You are on page 1of 6

Supplee et al. Vol. 15, No. 7 / July 1998 / J. Opt. Soc. Am.

B 1833

Response of a two-level atom to a


frequency-modulated
optically coherent pulse train

James M. Supplee
Department of Physics, Drew University, Madison, New Jersey 07940, and Department of Physics and Engineering
Physics, Stevens Institute of Technology, Hoboken, New Jersey 07030

Edward A. Whittaker
Department of Physics and Engineering Physics, Stevens Institute of Technology, Hoboken, New Jersey 07030

Keith Andrew
Department of Physics, Eastern Illinois University, Charleston, Illinois 61920

Received October 14, 1997; revised manuscript received March 16, 1998
We present an analytical expression for the response of a two-level atom to a frequency-modulated optically
coherent pulse train. The optical beam has sinusoidal frequency modulation and is chopped to have a square-
wave envelope. We assume that the laser pulses are short compared with the atomic-decay time, the pulse-
repetition time, and the modulation period. With this short-pulse assumption we are able to use a method
similar to Temkin’s [J. Opt. Soc. Am. B 10, 830–839 (1993)] and solve the optical Bloch equations in closed
form. © 1998 Optical Society of America [S0740-3224(98)02206-1]
OCIS codes: 020.1670, 190.0190, 300.6420.

1. INTRODUCTION become increasingly important as new high-power tun-


1–5 able laser sources emerge as a result of advanced semi-
Frequency-modulation (FM) spectroscopy is a useful
conductor laser research. For example, recent experi-
method for measuring absorption and dispersion and for
mental work in our laboratory involved sensitive pulsed
improving the sensitivity of optical spectroscopy measure-
ments. In the linear regime (optical intensity much less spectroscopic measurements using a frequency-
than the saturation intensity) absorption and dispersion modulated quantum-cascade semiconductor laser.17 The
are treated as functions of frequency, and the propagation quantum-cascade laser is a new laser device18 capable of
of the optical beam is calculated simply by addressing the delivering tunable high-power pulses at high-repetition
absorption and the dispersion of each Fourier component frequency in the infrared spectral region. While the
of the beam.1–4 In the nonlinear regime, however, ab- aforementioned published results were in the linear re-
sorption and dispersion are not simply functions of fre- gime, the power capabilities of the quantum-cascade laser
quency, since in this regime the optical beam itself modi- are steadily improving, and we are interested in deter-
fies the optical characteristics of the sample. One mining the effect that nonlinear optical effects may have
therefore needs to calculate the response of the material on the use of these devices in high-sensitivity spectro-
to the optical field. This usually involves working with scopic applications. This paper presents a technique for
the quantum-mechanical density matrix, or equivalently exploring these effects theoretically and will thus provide
(after a coordinate transformation), with the atomic po- a basis for interpreting future experiments that could si-
larization and inversion via the optical Bloch equations. multaneously involve sensitive detection, time resolution,
References 6–15 address the interaction of a frequency- and nonlinear optical effects.
modulated cw beam with a quantum-mechanical atom. A number of authors have studied frequency modula-
In this paper, however, we wish to calculate the response tion (or phase modulation) in combination with an optical
of a two-level atom to a frequency modulated and pulsed pulse train. These authors have addressed quite varied
beam. physical situations (despite having the superficial simi-
Our motivation for studying the response of an atom to larity of involving modulation and pulse trains). For ex-
a pulsed and modulated beam is this: Modulated absorp- ample, Gallagher and co-workers19–21 were able to obtain
tion spectroscopy affords a very sensitive means to detect FM spectra using a multimode pulsed dye laser. They
weakly absorbing species,16 and pulsed laser operation used large modulation frequencies, typically 1 to 12 GHz,
provides both access to higher peak power and the possi- so that the FM sidebands fell outside the power spectrum
bility to detect short-lived species. Understanding the of the (carrier) laser pulse. More recently, Eyler et al.22
response of an absorber to a pulsed modulated beam will have demonstrated FM spectra using pulsed amplifica-

