• Aucun résultat trouvé

Linear pulse propagation in a resonant medium : the adiabatic limit

N/A
N/A
Protected

Academic year: 2021

Partager "Linear pulse propagation in a resonant medium : the adiabatic limit"

Copied!
13
0
0

Texte intégral

(1)

HAL Id: jpa-00210499

https://hal.archives-ouvertes.fr/jpa-00210499

Submitted on 1 Jan 1987

HAL is a multi-disciplinary open access archive for the deposit and dissemination of sci- entific research documents, whether they are pub- lished or not. The documents may come from teaching and research institutions in France or abroad, or from public or private research centers.

L’archive ouverte pluridisciplinaire HAL, est destinée au dépôt et à la diffusion de documents scientifiques de niveau recherche, publiés ou non, émanant des établissements d’enseignement et de recherche français ou étrangers, des laboratoires publics ou privés.

Linear pulse propagation in a resonant medium : the adiabatic limit

B. Macke, J.L. Queva, F. Rohart, B. Segard

To cite this version:

B. Macke, J.L. Queva, F. Rohart, B. Segard. Linear pulse propagation in a resonant medium : the adiabatic limit. Journal de Physique, 1987, 48 (5), pp.797-808. �10.1051/jphys:01987004805079700�.

�jpa-00210499�

(2)

Linear pulse propagation in a resonant medium :

the adiabatic limit

B. Macke, J. L. Queva (*), F. Rohart and B. Segard

Université de Lille I, Laboratoire de Spectroscopie Hertzienne, Associé au CNRS, 59655 Villeneuve d’Ascq Cedex, France

(Requ le 5 août 1986, revise le 20 novembre, accepte le 13 decembre 1986)

Résumé.

2014

Grâce à un formalisme original d’opérateur nous donnons une description exacte de la propagation d’impulsions, dont l’enveloppe varie lentement (impulsions adiabatiques), dans un milieu lorentzien qui présente une dispersion anormale près de la résonance. Nous montrons que l’enveloppe complexe de l’impulsion émergente se déduit de celle de l’impulsion incidente en appliquant successivement un opérateur

de translation temporelle et un opérateur exp (s2 Z039B) où 039B est lui-même un opérateur, Z l’épaisseur optique du

milieu et s un (petit) paramètre d’adiabaticité caractérisant l’impulsion d’entrée. Le concept de vitesse de groupe est alors clairement relié à l’approximation d’ordre 0 par rapport à s2 Z (limite adiabatique) et nous

déterminons une limite supérieure rigoureuse des termes ainsi négligés. Le calcul est mené explicitement pour des impulsions incidentes d’enveloppe dn/dtn[exp(- 03B32t2) ] qui fournissent une base pour développer une impulsion quelconque de durée et d’énergie finies. Pour une large classe d’impulsions adiabatiques se propageant dans un absorbeur résonnant, nos calculs établissent de manière indiscutable que la forme de

l’impulsion sortante peut anticiper de façon significative celle de l’impulsion entrante, avec une distorsion aussi faible qu’on le veut. Nous montrons que ce résultat, corroboré par de récentes expériences, est tout à fait

compatible avec le principe de causalité. Il confirme que le concept de vitesse de groupe a un sens, même

quand celle-ci devient négative. Par ailleurs, notre technique d’opérateurs nous a permis de déterminer facilement les corrections d’ordre supérieur à la forme de l’impulsion. La discussion théorique est bien illustrée par des simulations numériques.

Abstract.

2014

Using an original operator formalism, we give an exact description of the propagation of slowly varying envelope pulses (adiabatic pulses) in a Lorentzian medium, which is anomalously dispersive close to

the resonance. The complex envelope of the output pulse is shown to be derived from that of the input pulse, by applying successively a time-translation operator and an operator exp(s2 Z039B), where 039B is itself an operator, Z the medium optical depth and s a (small) adiabaticity parameter characterizing the input pulse. The concept of group velocity is then clearly related to the 0th order approximation with respect to s2 Z (adiabatic limit),

and a rigourous upper bound to the neglected terms is fixed. The calculation is explicitely achieved for resonant input pulses of envelope dn/dtn[exp (- 03B32t2)] which provide a basis to expand any pulse of finite

duration and energy. For a wide class of adiabatic pulses propagating in a resonant absorber, our calculations

indisputably prove that the output pulse shape can significantly anticipate the input one with a distortion as low

as wanted. This result, supported by recent experiments, is shown to be quite consistent with the causality principle and confirms that the group velocity is a meaningful concept even when it becomes negative.

Moreover, higher order corrections to the pulse shape are easily derived by our operator technique. The

theoretical discussion is well supported by numerical simulations.

