities and thus may be themore reliable estimate of the unsaturated small signalgain. In any case, it is likely that the true threshold condi-tions were bounded by the values calculated from thesetwo methods. The discrepancies for higher currents can-not readily be explained quantitatively without a more de-tailed theory of laser operation well above threshold forthe unstable single-mode laser. However, the largerthreshold parameter values based on the power output areconsistent with the results of calculations by Caspersonwhich show increased average power for unstable pulsa-tions in comparison with the cw average power predictedby the steady-state solutions.

A distortion of the nearly linear mode pulling at lowgain appears as the laser is operated further above thresh-old. First a kink develops [Fig. 4(a)] and then an actual

break occurs [Fig. 4(c)). The region of the break is not tobe interpreted as the overlapping or hysteresis of single-frequency operation. Rather, observation of the beat fre-quency between the laser under study and the single-frequency laser shows the emergence of two equally strongbeat frequencies.

Bennett's original prediction of high-gain operationwas for a multivalued equation of state which would man-ifest itself as hysteresis in the detuning plots. The morerecent theoretical work has demonstrated that the steady-state single-frequency solution is unstable' so our ob-servations must be taken as measurements of the dynami-cal state of the self-pulsing laser and not of the predictedsteady-state solution. %e observe in Fig. 4 that instabili-ties have already emerged for these threshold conditionsand that the apparent multivalued-frequency region re-quires even higher threshold values with coincidentlystronger and more complex instabilities.

V. DISCUSSION

Bennett has provided an approximate formula for theanomalous mode pulling expected in a standing-wavelaser:

Cv„„=vi+ hys{vr)+b, yh, i,(vr),F/l.where v„„ is the cavity frequency, vr is the laser oscilla-tion frequency, F is the fractional loss per pass, L is thelength of the gain medium, hyz is the phase shift for theunsaturatrxi gain profile, and b,yh, r, is the phase shift forthe hole burned in the gain profile. Using values suitableor typical for our experimental setup, the curves shown inFig. 6 were generated using Bennett's approximate formu-la. The formula is limited in applicability to near thecenter of the Doppler-broadened profile and does not takeinto account power broadening of the hole or variation in

Page 5: Anomalous mode pulling, instabilities, and chaos in a single-mode, standing-wave 3.39-μm He-Ne laser

1308 R. S. GIOGGIA AND N. B. ABRAHAM 29

fQQ I p& s 0 I $1 I f s

)$ $ s d

)s s ~, ~ s

II I s s

e

~ 50

OCIU'Iu. QcsC

CdIClCL

OI

-iQQl a ~ ~ ~ s ~ a a ~ I a s s a s ~ y ~ a I ~

-SQ QCavity Frequency (MHz)

1QQ

OCIcrd)

0

-50CO

cd

~a «««

t i s a s I I s I i I i

"100 0 100Cavity Frequency (MHz)

FIG. 6. Plot of Bennett's mode [Eq. (1)] for steady-state laserfrequency versus cavity frequency using He-Ne parameters ofhvD ——300 MHz, Av, = 150 MHz, hvg ——25 MHz, I' =0.3,I. =0.1, and threshold values of r =1.0 (4), 1.67 (0), and 3.0(o).

the hole depth with detuning.The qualitative agreement between the theoretical

curves and the experimental ones shown in Fig. 4 indi-cates that even in the presence of the instabilities, the laseroperates close to the steady-state prediction for the opticalfrequency. Figure 7 shows an expanded scale version ofthe data in Fig. 4(c) including the weak sidebands that aregenerated by the instability. A reasonable agreement withthe theoretical curve for the steady-state solution is ob-tained.

The changes in the shape of the Lamb dip in the poweroutput versus detuning curve are distinctive with the onsetof strong self-pulsing. The dip appears to deepen and nar-row as the pulsing becomes stronger. The dip widens asthe laser is raised even further above threshold.

Simple low-gain interpretations of the Lamb dip cannoteasily be extended to this case. Not only is there strongmode pulling and the associated distortions predicted byBennett, but the unstable operation of the laser means thatthere are significantly strong optical sidebands to the mainlaser frequency. Often there are two equally strong opti-cal frequencies. No simple mapping is then possible be-tween the cavity frequency and operating frequency as thelatter no longer exists uniquely. Similarly, any simplehole-burning picture breaks down as well. A full dynami-cal model for the standing-wave case is likely to be re-quired to explain these observations in detail.