0740-3224/98/071833-06$15.00 © 1998 Optical Society of America


1834 J. Opt. Soc. Am. B / Vol. 15, No. 7 / July 1998 Supplee et al.

tion of a phase-modulated cw laser. This method allows E ~ t ! 5 E 0 exp~ iM sin v m t ! . (3)


for a narrow (unmodulated) pulse, so that a modest 400-
MHz modulation is sufficient to place the sidebands well For the FM case addressed in this paper, the envelope
outside the power spectrum of the carrier pulse. Both of E (t) describes the sinusoidal phase modulation. For
these groups19–22 addressed cases that essentially in- later use we note that the real and imaginary parts of
volved the interaction of a ground-state atom with a E (t) are
single modulated pulse; subsequent pulses repeat the
same experiment and are used for signal averaging. The
case we will consider below is quite different in that the Re~ E ! 5 E 0 cos~ M sin v m t ! ,
atom will not have time to return to the ground state be-
tween pulses, so that we are considering coherent effects Im~ E ! 5 E 0 sin~ M sin v m t ! . (4)
as later pulses interact with the still excited atom. For
our calculation then, optical coherence from pulse to pulse If this field is used in the optical Bloch equations, the
is essential. terms shown in Eqs. (4) will enter as time-dependent co-
Other authors have presented calculations addressing efficients in coupled differential equations. Such equa-
various cases (again different from our case) of the inter- tions are not always amenable to simple, closed-form
action of an atom with a phase-modulated and pulsed op- analysis for arbitrary field strength (nonlinear regime)
tical field. Peterson and Gavrielides23 use Fourier series and arbitrary modulation index M. How tractable the
to calculate the periodic response of the Bloch variables to answers are depends largely on what additional assump-
a train of quadratically chirped Gaussian pulses. Vi- tions and approximations can be applied to the circum-
tanov and Knight24 calculate the response of a two-level stances at hand. (See, for example, Refs. 6–15.)
atom to N identical pulses; they neglect relaxation and For a frequency-modulated beam the additional com-
are able to address a wide variety of pulse envelopes in- plexity of treating a pulsed beam can result, perhaps
cluding chirped pulses. They present a formalism for ex- ironically, in simpler Bloch equations than the case of a
tending single-pulse interaction results to the N-pulse frequency-modulated continuous beam. This brings us to
case. Hioe25 calculates the response of a two-level atom the central point of this paper: If the laser pulses are
to a single pulse with hyperbolic secant amplitude modu- sufficiently short [Eqs. (8) and (9) below], the real and
lation and with hyperbolic tangent frequency modulation. imaginary parts of the envelope (4) will not vary signifi-
Bradley26 is interested in the backscatter from mesos- cantly during the pulse. Therefore, as elaborated below,
pheric sodium and calculates the response of both two- the optical Bloch equations that apply during the pulse
level atoms and sodium atoms to a resonant, coherent are coupled differential equations with constant coeffi-
Gaussian pulse train. Bradley also considers the possi- cients (albeit different constants for each pulse). The
bility of phase modulation and concludes that modulation equations can therefore easily be solved in closed form.
can increase sodium backscatter by reducing velocity hole During the time that the pulse is off, the field is zero and
burning. Thomas27 considers the interaction of a multi- the Bloch equations have their simple free decay solu-
level system with a train of identical Gaussian pulses. tions. The solutions for ‘‘pulse off’’ and the solutions for
Milonni and Thode28 consider the response of a two-level ‘‘pulse on’’ can be joined continuously at the boundary.
atom to a perfectly coherent and unmodulated pulse The net result is that, with the short-pulse restriction, we
train. are able to find solutions to the optical Bloch equations for
In this paper we consider a situation different from all time. The solutions apply for arbitrary intensity and
those addressed by the authors mentioned above. We for any modulation index that does not violate the short-
consider an optical beam with sinusoidal frequency modu- pulse restriction [Eq. (8) below].
lation. The beam is first modulated, then chopped, but
leaving the pulses optically coherent from pulse to pulse.
Our pulses therefore simply consist of different sections
chopped from one continuous wave given in Eq. (1) below. 2. THEORY AND CALCULATIONS
This is not the same as the case of time-delayed identical A. Model
pulses. (Even without modulation this distinction is cru- How a two-level atom responds to an optical field can be
cial. For example, a chopped sine wave has the same studied via the optical Bloch equations29 in the dipole ap-
carrier phase in each pulse, whereas with identical but proximation and the rotating-wave approximation,
time-delayed copies of one pulse, the optical phase is to-
tally determined by the time delay. This distinction is
u̇ 5 2b u 2 Dv 1 k Im~ E ! w,
further discussed by Vitanov and Knight.24) The electric
field in our case (while the pulse is on) is a plane wave
given by the real part of v̇ 5 Du 2 b v 1 k Re~ E ! w,