Classification

Physics Abstracts

03.40K

-

41.10H - 42.10

1. Introduction.

The propagation of electromagnetic pulses in a dispersive linear medium is a very classical problem, widely accounted for in authoritative textbooks [1-4]

and in recent review papers [5-7]. When the pulses

have a slowly time-varying envelope or, equivalently,

a narrow Fourier spectrum (adiabatic pulses), it is

well known that the pulse envelope propagates in a transparent dispersive medium with the group vel-

ocity. The very first idea of this concept is due to Hamilton who distinguished, as far back as 1841 [8],

« the velocity of vibratory motion » from « the

Article published online by EDP Sciences and available at http://dx.doi.org/10.1051/jphys:01987004805079700

(3)

velocity of transmission of phase ». The group

velocity is now of standard use, not only in electro- magnetism but also in acoustics, quantum mechanics, etc. It is then quite surprising that such a

well-established concept may still be the subject of

discussions. The problem is that, in an anomalously dispersive medium, the group velocity may exceed the velocity of light in vacuum or become negative,

the latter condition being easily fulfilled in a resonant

absorber. For a long time, due to an incorrect reference to the causality and relativity principles, it

was considered that such velocities were without any

physical significance and could not represent signal propagation velocities. Indeed, it is clear that any identifiable peculiarity of the pulse (front, kink)

cannot propagate with a velocity larger than c [9, 7]

but this possibility may not be excluded for the overall shape of adiabatic pulses without marked front. Garrett and McCumber [10] have actually

shown that, under easily fulfilled conditions, a Gaussian pulse will propagate in a resonant absorber

at the group velocity, even when this velocity is negative. In this case, the peak of the output pulse anticipates the peak of the input field. This result have been later generalized to other pulse shapes [11-15] and is well supported by recent experiments involving laser pulses in a solid state sample [16],

millimetre wave pulses in a gas sample [17] or

ultrasonic pulses in superfluid 3He-B [14, 15]. Let us

note however that all the theoretical calculations are

approximate, being based on truncated series in the

frequency [10-12] or time [13] domain or on asympto- tic expansions of Fourier integrals [14, 15] (usually by the saddle point technique). They cannot then give a precise estimate of the pulse distortion.

This problem is not specific to the negative values

of the group velocity but is expected to be more

crucial in this pathologic case.

Thus one of the objectives of this paper is to fix an

upper bound to the pulse distortion, in order to give

a final confirmation of the physical relevance of the

negative group velocities. Its arrangement is as follows. In section 2, we first briefly recall, with slight extensions, the standard theory of group

velocity, apply it to the case of a resonant Lorentzian medium, and compare the approximate solutions given by this theory (adiabatic solution) with the

exact solutions obtained by computer simulations on

well-selected pulse shapes. The exact analytical

results are presented in section 3 : the correspond-

ence between the output and input pulse shapes is expressed in terms of well-conditioned operators, permitting us to give a rigourous upper bound to the

pulse distortion. This upper bound is explicitly

calculated in section 4, in the case of input pulse envelopes - d’/dt’[exp (- y 2 t2)] and of a pulse of

finite duration expanded in this basis. Higher order

corrections to the pulse envelope are given in

section 5 and section 6 is devoted to a summary of

our findings.

2. Group velocity and adiabatic solutions.

Let us consider a pulse of mean frequency wo

propagating in the z-direction in a linear medium. In the input plane (z = 0), its electric field may be written :

with

It is assumed in (2) that, the complex envelope EO(t) slowly varying at the scale of the mean period 2 vlcoo, its spectrum eo(u), centred on u = 0, is strictly null for I u I :> Cùo. The spectrum of

exp(icoot)Eo(t) then contains only positive fre- quencies w

=

coo + u, each of them propagating as exp [i ( Cù t - kz)]. The propagation constant k, de- pending on w, is complex for an absorbing medium (k = k1 - ik2 with kl, k2:> 0 for co :::. 0) but the absorption is usually weak on a wavelength path 2 w /ki (k 2kl, lkl kl). In any plane z, the

electric field becomes :

with ko

=

k(coo) and

Let us now consider input pulses, the envelope of

which is slowly varying at the scale not only of

2 7r/wo but also of any characteristic time of the medium (adiabatic pulses). Their spectrum eo(u) being very narrow, equation(4) approximately

amounts to :

Since any physical signal can be approximated as nearly as wanted by a function analytical in the complex plane, we may assume that EO(t) is analyti-

cal and, taking account of (2), write :

The result (6) extends the standard theory of group

velocity, allowed here to be complex [18], and, as we

shall see, is identical to the Oth order expansion of

the exact result according to a suitable adiabaticity

parameter (adiabatic solution).

(4)

Let us emphasize that, even at this order of

approximation, dramatic pulse reshaping may be present. This is easily seen by splitting the complex

time z dk/dw into its real part tl, which only entails

a time shift related to the real group velocity vg (with llvg

=

dklldw), and its imaginary part

t2 which will be generally responsible for both a frequency chirping and a distortion of the real

envelope of the pulse. An illustrative example is given by the camel-hump-shaped pulse

which is changed into the flat-topped pulse e(1 + y2t2)exp(- Y2 t2) for t2 = ± 1/)’ (Fig. 1).

An important exception is made by the Gaussian

pulses exp (- y2t2 ). The complex time delay only

affects the amplitude of their real envelope which

becomes

In the adiabatic limit, such pulses then propagate without shape change at the real group velocity vg whatever the frequency [10, 12, 13, 16]. Let us

note in passing that, in the two previous examples,

the peak value of I E (z, t ) I exceeds that of I Eo (t ) I .

