Universality class of the glassy random laser Jacopo Niedda1 2Giacomo Gradenigo3 4 2Luca Leuzzi2 1and Giorgio Parisi1 2 5 6 1Dipartimento di Fisica Universit a di Roma Sapienza Piazzale A. Moro 2 I-00185 Roma Italy

2025-05-06 0 0 1.97MB 17 页 10玖币
侵权投诉
Universality class of the glassy random laser
Jacopo Niedda,1, 2 Giacomo Gradenigo,3, 4, 2 Luca Leuzzi,2, 1, and Giorgio Parisi1, 2, 5, 6
1Dipartimento di Fisica, Universit`a di Roma “Sapienza”, Piazzale A. Moro 2, I-00185, Roma, Italy
2NANOTEC CNR, Soft and Living Matter Lab, Roma, Piazzale A. Moro 2, I-00185, Roma, Italy
3Gran Sasso Science Institute, Viale F. Crispi 7, 67100 L?Aquila, Italy
4INFN-Laboratori Nazionali del Gran Sasso, Via G. Acitelli 22, 67100 Assergi (AQ), Italy
5INFN, Sezione di Roma-1, P.le A. Moro 5, 00185, Rome, Italy
6Accademia Nazionale dei Lincei, Palazzo Corsini - Via della Lungara, 10, I-00165, Roma, Italy
By means of enhanced Monte Carlo numerical simulations parallelized on GPUs we study the
critical properties of the spin-glass-like model for the mode-locked glassy random laser, a 4-spin
model with complex spins with a global spherical constraint and quenched random interactions.
Implementing two different boundary conditions for the mode frequencies we identify the critical
points and the critical indices of the random lasing phase transition with finite size scaling techniques.
The outcome of the scaling analysis is that the mode-locked random laser universality class is
compatible with a mean-field one, though different from the mean-field class of the Random Energy
Model and of the glassy random laser in the narrow band approximation, that is, the fully connected
version of the present model. The low temperature (high pumping) phase is finally characterized
by means of the overlap distribution and evidence for the onset of replica symmetry breaking in the
lasing regime is provided.
I. INTRODUCTION
When light propagates through a random medium,
scattering reduces information about whatever lies across
the medium and the electromagnetic field, composed by
many interfering wave modes, provides a complicated
emission pattern as light undergoes multiple scattering.
If enough power is pumped into the medium multiple
scattering may support the population inversion of atoms
and molecules above some optical gap, yielding a random
laser [1–14]. Random lasers are made of an optically ac-
tive medium and randomly placed scatterers (sometimes
both in one [2]). The first provides the gain, the lat-
ter provides the high refraction index and the feedback
mechanism needed to lead to amplification by stimulated
emission. As opposed to ordered standard multimode
lasers, random lasers do not require complicated con-
struction and rigid optical alignment, have a low cost,
undirectional emissions, high operational flexibility and
give rise to a number of promising applications in the
field of speckle-free imaging [15, 16], granular matter
[9, 10], remote sensing [13, 17, 18], medical diagnostics
and biomedical imaging [13, 19–22], optical amplification
and optoelectronic devices [13, 23, 24].
Random lasers may show multiple sub-nanometer
spectral peaks above a pump threshold [2], as well as
smoother, though always disordered, emission spectra.
Depending on the material, and its optical and scattering
properties, random spectral fluctuations between differ-
ent pumping shots (i. e., different realizations of the same
random laser) may or may not vary significantly. A wide
variety of spectral features is reported [10, 12, 25–27],
depending on material compounds and experimental se-
tups. Random lasers can be built in very different ways,
luca.leuzzi@cnr.it
can be both solid or liquid, can be 2D or 3D, the optically
active material can be confined or spread all over the vol-
ume. Moreover, random lasers are, usually, open systems
where light can propagate in any direction rather than os-
cillating between well specific boundaries (mirrors) as in
standard lasers and the emission acquisition can only be
directional rather than on the whole solid angle. Finally,
also the scattering strength and the pumping conditions
may affect the emission.
In the last years experiments on a certain class of ran-
dom lasers provided evidence of particularly non-trivial
correlations between the shot-to-shot fluctuations of the
emission spectra. We will refer to those as glassy ran-
dom lasers [13, 28–32]. These special correlations are
predicted by a theory based on statistical mechanics of
complex disordered systems [33–35]. Indeed, it has been
shown that these fluctuations are compatible with an or-
ganization of mode configurations in clusters of states,
similar to the one occurring for complex disordered sys-
tems displaying multiequilibria, as the spin glasses. Such
a correspondence has been analytically explained prov-
ing the equivalence between the distribution of the Inten-
sity Fluctuation Overlaps (IFO) and the distribution of
the overlap between states, the so-called Parisi overlap,
the order parameter of the glass transition [36]. Though
the analytical proof assumes narrow-band spectra, such
that all modes - within their line widths - can be consid-
