Dynamic modulation of thermal emission - a tutorial Michela F. Picardi1Kartika N. Nimje1and Georgia T. Papadakis1a ICFO - Institut de Ciencies Fotoniques The Barcelona Institute of Science and Technology

2025-05-03 0 0 3.18MB 33 页 10玖币
侵权投诉
Dynamic modulation of thermal emission - a tutorial
Michela F. Picardi,1Kartika N. Nimje,1and Georgia T. Papadakis1, a)
ICFO - Institut de Ciencies Fotoniques, The Barcelona Institute of Science and Technology,
Castelldefels (Barcelona) 08860, Spain
(Dated: 17 February 2023)
Thermal emission is typically associated with a blackbody at a temperature above absolute zero, which ex-
changes energy with its environment in the form of radiation. Blackbody thermal emission is largely incoherent
both spatially and temporally. Using principles in nanophotonics, thermal emission with characteristics that
differ considerably from those of a blackbody have been demonstrated. In particular, by leveraging intrin-
sic properties of emerging materials or via nanostructuring at the wavelength or sub-wavelength scale, one
can gain control over the directionality, temporal coherence, and other more exotic properties of thermal
radiation. Typically, however, these are fixed at the time of fabrication. Gaining dynamic control of ther-
mal emission requires exploiting external mechanisms that actively modulate radiative properties. Numerous
applications can benefit from such thermal emission control, for example in solar energy harvesting, thermo-
photovoltaic energy conversion, radiative cooling, sensing, spectroscopy, imaging and thermal camouflage. In
this tutorial, we introduce thermal emission in two domains: the far-field, and the near-field, and we outline
experimental approaches for probing thermal radiation in both ranges. We discuss ways for tailoring the
spatial and temporal coherence of thermal emission and present available mechanisms to actively tune these
characteristics.
I. INTRODUCTION: THERMAL RADIATION
From the sun itself, to the discovery of fire, all the way
to incandescent light bulbs, radiation generated by hot
objects has been the world’s primary source of illumi-
nation consistently throughout history. While radiation
from burning objects was already known to our cavemen
ancestors, the fact that all bodies emit thermal radia-
tion if their temperature is above 0 K was only realized
in the nineteenth century. Only a small portion of this
radiation is visible to the unaided human eye, and it orig-
inates from bodies at very high-temperatures, such as the
sun. At a temperature of nearly 6000 K, the sun emits
light mainly at visible and near-infrared (IR) frequen-
cies. At lower, terrestrial temperatures, near 300 K, from
Planck’s law of thermal radiation1, the peak of a black-
body’s thermal emission spectrum shifts to the mid-IR
range, corresponding to wavelengths in the range of 5-20
µm.
The first connection between light and its thermal
properties is tied to the discovery of IR radiation by
William Herschel in 18002. Observing the spectrum of
sunlight, dispersed through a prism, Herschel utilized a
series of thermometers to measure the temperature of the
radiation corresponding to each color. He then placed a
thermometer beyond the visible red end and recorded a
high-temperature, unveiling the existence of invisible ra-
diation carrying thermal energy. Driven by the interest
to explain astronomical observations, numerous attempts
were subsequently made to describe thermal radiation, a
notable one resulting in the Rayleigh-Jeans law3, which
predicts the spectral radiance of a blackbody to be pro-
portional to λ4. While constituting a good approxima-
a)georgia.papadakis@icfo.eu
tion for long wavelengths, the Rayleigh-Jeans law hints
at an ultraviolet catastrophe at smaller wavelengths, for
which the spectral radiance would diverge. Max Planck
resolved the paradox by introducing his famous law of
thermal radiation1. Indeed, Planck’s theory reduces to
Rayleigh-Jeans law for large values of λ, yet it also ac-
curately predicts thermal emission from a blackbody at
short wavelengths.
Other than illumination, thermal radiation is also the
most abundant source of energy on our planet. The to-
tal solar daytime irradiance reaches 1,360 W/m2. This
power density suffices to meet the world’s energy needs
for an entire year, if one collected all the energy that
reaches the earth for just one hour. Harvesting this abun-
dant renewable energy source for electricity is the sub-
ject of a large portion of modern research4, with solar
photovoltaic cells having reached the stage of a mature
technology5,6. Solar photovoltaic cells harness the por-
tion of solar photons with frequencies in the visible range.