This apparent anomaly is due to our definition of the

complex envelope, which does not include the

Fig. 1. - Pulse reshaping in the adiabatic limit. A time shift of ± i / y changes the camel-hump-shaped input pulse

of envelope y2t2 exp(- Y2 t2) (dashed line) into a flat- topped pulse of envelope _ (1 + y2t2 ) exp (- Y2 t2)

(dotted line). Note that the output amplitude is maximum

when the input one is null. In the case of a Lorentzian medium, the required imaginary time shift is achieved for

a frequency detuning equal to the half width at half maximum (HWHM) of the line (1/T2) and an optical depth 1/yT2. The full line relates to the exact output shape, calculated in such a case ( y T2 = 1/8, optical depth Z/2

=

8). Only a weak reminiscence of the central well its observed. The three curves are normalized. The time unit is 1/y.

overall complex attenuation exp (- i ko z ) (see

Eq.(3)).

.

For adiabatic pulses of arbitrary shape, the distor- sion will be only negligible when the medium

absorption coefficient is negligible (k2 == 0) or pre-

sents an extremum with respect to the frequency (dk2/dw = 0). The most interesting features are expected in a resonant medium when the group

velocity becomes negative. To be definite, we will consider an idealized Lorentzian medium [1, 10-17].

Assuming that the region of anomalous dispersion is

narrow compared to the resonance frequency wm, such a medium is characterized by the approximate dispersion law :

where no is the frequency-independent refractive

index of an eventual host medium and 1 / T2 is the

width (HWHM) of the resonance line. Equation (7) only holds when the resonance absorption coefficient

« is small compared to the inverse of the wavelength

in the host medium [13]. This latter condition is in fact necessary, since otherwise any resonant adiaba- tic pulse would be fully absorbed after a few

wavelengths and no experiment would be feasible.

As expected, the group velocity, derived from

equation (7), is real and superluminal (vg :::- c/no ) on

exact resonance and becomes negative as soon as

a :::. 2 no T2/ c, a condition easily fulfilled by various

absorbers in different frequency domains [14-17] and quite consistent with the previous one.

Introducing the dimensionless quantities 0 (local

or retarded time), Z (optical thickness on resonance)

and 5 (mean detuning) such as :

Equation (6) becomes :

The real electric field & (z, t ) may also be written (see Eq. (3))

where E is the modified pulse envelope including the

overall complex attenuation :

On exact resonance, we get :

In the adiabatic limit, the pulse attenuation is

identical to the cw one and the output pulse shape

anticipates the input with the true time-advance

(5)

ZT2 (more exactly ZT2 - no z/c but the transit time

no z/c is usually negligible compared to ZT2).

Important advances, comparable to the pulse width

Tp, appear to be possible within the adiabatic limit,

that is without significant pulse distortion. Selecting pulses without frequency chirping (Eo(t) real), we

have verified this point by comparing, in such conditions, the adiabatic solution EO(O + Z) with

the exact solution E(Z, (), numerically computed

from the dispersion law (7) by a standard FFT procedure. The first example (Fig. 2), relating to the camel-hump-shaped-pulse, already considered, has

been chosen in order to show that the advance

phenomenon is not a specific attribute of Gaussian

or bell-shaped pulses, but affects any adiabatic

pulse. Let us note in passing the invalidity of the Crisp’s claim [11] that the output field has to be weaker than the input one at the same retarded time

0. Indeed, this rule clearly fails for 0 == 0 in the present case.

Fig. 2.

-

Propagation of a camel-hump-shaped pulse in a

resonant Lorentzian medium (s

=

T2/ Tp

=

1/32 ;

Z = 16). As expected, the output pulse envelope (full line) anticipates of ZT2

=

Tp/2 the input one (dashed line). The dotted line gives the difference between the former and the adiabatic solution. The time unit is the half-width of the central well in the input pulse envelope.

When the input pulse envelope is almost everywhere adiabatic but presents singularities, e.g.

a marked front, these singularities will be responsible

of the appearance of oscillatory patterns superim- posed on the slowly varying part of the output pulse shape. These patterns are identified or clearly related [15] to the high frequency precursors of Sommerfeld and Brillouin [1] and their wavefront propagates at

the luminal velocity c/no [9]. This is shown figure 3

in the case of a triangle-shaped symmetric pulse. As expected, no output signal is observed in the region

forbidden by the causality principle and « wiggles »

Fig. 3.

-

Propagation of a triangle-shaped pulse in a,

resonant Lorentzian medium. Excepted for the vicinity of

the slope discontinuities, the output pulse envelope (full line) is not distinguishable from the adiabatic solution

(dotted line). The time unit is the width (HWHM) of the input pulse envelope (dashed line).

develop just after the discontinuities of the slope of

the input pulse envelope. Outside these regions, the output pulse shape exactly agrees with the predic-

tions of the adiabatic approximation. All these

features are well explained by remarking that the medium response to Eo ( 9 ) is the time-integral of its

response to dEo/dO, which is the sum of three step functions of relative amplitude + 1, - 2 and + 1. It is

well known [19-21] that the medium-response to a single unit step strongly oscillates before reaching

the cw value e- z. After integration on 0, the transient oscillation and the steady state response will naturally lead respectively to the wiggles and the

linear parts of the output pulse shape.