E ~ t ! 5 E 0 exp@ i ~ v 0 t 1 M sin v m t !#
ẇ 5 2k Im~ E ! u 2 k Re~ E ! v 2 g ~ w 1 1 ! . (5)
5 E 0 exp~ iM sin v m t ! exp~ i v 0 t ! . (1)
u and v are, respectively, the (slowly varying) in-phase
The last arrangement shows the field written as a slowly
and in-quadrature parts of the atomic dipole moment; w
varying envelope, E(t), times the carrier,
is the (slowly varying) atomic inversion. The 1 in the
E ~ t ! 5 E ~ t ! exp~ i v 0 t ! , (2) third Bloch equation arises because in this form of the
Supplee et al. Vol. 15, No. 7 / July 1998 / J. Opt. Soc. Am. B 1835

equations we assume that all atoms would be in the decay, terms involving g and b disappear. (When we
ground state (w 5 21) in the absence of an external field. treat decay in the following subsection, we will consider a
We use a frame rotating at the carrier frequency v 0 . collisionless atom and set b 5 g /2. One need not neces-
Background information is given, for example, in texts by sarily commit to that restriction here, provided the pulse
Milonni and Eberly,30 Meystre and Sargent,31 Allen and time is short compared with all decay times.) The Bloch
Eberly,32 and Shore.33 We use detuning equations further simplify because, as discussed above,
Im(E ) and Re(E ) are approximately constant for the du-
D [ v atom 2 v 0 , (6) ration of one pulse. The optical Bloch equations there-
where v 0 is the laser carrier frequency and v atom is the fore become
atomic transition energy divided by \. k is the electric
dipole moment of the transition divided by \. g is the to-
u̇ 5 2Dv 1 b n w,
tal decay rate for all decay processes connecting the upper
and lower levels; b is the dipole-moment decay rate (1 /g
is often called T 1 , or the longitudinal lifetime, and 1 /b is v̇ 5 Du 1 a n w,
often called T 2 , or the transverse lifetime).
As mentioned above, if the optical Bloch Eqs. (5) in-
volve time-dependent coefficients, the resulting coupled ẇ 5 2b n u 2 a n v,
differential equations are intractable to various degrees
depending on what other assumptions apply to the model
or
at hand. For a pulsed FM optical field, both Im( E ) and

FG F GF G
Re( E ) are time dependent; the pulsing alone would
make Re( E ) a square wave, and the modulation affects 0 2D bn
u u
both the real and the imaginary parts of E via Eqs. (4). d
v 5 D 0 an v . (10)
We overcome the square-wave difficulty by using dt
Temkin’s34 approach; that is, we solve the Bloch Eqs. (5) w 2b n 2a n 0 w
separately for the time the pulse is on and for the time the
pulse is off, then match values at the boundary. We We introduce the abbreviations
overcome the second difficulty, the time-dependent modu-
lation, by making use of the short-pulse requirement; for
adequately short pulses, the values of Re(E ) and Im(E ) a n [ k Re@ E ~ t !# 5 x cos~ M sin v m t n ! ,
[Eqs. (4)] will change negligibly during the time that the
pulse is on.
b n [ k Im@ E ~ t !# 5 x sin~ M sin v m t n ! , (11)
B. Optical Bloch Solution during the Pulse
The key simplification in our model involves noting that where x 5 k E 0 is the Rabi frequency on resonance for the
the real and the imaginary parts of E, given by Eqs. (4), unmodulated beam. Recalling that the change in t is
will vary little during the pulse provided that the pulse negligible during the pulse, we simply replace the vari-
time is short enough that the argument of sine and cosine able t with the particular time at which the pulse of in-
does not vary much during the pulse. We are therefore terest is centered. We number the pulses, so that the
addressing the case where the variation of the argument first pulse starts at t 5 0 and is centered on t p / 2; the nth
satisfies the condition pulse is centered on t n 5 (n 2 1)t r 1 t p / 2. The net re-
d ~ M sin v m t ! ! 2 p . (7) sult is that a n and b n are constant, but different con-
stants for each pulse. a n and b n are, of course, not inde-
This condition is satisfied provided that the pulse dura- pendent since a n 2 1 b n 2 5 x 2 .
tion t p (which equals the variation in time d t) satisfies The ‘‘pulse on’’ optical Bloch Eqs. (10) are in the form
we sought: coupled differential equations with constant
2p
tp ! . (8) coefficients. What is more, the matrix in Eqs. (10) is an
Mvm antisymmetric 3 3 3 matrix, so it has especially simple
Since we are considering short pulses, it is also acceptable eigenvalues, 0 and 6 iV, where V [ ( x 2 1 D 2 ) 1/2 is the
to neglect atomic decay (following Temkin34) during the generalized Rabi frequency. This means that the trans-
time that the pulse is on, provided that the pulse is much formation matrix A n (below) will be orthogonal, and will
shorter than the atomic lifetime: show oscillations at the familiar generalized Rabi fre-
quency. The solution to Eqs. (10) is
1