For a single junction photovoltaic cell, the fundamental
efficiency limit of roughly 30% was derived by Shockley
and Queisser in 19617.
Together with the sun, terrestrial objects in the macro-
as well as in the micro-scale emit mid-IR radiation as a re-
sult of local heating. Examples range from exhaust gases
from a power plant, to a hot stove top, to an operating
microprocessor. Due to its abundance, this radiant heat,
often termed waste heat, presents a significant opportu-
nity for heat-to-electricity energy conversion, recycling,
and storage. A significant portion of waste heat resides
at temperatures below 1000 K8, and this, from a thermo-
dynamic point of view, hinders its efficient conversion.
Since the ultimate thermodynamic efficiency of a heat
engine is 1 TC/TH(the Carnot limit), where TCand
THare the temperatures of the cold and hot reservoirs,
respectively, reaching high efficiencies at low-grade waste
heat temperatures (TH<1000 K) is a challenge9.
This is the author’s peer reviewed, accepted manuscript. However, the online version of record will be different from this version once it has been copyedited and typeset.
PLEASE CITE THIS ARTICLE AS DOI: 10.1063/5.0134951
2
To harness waste heat, one can use a thermo-
photovoltaic system, which operates similarly to a solar
photovoltaic cell. In the thermo-photovoltaic case, how-
ever, instead of the sun, heat elevates the temperature
of a local thermal emitter, which in turn emits radiation
towards a low-band gap photovoltaic cell, whose band
gap energy corresponds to the average energy of photons
from the thermal emitter10. The characteristics of ther-
mal emission can be engineered to match the properties of
the corresponding photovoltaic cell, thus, theoretical pre-
dictions estimate thermo-photovoltaics performance near
thermodynamic limits5,11–13. Experimental demonstra-
tions of thermo-photovoltaic systems with efficiencies in
the range of 40% have already been reported, although
this peak efficiency was reached at high emitter temper-
atures (2400C)5.
Beyond energy and lighting, harnessing and controlling
thermal radiation finds applications in any technologi-
cal platform where regulating a temperature is relevant.
One notable example is radiative cooling14–23. Radia-
tive cooling is a technique via which the particular fea-
tures of the spectrum of the atmosphere’s transmittance
are exploited to decrease the temperature of objects on
earth24. In particular, the atmosphere is transparent be-
tween 8 and 13 µm, a frequency range which corresponds
to the peak of the emission of bodies at room tempera-
ture. Therefore, all bodies at room temperature emit
thermal radiation which can escape the atmosphere and
thermalize with the coldness of the universe at 3K. In
order to cool down, however, the body’s total energy bal-
ance needs to be negative: the energy the body absorbs
from its surroundings must be smaller than the energy it
emits25. For this reason, radiative cooling devices have
been designed using a nearly ideal mirror to reject solar
photons from the sun, while at the same time, they emit
efficiently at frequencies within the atmospheric trans-
parency window.26–28.
Other applications in which the control of thermal
radiation is crucial include thermal camouflaging29–33
and circuitry34–43. Applications in sensing44, thermal
imaging45,46, and spectroscopy47,48 are also relevant to
thermal photonics. The field of thermal photonics aims
to tailor the incoherent thermal emission from black-
bodies to meet the requirements of various applications.
Very interesting results demonstrating a high degree of
both spatial and temporal coherence have been recently
reported13.
This tutorial is structured as follows: first, we intro-
duce our notation and theoretical framework, and distin-
guish between the thermal near-field and far-field. We
briefly discuss the spectral and spatial coherence of both
near- and far- field thermal radiation, and comment on
the techniques adopted to measure thermal emission in
both ranges. Next, we describe prominent mechanisms
available for dynamic control of thermal emission and
their potential applications. We focus individually on
each distinct phenomenon which can be used to achieve
dynamic modulation, yet we note that, in practice, sev-
0.1
Energy (eV)
T=300K
T=600K
T=1000K
Spectral energy density (eV s/µm³) 108
0.2 0.3 0.4 0.5 0.6
0
0.5
1
1.5
2
FIG. 1. Spectral energy density for a blackbody calculated
using Eq. 1 for T= 300 K, 600 K and 1000K.
eral of the described phenomena may be combined to
design mid-IR photonic devices for applications.