The relative amplitude of the wiggles, compared

to the « adiabatic » part of the output pulse, obvious ly depends on the optical thickness Z and on the

nature of the singularities of the input pulse en- velope. Without entering into a detailed discussion which may be found elsewhere (see e.g. [15]), this is qualitatively interpreted by remarking that the wig- gles originate from the far wings of the spectrum e(u) where the approximation k(wo + u) - k(wo) + u dk/dw fails. It is well known [22] that the spectrum le(u)1 ] of a signal EO(t), the nth derivative of which presents a discontinuity, behaves asymptoti- cally as u- (n +1)*For a given optical thickness Z, the amplitude of the wiggles are then expected to

decrease with the order n of the discontinuity.

Moreover, for a given n, the part of the spectrum e(u) involved by the wiggles, being partially outside

the absorption band of the medium, the wiggles are

predicted to be less attenuated than the adiabatic

part of the pulse. As an illustrative example, figure 4

gives the output pulse shapes numerically computed

(6)

Fig. 4.

-

Study of the wiggle.s resulting from a discontinuity of the nth derivative of the input pulse envelope. (a) Input pulses. We have chosen pulses having the same amplitude and HWHM (taken as time unit), differing only by the order n of the discontinuity at their beginning and their end (n

=

3, 4, 5 ). The pulse pedestal lengthens very slightly with n but

this lengthening is negligible as soon as n -- 4. (b) (c) (d) (e) (f) Output pulse envelopes after propagation of the pulses (a) in a resonant Lorentzian medium (s

=

T2/ T P = 1/16), for different optical depths Z. The amplitude of the wiggles,

B

rapidly decreases (resp. increases) with n (resp. Z). Note that the maximum of the « adiabatic » part of the pulses

anticipates that of the input pulse of about ZT2, as expected in the adiabatic limit.

(7)

for different values of Z and input pulses, differing only by the order n of their discontinuities, but having the same width Tp (HWHM) and nearly

identical envelopes :

with

The dependence of the relative amplitude of the wiggles on Z and n agrees with the previous qualita-

tive predictions, as also with our experimental

observations [17]. Let us note in particular that the

conditions on the continuity of EO(t), to fulfil in

order to keep the wiggles at a moderate level,

become more and more severe as the optical thick-

ness increases.

3. Exact analytical expressions.

An exact expression of the output signal is given by

the Fourier integral (4), which leads to the concept of group velocity when k(wo + u) is expanded in a

power series, limited to its first order term - u. In the case of a resonant Lorentzian medium, it is easily

shown that the second order term - u 2 simply entails

the narrowing of any bell-shaped pulse [11, 23, 13],

the pulse maximum moving along with the group

velocity vg. Expansions of k(coo + u ) up to the third order term ~ u3 have only be considered when the term ~ u 2 vanishes, that is when the group velocity presents a minimum [18]. Indeed, it is obvious that the calculations rapidly become untractable as the order increases. More distressing, when the calcula- tions are made at a given order, e.g. at the 1st order

leading to the adiabatic solution discussed before, it

seems quite difficult to obtain an estimate of the

corresponding error and even to show the existence of an upper bound to this error.

It appears then more advisable to examine the

problem. in the time-domain [24, 13-15]. This is

achieved by transforming the frequency integral (4)

into the corresponding time-integral, namely the

convolution of the input field with the impulse

response of the medium [25]. The latter can be calculated in the case of a Lorentzian medium as

follows. Using the dimensionless quantities

d2 = (w - too) T2, 8, Z and 0 (see Eq. (8)), we

derive easily, from the dispersion law (7) and the

definition (10), the envelope E(Z, 0) corresponding

to the plane wave exp[i(wt - kz)] :

with

The impulse response F(0) of the medium is simply

the inverse Laplace transform L - 1 {f (s )} of the

system function f(s) [25]. Using the relation [26] :

and standard procedures of Laplace transform [26],

we get :

I - I 1

where U+ is the unit step function and Jo the Oth

order Bessel function. The modified envelope of the

output pulse, for an arbitrary input pulse envelope Ego (0 ), is then given by the convolution Eo (0 ) 0 F(0) which may be written:

with

Similar results have been obtained by other techni- ques [24, 13-15]. As expected, the expressions (18)

and (19) characterize a strictly causal system [8, 9, 25] (F (0 ) = 0 for 0 -- 0).

Equations (19) and (20) can be brought to tract-

able forms in particular cases, especially in the

adiabatic limit, and the adiabatic solution, discussed in section 2, is then retrieved by an asymptotic expansion of the integrals, evaluated by the saddle point method [15]. Unfortunately, like the technique

based on a series expansion of k(w ), such methods do not permit an estimate of the involved errors.

From another point of view, equation (19) may be

seen as an operator correspondence relating the output and input pulse envelopes. This integral form

is easily shown to be equivalent to the differential form:

with the boundary conditions :

Equations (21) and (22) can also be directly

derived from the dispersion law (7) or from the

linearized Bloch-Maxwell equations of a two-level

medium [13], within the slowly varying amplitude

and phase approximation [27].