F G FG
tp ! . (9)
g
u~ t ! u n8
In many cases it is justified to think of Eqs. (8) and (9) as v~ t ! v n8
roughly one restriction, not two, because experimental 5 A n~ t ! , (12)
w~ t ! w n8
situations meeting condition (9) will often meet condition 1 1
(8) as well.
We now rewrite the Bloch Eqs. (5) in the simplified
form that applies during the pulse. Because we neglect where
1836 J. Opt. Soc. Am. B / Vol. 15, No. 7 / July 1998 Supplee et al.

F G
a 2n 1 ~ D 2 1 b 2n ! cos V t 2a n b n ~ 1 2 cos V t ! 2 DV sin V t 2Da n ~ 1 2 cos V t ! 1 Vb n sin V t 0
1 2a n b n ~ 1 2 cos V t ! 1 DV sin V t b 2n 1 ~D 1
2
a 2n ! cos Vt Db n ~ 1 2 cos V t ! 1 Va n sin V t 0
A n~ t ! 5 .
V 2 2Da n ~ 1 2 cos V t ! 2 Vb n sin V t Db n ~ 1 2 cos V t ! 2 Va n sin V t D 2 1 x 2 cos V t 0
0 0 0 V2
(13)

The four by four format seems superfluous here, but it 3. RESULTS


will be needed in the next subsection when we include de-
The matrices A n and B above show how the Bloch vari-
cay as an inhomogeneous term in the optical Bloch equa-
ables evolve with time. We can consider time as a con-
tions. Equations (12) and (13) give u, v, and w during
tinuous variable, and use the dummy variable t to track
the nth pulse. u n8 , v 8n , and w 8n refer to the values of u, v,
the Bloch variables’ continuous evolution. In some cases,
and w at the beginning of the nth pulse. Also, in this
however, one might just need to calculate the effect of
particular context, t refers to the time elapsed since the
complete pulses on the atom. In that case, one can use
beginning of the nth pulse.
t 5 t p in Eq. (12), and a simple matrix multiplication
propagates the Bloch variables through one complete
C. Optical Bloch Solution between Pulses (Decay)
pulse. Similarly, using t 5 t r 2 t p in Eq. (15) will
Between pulses the optical field is zero and the atom de-
propagate the Bloch variables through one ‘‘dead time’’
cays. With zero field present, and considering a collision-
between pulses. For example, if an atom starts in the
less atom with b 5 g /2, the Bloch Eqs. (5) simplify to
ground state, then at a time t after the jth pulse turns off,
the atom’s Bloch vector is given by

FG FG
g
u̇ 5 2 u 2 Dv,

F) G
2 u 0
1
v 0
5 B ~ t ! A j~ t p ! B ~ t r 2 t p ! A n~ t p ! .
g w n5j21 21
v̇ 5 Du 2 v, 1 1
2
(17)
The matrix A, which propagates the Bloch variable
ẇ 5 2g ~ w 1 1 ! . (14) through a pulse, contains a time dependence conveyed via
the subscript. This is because the pulse period and the
The solution to these equations is modulation period are different, so each pulse arrives

F G FG
when the modulation is at a different phase. Informa-
tion about the phase of the modulation at the arrival of
u~ t ! u n9
the nth pulse is conveyed through the t n by using Eqs.
v~ t ! v n9
5 B~ t ! , (15) (11) and (13). (We use t as a dummy variable to track
w~ t ! w 9n the time since the last change in the square-wave ampli-
1 1 tude envelope, but we use t to represent the total elapsed
time in the experiment.) We make no assumptions about
where whether or not the modulation period and pulse period