II. PLANCK’S LAW OF THERMAL RADIATION
Planck’s law describes how the temperature of a body
determines the amplitude and spectrum of its thermal
emission1. It states that the spectral energy density (en-
ergy per unit volume), φ(ω), for an emitter at tempera-
ture Tand frequency ωcan be expressed as:
φ(ω) = E(ω, T )×DOS,(1)
where E(ω, T ) is the mean thermal energy per photon,
given by E(ω, T ) = ¯N(ω, T ), with ¯hbeing the re-
duced Planck’s constant, and N(ω, T ), the photon num-
ber, which is determined via the Bose-Einstein distribu-
tion:
N=1
e
¯
kBT1
,(2)
where kBis the Boltzmann constant. The factor DOS in
Eq. 1 is the photonic density of states, which quantifies
the number of photonic states available per frequency
and unit volume.
Via Eq. 1, it already becomes possible to identify
a mechanism available for dynamic tuning of thermal
emission: modulation of the temperature of a blackbody
emitter. As shown in Fig. 1, changing the temperature
in Eq. 1 leads to significant changes in both the central
frequency and the amplitude of the spectral density
of a blackbody emitter. Tuning the temperature of
an emitter is what enables thermo-optical modulation,
as well as a mechanism to induce transitions in phase
change materials, as we will discuss later on.
This is the author’s peer reviewed, accepted manuscript. However, the online version of record will be different from this version once it has been copyedited and typeset.
PLEASE CITE THIS ARTICLE AS DOI: 10.1063/5.0134951
3
Planck’s law assumes that the distance from an emit-
ting body, at which the spectral flux is measured, as well
as the dimensions of the body itself, are larger than the
thermal wavelength, which is given by:49:
λT=π2/3¯hc
kBT.(3)
As we shall see below, raising these assumptions and con-
sidering separation distances much smaller than λTyields
thermal emission and radiative heat transfer rates that
can exceed the blackbody limit50.
III. FAR-FIELD AND NEAR-FIELD RADIATIVE HEAT
TRANSFER
With respect to λTas defined in the previous section,
we broadly distinguish between two spatial regions: the
region that is sufficiently far away from a thermal emit-
ter, for which dλT, or far-field (FF), and the region in
close proximity to the emitter for which dλT, termed
thermal near-field (NF). We note that an alternative def-
inition to NF that takes into account the emitter size
and is irrespective of its temperature is also discussed in
literature51,52.
In order to understand the key differences between
these two regions, let us start by considering a local
monochromatic emitter in an otherwise lossless, isotropic
and homogeneous background, placed at r=r0. The
electric field at a point rcan be described as a plane
wave:
E(r, ω) = E0ei[k·(rr0)ωt],(4)
where E0is the amplitude of the field at r0,kis the
wavevector, ωis the angular frequency and tis time.
If the wavevector kis a real number, then at a distance
r from the source, the field acquires a phase eik·(rr0)
while its amplitude is preserved. We shall refer to light
with such a wavevector as a propagating wave. By defi-
nition, FF radiation is entirely comprised of propagating
waves. In Fourier space, as shown in Fig. 2(c), propa-
gating waves are confined to the volume inside the light-
cone defined by k0=nω
c,nbeing the refractive index
of the surrounding medium and cthe speed of light in
vacuum51. By contrast, if the transverse component of
the wavevector extends to values of koutside the light-
cone, the wavevector in Eq. 4 becomes complex, resulting
in an exponentially decaying wave away from its source,
also termed evanescent wave.