Equation (21) admits an infinity of steady state

solutions E(Z, 0 ) = f (Z) Eo [ 0 + g(Z)]. They are

such as :

with, according to equation (22), f(0) = 1,

g(0)

=

1. The corresponding input pulse envelopes

(8)

Eo (0 ) are all of unbounded amplitude and, strictly speaking, no pulse of finite energy can propagate without distortion, as the well-known hyperbolic

secant pulses in the related nonlinear problem [28, 29]. Note that, for /(Z)=exp(-Z/l+f6) and g (Z)

=

ZI (1 + 1 6 f, Eo(8), given by equation (23),

reduces to a8 + b. In this remarkable case, the exact solution of equation (21) is rigorously equal to the

adiabatic solution given by equations (9) and (11), as already mentioned (see Fig. 3).

Equation (21) also implies a general property of the pulse energy, relating to the causality principle.

From (21) and the complex conjugate equation, we

get easily :

Whatever the sign of the group velocity, the energy propagates in the z-direction and equation (24)

shows that the total energy, having crossed the

output plane at any retarded time 0

is always weaker than the energy having left the input plane at the same retarded time. This novel rule of « energetical causality » replaces the Crisp’s

rule [11] ( ( E (Z, e ) I ) , I E (o, 9 ) I which fails in

some cases (see Sect. 2) and complements the rule of

strict causality.

As indicated before, the differential equation (21)

is convenient to establish an operator correspond-

ence between the output and input fields, more

suitable than that given by the integral equation (19).

We summarize here the demonstration which is detailed in appendix A. We search first for solutions of equation (21) in the form of a series V bn(9 ) Zn

and by summing the resulting series, we get:

where A is the linear operator such as :

Equations (25) is then transformed by exploiting the properties of A, in particular the fact that A com- mutes with the time-derivative operator a

=

a / a e.

Introducing the reduced complex envelop E(Z, 0),

that is f (Z, 0 ) cleared of the overall complex

attenuation exp( - Z/1 + is) (see Eq. (11», we

get :

The adiabatic solution (9) is immediately identified

to the exact one (27) if the exponential operator is approximated by 1 (identity operator). The error

involved in this approximation may be characterized

by the uniform norm AE of the difference of the exact solution and of the adiabatic one.

An upper bound to AE is easily obtained by expanding in series the exponential operator of equation (27) and remarking that:

Scaling the time with a time Tp characterizing the input pulse envelope, instead of T2, we get:

with

and where a henceforth indicates the derivative

according to T. s is an adiabaticity parameter, the smaller the less the input field envelope varies during T2. The adiabatic solution reads

and appears as a Oth order solution with regard to

s2Z/(l + S2). For a given input pulse, the validity

of the adiabatic approximation depends then on the optical depth ZI(l + 8 2) as expected. Moreover, examining equation (29), it seems reasonable to

think that AE can be kept at a very low level if

s2 Z/ (1 + 52)_ , 1. A significant pulse advance, com- parable to rp, is achieved if sZ/ (1 + 52) _ 1. Both

conditions are consistent if s 1 but a more explicit

calculation of AE is however required.

4. Upper bound to the pulse distortion : explicit

calculations.

For an arbitrary detuning between the mean fre- quency wo of the pulse and the resonance frequency

w m, the pulse distortion originates both from the

complex value of the time delay involved in the

adiabatic approximation (see Sect. 2 and, e.g.,

Fig. 1) and from the terms neglected within this approximation. In this section, we focus our atten- tion to the latter source of distortion by examining

the case of exact resonance, where the adiabatic

approximation predicts that the envelope of the

output pulse anticipates that of the input pulse

(9)

without any distortion. Equations (31) and (29) then

reduce to :

Our objective is to show that a significant pulse advance, say of the order of the pulse duration (i.e.

sZ =-= 1), is compatible with a low distortion. This

requires that an upper bound is fixed to the pulse distortion, not only for a particular envelope Eo( 7" ) but for a set of envelopes providing a basis to expand any adiabatic envelope.

An explicit calculation is actually feasible for the basis provided by the Gaussian envelope exp(- T 2)

and its successive derivatives. The time-unit is here the half-width at lle of the Gaussian pulse. Using

the properties of the Hermite polynomials Hn and of

the factorial function r [30-32], we get :

An upper bound to the distortion AE p 2 of the envelope of a pulse such as Eo (T )

=

aP e -T (p>_O)

is then easily derived from (33).

We obtain :

with A

=

4 s2 Z. This series converges as soon as A 1 and explicit sums or upper bounds are given in appendix B. It is obviously necessary that A « 1 in order that the pulse distortion may be low. An

equivalent to the series is then given by its first

term :

This quantity has to be compared to the peak amplitude Ep = 11 9P e- r2 110, of the corresponding input pulse in order to precise the relative distortion

åEp/Ep. An estimate of EP is given by (34) and, taking account of the definitions (30), we get :

Recall that in this formula rp is the half-width at

1/e of the basic Gaussian pulse (p

=

0). This time is

not quite suitable to characterize the pulses corre- sponding to higher order derivatives 9Pe- 2 which

oscillate faster and faster as p increases (p -- 2). It is

more advisable to characterize these pulses by a time proportional to the pseudo-period of their fastest oscillations. This pseudo-period roughly depends on

p as (p + 1/2)-1z [32], that is nearly as (p + 1)- 1/2 (p -- 2). Taking then T pi N/P + 1 as the new defini-

tion de Tp, approximately suitable whatever p, (37)

reads :

As expected, for a given pulse advance, comparable

to the characteristic time of the pulse, that is for sZ

=

T2 Z/Tp =1, the distortion can be kept at a

level as low as wanted by taking s

=

T2/ T P small

enough.