F G
have an integer ratio. Because each pulse is slightly dif-
ferent, it is not possible (at least in general) to find the
x cos D t 2x sin D t 0 0 steady-state behavior of the atom by finding the eigenvec-
x sin D t x cos D t 0 0 tor of BA having eigenvalue 1. With identical pulses or
B~ t ! 5 . (16) with simply commensurate periods, the eigenvalue analy-
0 0 x2 x2 2 1
sis can be helpful.26,34
0 0 0 1 Equation (17), along with the equations giving A and
B, is the main result of this paper. With Eq. (17), one
Here x [ exp(2gt /2), simply for abbreviation. The no- can study the Bloch vector behavior while varying the pa-
tation used in these pulse-off equations is parallel to that rameters of interest. There is a large parameter space to
of the previous pulse-on equations. That is, Eqs. (15) and explore, since one can vary M, v m , D, g, t r , t p , and pulse
(16) give u, v, and w during the pulse-off time between area. We give examples in Figs. 1–3.
the nth pulse and the next. u n9 , v n9 , and w n9 refer to the Figure 1 shows inversion w versus detuning for fixed
values of u, v, and w at the end of the nth pulse; that is, values of the other parameters. The results are plotted
they are the ‘‘initial’’ values for the purposes of this time after 30 complete pulse cycles and also after 50 complete
interval. In a fashion parallel to that above, the t of Eqs. pulse cycles. The similarity of the two curves shows that
(15) and (16) refers to the time elapsed since the end of the overall structure is not critically sensitive to the num-
the nth pulse. Of course Eqs. (15) and (16) simply de- ber of pulses, provided one waits long enough for tran-
scribe the decay of the atom in the absence of an optical sients to die away. Transients are gone after a number
field. of pulses greater than ;4/g t r (Ref. 34). However, it is
Supplee et al. Vol. 15, No. 7 / July 1998 / J. Opt. Soc. Am. B 1837

v m 5 0.307 simply rigs the beating so that the inversion


will instantaneously be high after 30 cycles, which was
the number of cycles chosen in plotting Fig. 2.

4. CONCLUSION
The results above give the Bloch-vector behavior for a
two-level atom interacting with a frequency-modulated
pulse train. They can be used to explore parameter
Fig. 1. Inversion versus detuning. M 5 0.25, v m
space by varying M, v m , D, g, t r , t p and pulse area over
5 0.15 GHz, g 5 0.02 GHz, t r 5 10 ns, t p 5 0.2 ns, pulse area
a wide range, subject only to the short-pulse restriction
is p /8. The major peaks are at zero detuning and the first opti-
cal Ramsey fringe (t r D 5 2 p ). The modulation sidebands are given in Eqs. (8) and (9). In Figs. 1–3 we have illustrated
clearly visible. The solid curve shows w after 30 complete cycles some of the ways that varying parameters can affect in-
(the instant the 31st pulse turns on). The dashed curve shows version. This formalism can also be used to investigate
w after 50 complete cycles. the potential for using frequency-modulated pulse trains
in sensitive spectroscopic measurements of transient
atomic and molecular states.

ACKNOWLEDGMENTS
This work was partially supported by the National Sci-
ence Foundation under grant DMI-9313320.