ered at the same frequency [33, 37, 38], numerical simu-
lations have provided evidence that the onset of nontriv-
ial distributions of IFO and Parisi overlap distributions
occur at the same (critical) temperature also in realis-
tic models for random multimode lasers [39]. In these
models, the four-waves non-linear mixing between elec-
tromagnetic field modes is controlled by a deterministic
selection rule depending on modes frequencies, termed
mode-locking. In mode-locked lasers interactions are pos-
sible only for the quadruplets of modes whose frequencies
arXiv:2210.04362v3 [cond-mat.dis-nn] 24 Feb 2023
2
ωksatisfy the condition
|ωk1ωk2+ωk3ωk4|< γ, (1)
with γbeing the typical line-width of the modes. We
will refer to Eq. (1) as Frequency Matching Condition
(FMC). In standard mode-locked lasers such selection
rule is implemented by ad hoc nonlinear devices (e.g., sat-
urable absorbers for passive mode-locking [40]) that are
not there in random lasers. As hypothesized in [41] and
recently experimentally demonstrated in [12], though, in
random lasers mode-locking occurs as a self-starting phe-
nomenon. We call an interaction network built on the
mode-locking selection rule in Eq. (1) a Mode-Locked
(ML) graph.
From the point of view of statistical mechanics of com-
plex disordered systems, random lasers represent, so far,
the only physical system where the relevant degrees of
freedom, namely the complex amplitudes of the light
modes, naturally form a dense interaction network of
the kind for which replica symmetry breaking mean-field
theory [42] is proved to work, as in high dimension spin-
glasses or structural glasses made of hard spheres [43].
It is not by chance that random lasers are, so far, the
only complex disordered system providing experimental
evidence of a continuous replica symmetry-breaking pat-
tern [13, 28–32]. Actually, mean-field theory for an infi-
nite number of replica symmetry breakings has rigorously
been derived [44, 45] only for fully connected systems, in-
cluding the random laser model in the narrow-band ap-
proximation [33, 38]. Using the cavity method it is, then,
possible to compute a replica symmetry breaking (RSB)
phase also in systems with sparse interactions[46] (the
Viana-Bray model, for instance [47–49]). Still, the cor-
rect mean-field theory which describes ML random lasers
has yet to be found, due to some peculiarities of the in-
teraction between light modes that will be detailed in the
following.
In this work we resort to Monte Carlo numerical simu-
lations of the dynamics of a leading model for multimode
random lasers, the Mode-Locked (ML) 4-phasor model
[33, 34, 38, 50].
Even if the phenomenology of the model is quite rich
already in the narrow bandwidth approximation, going
beyond the fully-connected case is necessary to achieve
a realistic description of random lasers in the spin-glass
theoretical framework. If Nis the number of modes, the
FMC leads to O(N) dilution in the interaction graph: the
total number of interactions, which is of order O(N4) in
the complete graph, is, thus, reduced to O(N3) in the
diluted graph [51].
Therefore, as far as the the interaction graph is con-
cerned, the ML 4-phasor model places itself in an inter-
mediate position between the complete and the sparse
graph, the latter being the case where the number of
couplings per variable does not scale with Nin the ther-
modynamic limit. The analytical solution of a spin-glass
model in such an intermediate regime of dilution is a
very hard problem to address, since standard mean-field
techniques such as RSB theory, [52, 53], do not straight-
forwardly apply and the cavity method for sparse [48] or
diluted dense networks [54] does not allow to devise close
equations for global order parameters and provide a fully
explicit solution. Eventually, to the best of our knowl-
edge, no spin-glass model has been solved exactly out of
the fully connected or the sparse case. Hence, one needs
to perform numerical simulations in order to investigate
the physics of the model.
The ordered version of the ML 4-phasor model has
been extensively studied through numerical simulations
in [55, 56], where the essential consequences of the FMC
on the topology of the interaction graph have been inves-
tigated. In particular, the dilution induced by the FMC
Eq. (1) has been compared with a random dilution of
the same order, revealing important differences between
the two cases. The random diluted graph has a homo-
geneous topology and its phenomenology is compatible
with the mean-field solution [37, 57]. On the other hand
the inhomogeneities induced by the FMC lead to a graph
characterized by a correlated topology and its behaviour
significantly differs from the homogeneous mean-field so-
lution because of the onset of phase waves, at least at all
simulated N. Already in the ordered case, thus, the ML
model might display very strong finite size effects. The
more so when quenched disordered couplings are consid-
ered.
Large sizes are hard to simulate because the mode vari-
ables are continuous (complex) numbers and because the
total number of interactions grows like N3with the num-
ber Nof modes. Because of these effects it has not been