We have compared, in various cases, the true

distortion with the upper bound given by (35).

Figure 5 shows the output pulse envelope obtained

for a Gaussian input pulse. The optical depth (Z 8 and the adiabatic parameter (s

=

B/Log 2/Z) have been chosen in order that the pulse

advance is equal to the width (HWHM) of the input pulse. and the distortion is moderate but significant (22 %). This distortion is very close to the upper- bound given by (35) (24 %). A second example is given by the normalized camel-hump-shaped pulse

of figure 2 (dashed line) :

An upper bound to the pulse distortion is then

e(,&EO12 + åE2/4) that is about 18 °7o with the

Fig. 5.

-

Checking of the upper bound to the pulse

distortion and of the higher order corrections. The dashed and full lines are respectively the envelopes of a Gaussian input pulse and of the corresponding output pulse for

Z

=

8, s

=

0.104. The time unit is the HWHM of the input pulse. The circles (0) relates to the pulse distortion, i.e.

the difference between the output pulse envelope and the input one, advanced of ZT2 (adiabatic solution). Its peak

value (24 %) is very close to its theoretical upper bound

(22 %). The dotted line gives the correction to the adiabatic solution derived from the expansion (45). As expected, the main effect of this correction is a narrowing

of the output pulse.

(10)

parameters of figure 2 (s = 1/32 ; Z = 16 ; À

=

1/16 and then AEo = 0.03, AE2

=

0.2). This value is

again not far beyond the true distortion (=..e 10 %).

This situation seems to be general when the distor- tion is moderate.

The above procedure is easily extended to any normalized pulse, of finite energy, which can be’

expanded in the basis provided by e- T2 and its

successive derivatives :

An upper bound to the pulse distortion is then :

where the AEp are given by equation (35). In general, the moduli lap I decrease vs. p more rapidly

than p1l4/2P/2 JP! [30] but it is not sufficient to

prevent the series (41) to diverge. This is not surprising since a pulse of arbitrary envelope is always heavily distorted. On the contrary, if the

input pulse envelope is such that the I ap ( decrease

vs. p as (or faster than) XPIF p + with

jc 1/2, the series (41) converges to a value small

compared to unity in the adiabatic limit (A « I ) and

the corresponding adiabatic pulses can propagate by taking a significant advance with negligible distor- tion, as the related Gaussian pulses [11].

The precise mathematical characterization of these

strictly adiabatic pulses is beyond the scope of this paper. Moreover the previous definition seems to be too restrictive : the convergence of the series (41) is

a sufficient condition but is not a necessary one since it results from successive overestimates. The concept of adiabatic pulse may be heuristically extended to

any pulse of analytical envelope [13]. To such a pulse, it is possible to associate a scaling time

Tp, related to the convergence of the series expansion

of EO(t) in any point:

(recall that Eo(t) is normalized). The adiabaticity

condition A « 1 reads then :

Numerical simulations, made with various pulse envelopes, have shown that the pulse distortion is

always negligible if the condition (43) is fulfilled. For

an ideally smooth bell-shaped pulse, the characteris- tic time Tp, given by equation (42), does not depart significantly from the pulse half-duration (HWHM).

On the contrary, if any derivative d’Eoldtn presents

a discontinuity (often hardly visible on the input

pulse shape), rp becomes null, the adiabatic approxi-

mation breaks down and the singularity leads to a wiggle propagating at the velocity c, in accordance with the causality principle. This fact is clearly

evidenced by the simulations of figures 3 and 4, and by the experimental signal of figure 2 in re-

ference [17].

From an experimental point of view, the envelop-

pe EO(t) of a realistic pulse, having a finite duration,

is obviously not analytical. It can however be

approximated as nearly as wanted by an analytical

function EOA(T) and the two envelopes EO(t) and EOA(T) are equivalent if their difference eo(t) leads to

an output signal E (z, t ) below the tolerated distor- tion. A general upper bound to c (z, t ) is given in

reference [33]. Concretely, if the discontinuities of the true pulse envelope are of sufficiently large order

n, they lead to wiggles of negligible amplitude and

the previous results apply by restricting n to

n - 1 in equation (42). The order n depends on the optical depth which is experimentally limited by the sensitivity of the output detection (note that an optical depth Z = 12 leads to a pulse attenuation

larger than 100 dB !). The simulations of figure 4b

resp. d and f) show that a discontinuity of order 5 (resp. 4 and 3) raises a wiggle of amplitude weaker

than the distortion of the adiabatic part of the pulse

for Z = 16 (resp. 12 and 8). A more quantitative

discussion on the amplitude of the wiggles can be

found in reference [15].