Study of the detuning curves indicates that the coexist-ing frequencies are within 20 MHz of each other. Theswitchback in the steady-state curves from the Bennetttheory shows a separation of the curves thai is more near-

ly equal to the homogeneous linewidth. The combinationof a narrow Lamb dip under stable operating conditionsand the close spacing of the coexisting frequencies maysuggest that either or both of the pressure-broadening ef-fects or the natural linewidth have been overestimated pre-viously. Detailed study of standing-wave laser operatingconditions in the presence of single-mode instabilities will

FIG. 7. Expanded scale plot of Fig. 4(d) data including fre-

quencies of sidebands to the primary laser frequencies.

be required before these discrepancies can be resolved.Preliminary studies have indicated qualitative agreementwith the features of an enhancement of the instability atthe Lamb dip, a dip in the pulsing frequency at the Lambdip, and a narrow range of detunings that support the in-stability.

VI. SUMMARY

Studies of the single-mode laser operation in a high-gain medium using a low-g, Fabry-Perot cavity indicateboth spontaneous pulsations and anomalies in the modepulling. Features include a narrow range of pulsations atthe Lamb dip, and an increase in the depth (and narrow-ing of the width) of the dip in the power output versus de-tuning curves. A link is thus established between thedispersive effects of hole burning predicted for steady-state laser operation and the more recently predicteddispersion-induced instabilities for hole burning in inho-mogeneously broadened lasers.

Note added in proof. It has been brought to our atten-tion that the high intensity gas laser theory papers ofRefs. 37—39 predict population pulsations in "steadystate" which may be the precursors of a Lamb dipenhanced laser instability. We also note with interest anew paper on mode pushing in a single mode laser.

VII. ACKNOWI. EDGMENTS

The authors wish to thank R. Tench, R. MacDowall, J.Bentley, M. Maeda and J. Wesson for their preliminarywork on this laser system and associated designs. Usefuldiscussions with N. J. Halas and S. P. Adams aided in theexperiments reported here. Helpful conversations with L.Casperson, L. Lugiato, P. Mandel, and L. Narducci aidedin the interpretation of these results. This work was sup-ported in part by grants from Research Corporation andthe National Science Foundation (Cxrant No. ECS-82-10263) and by an Alfred P. Sloan Research Fellowship(N.B.A.).

Page 6: Anomalous mode pulling, instabilities, and chaos in a single-mode, standing-wave 3.39-μm He-Ne laser

ANOMALOUS MODE PULLING, INSTABILITIES, AND CHAOS. . . 1309

'On sabbatical leave from Department of Physics, WidenerUniversity, Chester, PA 19013 (present address).

W. R. Bennett, Jr., Phys. Rev. 126, 580 (1962).W. R. Bennett, Jr., Appl. Opt. Suppl. No. 2, 3 (1965).

3W. R. Bennett, Jr., The Physics of Gas Lasers (Gordon andBreach, New York, 1977).

4W. E. Lamb, Jr., Phys. Rev. 134, A1429 (1964); M. Sargent,III, M. Scully, and W. E. Lamb, Jr., Laser Physics (Addison-

Wesley, Reading, Mass. , 1974).5B. L. Gyorffy, M. Borenstein, and W. E. Lamb, Jr., Phys. Rev.

169, 340 (1968).W. R. Bennett, Jr., S. F. Jacobs, J. T. LaTourette, and P. Rabi-

nowitz, Appl. Phys. Lett. 5, 56 (1964).7H. S. Boyne, M. M. Birky, and W. G. Schweitzer, Appl. Phys.

Lett. 7, 62 (1965).sE. I. Tsetsegova, Opt. Spektrosk. 31, 319 (1971) [Opt. Spec-

trosc. (USSR) 31, 172 (1971)].9L. W. Casperson and A. Yariv, Phys. Rev. Lett. 26, 293 (1971).~OP. W. Smith and P. J. Maloney, Appl. Phys. Lett. 22, 667

(1973).J. Bentley and N. B.Abraham, Opt. Commun. 41, 52 (1982).N. B. Abraham, T. Chyba, M. Coleman, R. S. Gioggia, N. J.Halas, L. M. Hoffer, S.-N. Liu, M. Maeda, and J. C. Wesson,in Laser Physics, Vol. 182 of Lecture Notes in Physics, edited

by J. D. Harvey and D. F. Walls (Springer, New York, 1983),p. 107.L. W. Casperson and A. Yariv, Appl. Phys. Lett. 17, 259(1970).

t4L. W. Casperson, IEEE J. Quantum Electron. OE-14, 756(1978).L. W. Casperson, Phys. Rev. A 21, 911 (1980);23, 248 (1981).