REFERENCES
1. G. C. Bjorklund, ‘‘Frequency-modulation spectroscopy: a
Fig. 2. Inversion (after 30 complete cycles) versus modulation new method for measuring weak absorption and disper-
frequency. M 5 0.25, D 5 0.314 GHz, g 5 0.02 GHz, t r sion,’’ Opt. Lett. 5, 15–17 (1980); G. C. Bjorklund, ‘‘Method
5 10 ns, t p 5 0.2 ns, pulse area is p /8. w remains near 21 un- and device for detecting a specific spectral feature,’’ U.S.
til the modulation frequency is sufficient to make one sideband patent 4,297,035 (October 27, 1981).
nearly resonant with the atom. This curve is very sensitive to 2. J. M. Supplee, E. A. Whittaker, and W. Lenth, ‘‘Theoretical
the number of pulses. description of frequency modulation and wavelength modu-
lation spectroscopy,’’ Appl. Opt. 33, 6294–6302 (1994), and
references therein.
3. G. C. Bjorklund, M. D. Levenson, W. Lenth, and C. Ortiz,
‘‘Frequency modulation (FM) spectroscopy: theory of line-
shapes and signal-to-noise analysis,’’ Appl. Phys. B: Photo-
phys. Laser Chem. 32, 145–152 (1983).
4. J. A. Silver, ‘‘Frequency-modulation spectroscopy for trace
species detection: theory and comparison among experi-
mental methods,’’ Appl. Opt. 31, 707–717 (1992).
5. Xiang Zhu and D. T. Cassidy, ‘‘Modulation spectroscopy
with a semiconductor diode laser by injection-current
modulation,’’ J. Opt. Soc. Am. B 14, 1945–1950 (1997).
Fig. 3. Inversion versus time. M 5 0.25, D 5 0.314 GHz, v m
6. G. S. Agarwal, ‘‘Frequency-modulated spectra of coherently
5 0.307 GHz, g 5 0.02 GHz, t r 5 10 ns, t p 5 0.2 ns, pulse area driven systems,’’ Phys. Rev. A 23, 1375–1381 (1981).
is p /8. (Parameters correspond to a peak in Fig. 2.) The main 7. N. Nayak and G. S. Agarwal, ‘‘Absorption and fluorescence
oscillation, period 439 ns, is the beating of the pulse-repetition in frequency-modulated fields under conditions of strong
frequency with the second harmonic of the modulation frequency. modulation and saturation,’’ Phys. Rev. A 31, 3175–3182
(1985).
8. H.-R. Xia, J. I. Cirac, S. Swartz, B. Kohler, D. S. Elliott, J.
not possible to plot genuine steady-state results for this
L. Hall, and P. Zoller, ‘‘Phase shifts and intensity depen-
problem, because the problem is not strictly periodic. dence in frequency-modulation spectroscopy,’’ J. Opt. Soc.
Figure 2 shows inversion versus modulation frequency, Am. B 11, 721–730 (1994).
with the other parameters fixed. Figure 2 (unlike Fig. 1) 9. W. M. Ruyten, ‘‘Magnetic and optical resonance of two-level
is extremely sensitive to the number of pulses even after a quantum systems in modulated fields. I. Bloch equation
approach,’’ Phys. Rev. A 42, 4226–4245 (1990).
long time. This is because changing v m can cause sig- 10. M. A. Kramer, R. W. Boyd, L. W. Hillman, and C. R. Stroud,
nificant beating between the modulation frequency and Jr., ‘‘Propagation of modulated optical fields through
the pulse-repetition frequency. To illustrate the impor- saturable-absorbing media: a general theory of modula-
tance of this beating, we plot w(t) in Fig. 3, and we choose tion spectroscopy,’’ J. Opt. Soc. Am. B 2, 1444–1455 (1985).
11. A. Schenzle, R. G. DeVoe, and R. G. Brewer, ‘‘Phase-
the modulation frequency that corresponds to one of the
modulation laser spectroscopy,’’ Phys. Rev. A 25, 2606–
large maxima in Fig. 2. Figure 3 therefore illustrates 2621 (1982).
two interesting points. First, we see beating with a pe- 12. W. M. Ruyten, ‘‘Comment on absorption and fluorescence in
riod of 439 ns, which is the beat period between the pulse- strong frequency–modulated and amplitude-modulated
repetition frequency and twice the modulation frequency. fields,’’ Phys. Rev. A 39, 442–444 (1989).
13. A. V. Alekseev and N. V. Sushilov, ‘‘Analytic solutions of
Second, Fig. 3 helps us understand the structure in Fig. 2: Bloch and Maxwell–Bloch equations in the case of arbitrary
It is not the case that v m 5 0.307 achieves much higher field amplitude and phase modulation,’’ Phys. Rev. A 46,
average inversion than, say, v m 5 0.314; it is rather that 351–355 (1992).
1838 J. Opt. Soc. Am. B / Vol. 15, No. 7 / July 1998 Supplee et al.