5. Higher order corrections to the pulse shape.

Up to now we have only considered the adiabatic

solution, that is the Oth order solution with respect to the generalized adiabatic parameter A (A

=

4 s 2 Z ),

and we have fixed an upper bound to the error

involved by this approximation. In fact our operator formalism is quite efficient to derive the higher

corrections to the pulse shapes. This is achieved by using repetitively the procedures having led to equation (27). We get (see Appendix A) :

that is

where s’

=

sl(l + i 8 ) and Z’

=

Z/ (1 + i5) reduce

to s and Z on exact resonance and where Pn(Z’) are polynomials such as :

of degree n/2 (resp. (n - 1 )/2) for n even (resp.

odd). Equations (45) and (46) generalize the result

previously obtained in the resonant case by an other

(11)

technique [13]. It is easily shown that the polyno-

mials P,,(x) fulfil the recurrence relation (for n :::. 1) :

From equations (46) or (47) we get easily :

As expected the lowest order correction in s to the

pulse shape is the 2d order one. At this order

equation (45) reads :

When the input pulse is resonant and bell-shaped of

real and symmetrical envelope, the correction given by equation (49) entails only a compression of the pulse, its maximum still propagating with the nega- tive group velocity [11, 23, 13]. Obviously the higher

order corrections involve more substantial distortion.

It is interesting to test the expansion (45) in the

situation where the pulse advance is comparable to

the pulse width, that is for sZ =-- 1. In order to obtain

a result valid at the pth order in s, it is then necessary to use an expansion (45) up to the order 2 p (resp.

2 p -1) for p even (resp. odd). Such a calculation has been made in the case of the Gaussian pulse of figure 5. The correction, calculated with p

=

4, perfectly reproduces the difference between the adiabatic solution and the exact one. This supports the exactness of the expansion (45) and shows that the correction proportional to aEo (,r + s’ Z’) ap-

pearing in the approximate calculation of re-

ference [15] is irrelevant.

6. Conclusion.

To summarize, the problem of the propagation in a

linear resonant absorber of electromagnetic pulses

of very slowly varying envelope has been revisited.

The medium is assumed to be Lorentzian and such that the group velocity is largely negative on re-

sonance. It is shown that, for so-called adiabatic

pulses, the shape of the output pulse can actually anticipate that of the input one with a distortion as

low as wanted. This gives a final support to the idea that the group velocity is a meaningful physical concept, even when it becomes negative. This result has been previously stated by Garret and McCumber for Gaussian pulses [10] but the existence of an

upper bound to the pulse distortion was not proved.

This is achieved here analytically by an operator

technique, in the case of Gaussian and Gaussian-

derivative pulses, and of a particular class of pulses expanded in the basis provided by the previous ones.

This formalism is also shown to be quite efficient to

derive the corrections of higher order to the pulse shape.

Numerical simulations show that the negative- velocity propagation can be observed on a rather

wide class of smooth pulses, the envelope of which

evolves slowly at the scale of T2/ JZ. Our results are

compared with those obtained by more standard procedures and some general results concerning the

linear pulse propagation are given in the course of

the paper.

Acknowledgments.

One of us (BM) would like to thank E. Varoquaux

for an enjoying and fruitful discussion.

Note added in proof. - In a very recent paper

(Phys. Rev. A 34 (1986) 4851), published after the

submission of the present paper, Tanaka et al. give

some results on the propagation of Gaussian pulses

which might seem conflicting with our ones. Without entering here into a detailed discussion, we wish to emphasize that the computer simulations of Tanaka et al. have been made in the case of a heavily absorbing medium (a == 8/A on resonance) and of

ultrashort pulses (Tp = 20 periods of the carrier and

comparable to T2). In such conditions the adiabatic

approximation is obviously invalid and it is then not

suprising that the pulse propagation is not described by the group velocity.

Appendix A.

We search for solutions of (21) in the form of a series

expansion :

Substituting this form into equations (21) and (22),

we get :

where A is a linear operator such that

Equation (A.I) reads then :

Derivating equation (A.3) with respect to 0 and

integrating it per parts, we get respectively :

(12)

and

Using equation (A.7), equation (A.4) is immediately

transformed into :

that is

where £(Z, ()) is the complex envelope cleared of

the overall complex attenuation exp - 1 + i 5 ( l + 1 6 )

Using again equation (A.7) and taking equation (A.8) into account, we get then :

Noting that

we obtain immediately the result given by equation (27) in the main text.

The technique used to derive equation (A.11)

from equation (A. 10) can obviously be continued in

order to determine the higher order corrections.

Exploiting equations (A.7) and (A.8) iteratively we get easily :

Insofar as this operator expansion applied to EO(O) converges, (A.13) may be written :

and the substitution of this result in (A.11) simply

leads to (44) of the main text.

Appendix B.

The series

appearing in (35) converges for A 1. It can be calculated (resp. bounded) for p odd (resp. even) by using the expression of the gamma function for

integer (resp. half-integer) values of the variable [31]

We get in particular

Similarly for any odd value of p (p

=

2 k + 1 with

k > 0), we obtain :

Observing that

I" ."J

and that

we get:

Note that this result includes in fact the previous

one, obtained for k

=

0 (Eq. (B.4)).

The situation is slightly more complicated when p

is even. The series S2 k are however easily bounded.

For k

=

0 (Gaussian pulse), the general term of the

series is

and

Similarly, for any positive value of k, we get

and, using again (B.6) and (B.7),

This equation, proved for k > 0, recovers also the

case k = 0 (Eq. (B.10)).