possible so far to identify the universality class of the
modes. By looking at the specific heat behaviour, it has
been observed [39] that for a O(N) dilution having a ran-
dom homogeneneous or a deterministic topology for the
same model makes a great difference in terms of inter-
polation of the critical properties in the thermodynamic
limit. A random homogeneous O(N) dilution of the fully
connected network allows to see, already at relatively
small sizes, a glass transition of the mean-field kind in
the same universality class of the Random Energy Model
(REM), which is the reference mean-field model for disor-
dered systems with non-linear interactions. On the other
hand the deterministic dilution yields apparently a dif-
ferent result.
To unravel such possible difference here we carefully in-
vestigate the universality class of the ML 4-phasor model,
providing simulations of systems of large enough sizes,
large statistics and, above all, introducing a trick to dras-
tically reduce finite size effects.
After a description of the model in Section II, in Sec-
tion III we explain the strategy used to reduce the finite
size effects due to the heterogeneous FMC dilution. In
Section IV we present a simple argument to get the expo-
nent for the finite-size scaling (FSS) regime of the specific
heat in the REM and generalize it deriving boundaries for
the critical exponents of a generic mean-field universality
class. We, then, compare this prediction to the specific
3
heat behaviour in the equilibrium numerical simulations
and, through FSS analysis, we assess that the scaling of
the specific heat near the glass transition temperature is
compatible with a mean-field theory, which is the main
outcome of the present work. Eventually, in Section V
we present the behaviour of the overlap probability distri-
bution upon lowering the temperature across the random
lasing transition, that turns out to be a glass transition.
The trick used to reduce finite-size effects turns out to be
useful in identifying more clearly signatures of glassiness.
II. THE MODE-LOCKED 4-PHASOR MODEL
The ML 4-phasor model has its roots in the quantum
theory of the electromagnetic field and matter interaction
in an open system. A full account of the derivation of the
classical stochastic dynamics from the quantum many-
body dynamics of light coupled with matter can be found
in [34].
The main point is that by considering laser media
where the characteristic time of atomic pump and loss are
much shorter than the lifetimes of the resonator modes,
the atomic variables, i.e., matter fields, can be removed
obtaining non-linear equations for the electromagnetic
field alone.
The stochastic differential equation for the time evolu-
tion of the modes akreads as
dak1
dt =X
k|FMC(k)
g(2)
k1k2ak2
+X
k|FMC(k)
g(4)
k1k2k3k4ak2ak3ak4+ηk1(t),(2)
where the expression of the sum over the indices ksatisfy-
ing a FMC, like (1), will be soon clarified in Eq. (4). The
dynamic variables are ak(t) = Ak(t)ek(t), the complex
amplitudes of the light modes comprised by the discrete
spectrum of the electromagnetic field
E(r, t) =
N
X
k=1
ak(t)ektEk(r) + c.c. (3)
where Ek(r) is the space-dependent wavefunction of the
mode with frequency ωk. The noise is taken as a white
noise hηk(t)i= 0, hηj(t)ηk(t0)i= 2T δjkδ(tt0), as we
will later discuss. The amplitudes ak(t) are the remnant
of the original creation and annihilation operators of the
electromagnetic field quantization, which have been de-
graded to complex numbers in the semiclassical approx-
imation. By slow amplitude mode it is meant that the
time scale of the amplitude dynamics is larger than the
time scale defined by the frequency of the mode, i.e.,
ω1
k. Therefore, in the slow amplitude approximation
the phases ektcan be averaged out, which, in Fourier
space, taking the Fourier transform of Eq. (3), implies
that ak(t)'ak(t, ω)δ(ωωk). Lasing modes are slow
amplitude modes by definition, since they are character-
ized by a very narrow linewidth γaround their frequency
ωk. The time average of the fast oscillations ektleads
to the sum termed FMC in the equation (2). The general
expression for 2n-body interactions reads as
FMC(k) : |ωk1ωk2+··· +ωk2n1ωk2n|.γ, (4)
of which Eq. (1) is the case n= 2. The FMC acts as a
selection rule on the modes participating in the interac-
tions.
The linear terms in Eq. (2) yield different contribu-
tions possibly depending on cavity gain and losses and
atom-field interaction inside the disordered medium. The
latter expression is the most relevant one in the dynam-
ics:
g(2)
k1k2ωk1ωk2
{x,y,z}
X
αβ ZV
drαβ(r)Eα
k1(r)Eβ
k2(r),(5)
where (r) is the dielectric permittivity tensor and the
integral is extended over the entire volume Vof the
medium. In particular, the diagonal elements of g(2)
k1k2
represent the net gain curve of the medium (i.e., the gain
reduced by the losses), which plays an important role
mainly below the lasing threshold.
The non-linear couplings g(4)
k1k2k3k4are given by the spa-
tial overlap of the electromagnetic mode wavefunctions
modulated by a non-linear optical susceptibility χ(3)
g(4)
k1k2k3k4
4
Y
j=1
ωkj
{x,y,z}
X
αβγδ ZV
drχ(3)
αβγδ ({ωk};r)
×Eα
k1(r)Eβ
k2(r)Eγ
k3(r)Eδ
k4(r),(6)