~6M. L. Minden and L. W. Casperson, IEEE J. Quantum Elec-tron. OE-18, 1952 (1982).L. W. Casperson, in Laser Physics, Vol. 182 of Lecture Votesin Physics, edited by J. D. Harvey and D. F. Walls (Springer,New York, 1983), p. 88.

~8S. Hendow and M. Sargent III, Opt. Commun. 40, 385 (1982);43, 59 (1982).

~9P. Mandel, Opt. Commun. 44, 400 (1983); 45, 269 (1983); inCoherence and Quantum Optics V, edited by L. Mandel and E.Wolf (Plenum, New York, in press).L. A. Lugiato, L. M. Narducci, D. K. Bandy, and N. B. Abra-ham, Opt. Commun. 46, 115 (1983); in Coherence and Quanturn Optics V, edited by L. Mandel and E. Wolf (Plenum, NewYork, in press).

~~M. Maeda and N. B.Abraham, Phys. Rev. A 26, 3395 (1982).~~R. S. Gioggia and N. B.Abraham, in Coherence and Quantum

Optics V, edited by L. Mandel and E. Wolf (Plenum, NewYork, in press).R. S. Gioggia and N. B. Abraham, Phys. Rev. Lett. 51, 650(1983).L. E. Urbach, S.-N. Liu and N. B.Abraham, in Coherence andQuantum Optics V, edited by L. Mandel and E. Wolf {Plenum,New York, in press).R. S. Gioggia and N. B. Abraham, Opt. Commun. (to be pub-lished).D. Hotz, Appl. Opt. 4, 527 (1965); Physica (Utrecht) 33, 746(1967).

~7I. P. Konovalov, A. I. Popov, and E. D. Protsenko, Opt. Spek-trosk. 33, 198 (1972).

~ E. Kotlikov, G. Todorov, and M. Chaika, Opt. Spektrosk. 30,185 (1971).

I. M. Beterov, Yu. A. Matyugin, and V. P. Chebotvae, Opt.Spektrosk. 28, 357 (1970).A. F. Mukhamedgalieva, V. M. Tatarenkov, A. N. Titov, andA. V. Uspenskii, Opt. Spektrosk. 30, 798 (1971).M. Ohi, Jpn. J. Appl. Phys. 12, 1377 (1973).W. Dietel and G. Jahn, Z. Ang. Phys. 26, 206 (1969).

33A. Dienes, IEEE J. Quantum. Electron. OE-5, 162 {1969).3'. T. Mevries, A. Dienes, and N. George, IEEE J. Quantum.Electron. OE-6, 117 (1970).M. A. Gubin, A. I. Popov, and E. D. Protsenko, Opt. Spek-trosk. 25, 756 (1968).K. R. Manes and A. E. Siegman, Phys. Rev. A 4, 373 (1971)~

37S. Stenholm and W. E. Lamb, Jr., Phys. Rev. 181, 618 (1969).3 B.J. Feldman and M. S. Feld, Phys. Rev. A 1, 1375 (1970).

S. Stenholm, Phys. Rev. A 2, 2089 (1970); S. Stenholm, Prog.Quantum Electron. 1, 239 (1971).

4oR. Graham and Y. Cho, in Coherence and Quantum Optics V,

edited by L. Mandel and E. Wolf (Plenum, New York, inpress); Opt. Commun. 47, 52 (1983).

~ J.-P. Eckmann, Rev. Mod. Phys. 53, 643 (1981).C. Grebogi, E. Ott, and J. A. Yorke, Phys. Rev. Lett. 51, 339(1983).

M. L. Minden and L. W. Casperson (private communication).~J. C. Wesson, M. Maeda, and N. B. Abraham, J. Opt. Soc.

Am. 72, 1811 (1982).4~L. Lynds, A. W. Angelbeck, and B. Burdick, Opt. Commun.

47, 335 (1983).