As checked on several examples, the upper bound

given by these equations is in fact close to thb exact

sum of the series.

(13)

References

[1] BRILLOUIN, L., Wave Propagation and Group Veloci- ty (Academic Press, New York) 1960.

[2] BORN, M. and WOLF, E., Principles of Optics (Pergamon, Oxford) 1975, p. 23.

[3] JACKSON, J. D., Classical Electrodynamics (Wiley,

New York) 1962, p. 211.

[4] SOMMERFELD, A., Optics (Academic Press, New York) 1954.

[5] VAINSHTEIN, L. A., Sov. Phys. Usp. 19 (1976) 189.

[6] LIGHTHILL, M. J., J. Inst. Math. Appl. 1 (1965) 1.

[7] PRASAD, P., J. Math. Phys. Sci. 14 (1980) 71.

[8] HAMILTON, W. R., Proc. Irish Acad. 1 (1841) 341.

[9] SOMMERFELD, A., Physik. Z. 8 (1907) 841.

[10] GARRETT, C. G. B. and MCCUMBER, D. E., Phys.

Rev. A 1 (1970) 305.

[11] CRISP, M. D., Phys. Rev. A 4 (1971) 2104.

[12] PURI, A. and BIRMAN, J. L., Phys. Rev. A 27 (1983)

1044.

[13] MACKE, B., Optics Commun. 49 (1984) 307.

[14] AVENEL, O., VAROQUAUX, E. and WILLIAMS, G. A., Proc. 17th Int. Conf. Low Temperature Physics, Ed. Eckern V., Schmid A., Weber W.

and Wühl H. (North-Holland, Amsterdam) 1984, p. 767.

[15] VAROQUAUX, E., WILLIAMS, G. A. and AVENEL, O., Phys. Rev. B 34 (1986) 7617.

[16] CHU, S. and WONG, S., Phys. Rev. Lett. 48 (1982)

738.

[17] SEGARD, B. and MACKE, B., Phys. Lett. 109A (1985)

213.

[18] JOHNSON, D. L., Phys. Rev. Lett. 41 (1978) 417.

[19] SHIREN, N. S., Phys. Rev. Lett. 15 (1965) 341 and

397.

[20] CRISP, M. D., Phys. Rev. A 1 (1970) 1604.

[21] MACKE, B. and ROHART, F., Optica Acta 28 (1981)

1135.

[22] ARSAC, J., Transformation de Fourier et théorie des distributions (Dunod, Paris) 1961, p. 46.

[23] CHU, S. and WONG, S., Phys. Rev. Lett. 49 (1982)

1293.

[24] HARTMANN, H. F. and LAUBEREAU, A., Optics

Commun. 47 (1983) 117.

[25] PAPOULIS, A., The Fourier integral and its applica-

tions (McGraw Hill, New York) 1962.

[26] ABRAMOWITZ, M. and STEGUN, I. A., Handbook of

Mathematical Functions (Dover, New York) 1970, p. 1020.

[27] TANG, C. L. and SILVERMANN, B. D., Physics of Quantum Electronics, Ed. Kelly P. L., Lax B.

and Tannenwald P. E. (McGraw Hill, New York) 1966, p. 280.

[28] MCCALL, S. L. and HAHN, E. L., Phys. Rev. Lett. 18 (1967) 908.

[29] ALLEN, L. and EBERLY, J. H., Optical Resonance

and two-level atoms (Wiley, New York) 1975, p. 78.

[30] SZEGÖ, G., Orthogonal Polynomials (American

Mathematical Society, Colloquium Publication) 1978, Vol. 23.

[31] Reference [26], p. 255.

[32] Reference [26], p. 787 and p. 924.

[33] MACKE, B., ZEMMOURI, J. and SEGARD, B., Optics

Commun., 59 (1986) 4851.

Références

Documents relatifs

Abstract: We prove a new upper bound for the first eigenvalue of the Dirac operator of a compact hypersurface in any Riemannian spin manifold carrying a non-trivial twistor

In this paper, we propose a new encoding scheme which is used in a hybrid resolution: a metaheuristic is matched with a list algorithm (Next Fit, First Fit, Best Fit) to solve the

Au total, la prévalence des troubles liés à l’abus d’alcool est fréquente dans le trouble bipolaire, et plus élevée chez les patients consultant pour leur trouble que

Les valeurs de gain et de facteur de bruit pour plusieurs valeurs de puissance d’entrée du signal dans l’amplificateur à fibre dopée erbium faiblement multimode

L’archive ouverte pluridisciplinaire HAL, est destinée au dépôt et à la diffusion de documents scientifiques de niveau recherche, publiés ou non, émanant des

Representation of the obtained failure mechanism u for the square plate problem with simple supports (left) and clamped supports (right) using the P 2 Lagrange element.

It is the purpose of this manuscript to formulate a numerical model that is capable of describing dynamic shear band propagation in a porous medium, with a solid skeleton and

In conclusion, using two kinds of anti-βIV spectrin antibodies, NT for βIVΣ1 and SD for both βIVΣ1 and βIVΣ6 splice variants, we found a switch in the expression levels