where, again, the integral is over the whole volume of the
medium. In general, both the linear and the non-linear
couplings are complex numbers and can be written as
[50]
g(2)
k1k2=Gk1k2+iDk1k2,(7)
g(4)
k1k2k3k4= Γk1k2k3k4+ik1k2k3k4.(8)
In the standard laser case the linear couplings are di-
agonal and the non-linear ones can be safely considered
as constant: in this case, Dkis the group velocity dis-
persion coefficient and ∆ is the self-phase modulation
coefficient, responsible for the Kerr effect [40]. In the
purely dissipative limit [34, 37], i.e. Dk1k2Gk1k2and
k1k2k3k4Γk1k2k3k4, which in standard laser theory
corresponds to neglect the group velocity dispersion and
the Kerr effect, the dynamics of Eq. (2) becomes a po-
tential differential equation
dak1
dt =H[a]
ak1(t)+ηk1(t),
4
with a Hamiltonian function given by
H=X
k|FMC(k)
Gk1k2ak1ak2
X
k|FMC(k)
Γk1k2k3k4ak1ak2ak3ak4+ c.c..(9)
In principle, the noise is correlated, i.e. hηk1ηk2i 6=
δk1k2. However, it can be diagonalized by changing ba-
sis of dynamic variables: the decomposition of resonator
modes into a slow amplitude basis is not unique [58] and
one can use this freedom to build a basis in which the
noise has no correlations. The diagonalization of the
noise can be done at the cost of having non-diagonal lin-
ear interactions, which is not a real complication in the
random laser case, since linear couplings already have
off-diagonal contributions accounting for the openness of
the cavity.
The laser dynamics is brought to stationarity by gain
saturation, a phenomenon connected to the fact that,
as the power is kept constant, the emitting atoms pe-
riodically decade in lower states saturating the gain of
the laser. In the same way the dynamics induced by
the Hamiltonian Eq. (9) eventually reaches a stationary
regime, when a constraint on the total energy contained
in the system is added. This argument was first pro-
posed for standard multimode lasers in Refs [37, 57].
In fact, lasers are strongly out of equilibrium: energy
is constantly pumped into the system in order to keep
population inversion and stimulated emission, and in the
case of cavityless systems also compensate the leakages.
However, a stationary regime can be described as if the
system is at equilibrium with an effective thermal bath,
whose effective temperature (a “photonic” temperature)
accounts both for the amount of energy E=N stored
into the system because of the external pumping and for
the spontaneous emission rate. The latter is proportional
to the kinetic energy of the atoms, e. g., to the heat bath
temperature T. Eventually, the external parameter driv-
ing the lasing transition turns out to be [33, 34, 50]
Tphotonic =T
2.(10)
One can also introduce the pumping rate P[41] as the
inverse of the square root of this ratio:
P2=2
T=1
Tphotonic.
In order to mathematically model the gain saturation,
an overall spherical constraint can be imposed on the
amplitudes fixing the total optical intensity in the system
N
X
k=1 |ak|2=N. (11)
The precise value of the couplings in the Hamiltonian
Eq. (9) requires the knowledge of the spatial wavefunc-
tions of the modes, see Eqs. (5) and (6), which is not
available in random lasers, since they are characterized
by a complicated spatial structure of the modes. If, as
it apparently occurs in glassy random lasers, modes are
spatially extended to wide regions of the optically ac-
tive compound, each mode is nonlinearly interacting with
very many others. We will implement such an “extended
modes approximation” [59] in our model, where the only
relevant factor in the mode-coupling is the FMC, rather
than spatial confinement of light modes. In this case,
because of thermodynamic convergence, each coupling
coefficient will be smaller and smaller as the number of
modes increases.
In principle, all couplings involving the same mode
will be correlated. However, because of the spatially
etherogeneous optical nonlinear susceptibility in (6) and
the fact that each coupling coefficient vanishes as Nin-
creases, the role of correlation will be qualitatively neg-
ligible as far as the system displays enough modes. For
this reason, the couplings will be taken as independent
Gaussian random variables in the present work:
P(Jk1···kp) = 1
q2πσ2
p
exp (J2
k1···kp
2σ2
p),(12)
with p= 2,4 and σ2
pN2pto ensure the extensivity of
the Hamiltonian and where some rescaling of the modes
and coefficients (gJ) has been performed [34, 41].
Eventually, the stationary properties of the system can
be described by a model whose Hamiltonian is
H[a] = H2[a] + H4[a],(13)
where
H2[a] = X
k|FMC(k)
Jk1k2ak1ak2+ c.c.
H4[a] = X
k|FMC(k)
Jk1k2k3k4ak1ak2ak3ak4+ c.c. (14)
As mentioned in section II when introducing the dissi-
pative limit we will consider the J’s as real parameters,
without loss of generality. The effective distribution for
the phasor configuration a={a1, . . . , aN}will, eventu-
ally, be
P[a]eβH[a]δ N
N
X
k=1 |ak|2!,(15)
where βis the inverse temperature.
III. FREQUENCY MATCHING CONDITION
WITHOUT EDGE-BAND MODES
The FMC Eq. (4) is the most peculiar aspect of the
ML 4-phasor model, since it defines the topology of the
interaction network. The full inclusion of the FMC in
摘要:

UniversalityclassoftheglassyrandomlaserJacopoNiedda,1,2GiacomoGradenigo,3,4,2LucaLeuzzi,2,1,andGiorgioParisi1,2,5,61DipartimentodiFisica,UniversitadiRoma\Sapienza",PiazzaleA.Moro2,I-00185,Roma,Italy2NANOTECCNR,SoftandLivingMatterLab,Roma,PiazzaleA.Moro2,I-00185,Roma,Italy3GranSassoScienceInstitute...

展开>> 收起<<
Universality class of the glassy random laser Jacopo Niedda1 2Giacomo Gradenigo3 4 2Luca Leuzzi2 1and Giorgio Parisi1 2 5 6 1Dipartimento di Fisica Universit a di Roma Sapienza Piazzale A. Moro 2 I-00185 Roma Italy.pdf

共17页,预览4页

还剩页未读, 继续阅读

声明:本站为文档C2C交易模式,即用户上传的文档直接被用户下载,本站只是中间服务平台,本站所有文档下载所得的收益归上传人(含作者)所有。玖贝云文库仅提供信息存储空间,仅对用户上传内容的表现方式做保护处理,对上载内容本身不做任何修改或编辑。若文档所含内容侵犯了您的版权或隐私,请立即通知玖贝云文库,我们立即给予删除!

相关推荐

分类:图书资源 价格:10玖币 属性:17 页 大小:1.97MB 格式:PDF 时间:2025-05-06

开通VIP享超值会员特权

  • 多端同步记录
  • 高速下载文档
  • 免费文档工具
  • 分享文档赚钱
  • 每日登录抽奖
  • 优质衍生服务
/ 17
客服
关注