14. Ping Koy Lam and C. M. Savage, ‘‘Complete atomic popu- Bloch equations to a phase modulated pulse train; an appli-
lation inversion using correlated sidebands,’’ Phys. Rev. A cation to mesospheric sodium,’’ Opt. Commun. 104, 53–56
50, 3500–3504 (1994). (1993).
15. S. Feneuille, M.-G. Schweighofer, and G. Oliver, ‘‘Response 24. N. V. Vitanov and P. L. Knight, ‘‘Coherent excitation of a
of a two-level system to a narrow-band light excitation com- two-state system by a train of short pulses,’’ Phys. Rev. A
pletely modulated in amplitude,’’ J. Phys. B 9, 2003–2009 52, 2245–2261 (1995).
(1976). 25. F. T. Hioe, ‘‘Solution of Bloch equation involving amplitude
16. Jun Ye, Long-Sheng Ma, and J. L. Hall, ‘‘Ultrasensitive de- and frequency modulations,’’ Phys. Rev. A 30, 2100–2103
tections in atomic and molecular physics: demonstration (1984).
in molecular overtone spectroscopy,’’ J. Opt. Soc. Am. B 15, 26. L. C. Bradley, ‘‘Pulse-train excitation of sodium for use as a
6–15 (1998). synthetic beacon,’’ J. Opt. Soc. Am. B 9, 1931–1944 (1992).
17. K. Namjou, S. Cai, E. A. Whittaker, J. Faist, C. Gmachl, F. 27. G. F. Thomas, ‘‘Excitation of a multilevel system by a train
Capasso, D. L. Sivco, and A. Y. Cho, ‘‘Sensitive absorption of identical phase-coherent Gaussian-shaped laser pulses,’’
spectroscopy with a room-temperature distributed-feedback Phys. Rev. A 41, 1645–1652 (1990).
quantum-cascade laser,’’ Opt. Lett. 23, 219–221 (1998). 28. P. W. Milonni and L. E. Thode, ‘‘Theory of mesospheric so-
18. J. Faist, F. Capasso, C. Sirtori, D. L. Sivco, A. L. Hutchin- dium fluorescence excited by pulse trains,’’ Appl. Opt. 31,
son, and A. Y. Cho, Appl. Phys. Lett. 70, 2670–2672 (1997). 785–800 (1992).
19. T. F. Gallagher, R. Kachru, F. Gounand, G. C. Bjorklund, 29. R. P. Feynman, F. L. Vernon, Jr., and R. W. Hellwarth,
and W. Lenth, ‘‘Frequency-modulation spectroscopy with a ‘‘Geometrical representation of the Schrödinger equation
pulsed dye laser,’’ Opt. Lett. 7, 28–30 (1982). for solving maser problems,’’ J. Appl. Phys. 28, 49–52
20. N. H. Tran, R. Kachru, T. F. Gallagher, J. P. Watjen, and G. (1957).
C. Bjorklund, ‘‘Pulsed frequency-modulation spectroscopy 30. P. W. Milonni and J. H. Eberly, Lasers (Wiley, New York,
at 3302 Å,’’ Opt. Lett. 8, 157–159 (1983). 1988), especially Chaps. 6 and 8.
21. N. H. Tran, R. Kachru, P. Pillet, H. B. van Linden van den 31. P. Meystre and M. Sargent III, Elements of Quantum Op-
Heuvell, T. F. Gallagher, and J. P. Watjen, ‘‘Frequency- tics (Springer-Verlag, New York, 1990).
modulation spectroscopy with a pulsed dye laser: experi- 32. L. Allen and J. H. Eberly, Optical Resonance and Two-Level
mental investigations of sensitivity and useful features,’’ Atoms (Dover, New York, 1987).
Appl. Opt. 23, 1353–1360 (1984). 33. B. W. Shore, The Theory of Coherent Atomic Excitation:
22. E. E. Eyler, S. Gangopadhyay, N. Melikechi, J. C. Bloch,
Simple Atoms and Fields (Wiley, New York, 1990), Vol. 1,
and R. W. Field, ‘‘Frequency-modulation spectroscopy with
especially Chap. 8.
transform-limited nanosecond laser pulses,’’ Opt. Lett. 21,
34. R. J. Temkin, ‘‘Excitation of an atom by a train of short
225–227 (1996).
pulses,’’ J. Opt. Soc. Am. B 10, 830–839 (1993).
23. P. Peterson and A. Gavrielides, ‘‘Periodic response of the

You might also like