The scenario of two objects exchanging radiative heat
in the FF is shown in Fig. 2 (a) for the case of planar
surfaces at different temperatures. By contrast, when the
objects are placed in close proximity (dλT), we shall
refer to them as being in the NF of one another, Fig. 2
(b), where interesting phenomena beyond Planck’s law
can emerge53.
far eld
near eld
(d=3μm)
near eld
(d=1μm)
near eld heat ux
(TH² - TC²)
evanescent
modes
far eld heat ux
x
y
σ (TH⁴ - TC)
k₀ =
propagating
modes
c
z
x
yz
FIG. 2. Schematic depiction of two planar surfaces exchang-
ing thermal radiation in (a) the far-field (FF) and (b) near-
field (NF). FF, propagating waves preserve their amplitudes
with their phase oscillating away from the planar surfaces,
while the amplitude of evanescent waves decays exponen-
tially. (c, d) Schematic depiction of FF and NF waves in
Fourier space. For propagating waves, the amplitude of the
wavevector (k) is bounded (c) inside the light cone, delim-
ited by k0=nω
cand denoted by dashed lines, whereas (d)
evanescent waves lie outside the lightcone. In (e), the NF
heat flux calculated between two SiC planar surfaces with
0< ω < 10 ΩSiC , ΩSiC being the resonant frequency of bulk
SiC, and k0< kt<1
dat different distances is shown in com-
parison to the FF, as a function of the temperature difference
T=THTC.
A. Far-field
In the FF, we can express the DOS in Eq. 1 as:
DOSFF =ω2
π2c3.(5)
Plugging this in Planck’s law (Eq. 1), we can evaluate
the electromagnetic energy emitted from a blackbody at
temperature T. It is useful to introduce now a quantity
called emissivity, which serves as a metric of the efficiency
of an emitting body as a thermal emitter54. It is defined
as the ratio between the energy emitted by a body at
temperature Tand that emitted by a blackbody at the
same temperature. Hence, blackbodies are characterised
by unitary emissivity, i.e.: EBB = 1, at all frequencies.
The emissivity, E, obtains values between 0 (a perfect
reflector) and 1 (a perfect emitter). We note that the
This is the author’s peer reviewed, accepted manuscript. However, the online version of record will be different from this version once it has been copyedited and typeset.
PLEASE CITE THIS ARTICLE AS DOI: 10.1063/5.0134951
4
emissivity, as defined here, is a FF property of emitters
with no counterpart in the NF. Due to Kirchhoff’s law of
thermal radiation55, by reciprocity, Eis also equal to the
optical absorption Aof a material. Often, in literature,
a more realistic emitter is considered, termed greybody,
characterized by a constant emissivity E<EBB . This
is still an idealization as emissivity in general depends
on frequency, but can be a good approximation for some
materials in given spectral ranges. As an example, in
Fig. 3, the spectral energy density of a greybody with
E= 0.25 EBB (orange curve) is compared to that of a
blackbody (red curve).
The DOSFF of equation 5 is the density of states eval-
uated in vacuum. If, instead of vacuum, the emitter is
embedded into a lossless medium, the intensity of the ra-
diation scales proportionally to n2=ε, where nis the
refractive index of that medium and εis its dielectric
function56. This still holds if nis both a function of fre-
quency and position. The dielectric function of materials
is a key property for actively tailoring thermal emission,
as will be discussed below. In the most general case, εis
a 3 ×3 tensor with up to 6 independent degrees of free-
dom. For isotropic materials εis a scalar quantity. In the
following we will see examples of dynamic control of the
scalar dielectric function to modulate thermal emission,
such as thermo-optical and electrostatic modulations, as
well as mechanisms in which the tensor nature of εis fun-
damental, as is the case for magneto-optical modulation.
Let us consider the FF heat exchange between two
greybodies as in Fig. 2 (a). The heat transfer be-
tween them can be accurately described by the Stefan-
Boltzmann law. Upon integrating spatially and spec-
trally the heat flux exchange of two bodies, one at
temperature TCand the other at temperature TH, for
TH> TC, the Stefan-Boltzmann law predicts a total heat
exchange that is proportional to (T4
HT4
C). We therefore
note that, in the FF, thermal emission as well as the to-
tal radiative heat transfer is independent of the distance
from the emitter.
On the other hand, when the heat exchange between
the two bodies occurs at distances comparable or below
the thermal wavelength λT, as in Fig. 2 (b), the modes
that dominantly contribute to the heat transfer do not
follow Stefan-Boltzmann law (blue and green curves), as
can be seen in Fig. 2(e). This is explained in the following
section.
B. Near-field
In close proximity to the emitter, the radiative heat
flux is mediated by evanescent fields that are highly local-
ized close to the emitting body, and quickly decay away
from it, as shown for example in Fig. 2 (b) for the case
of two planar bodies. In the NF, the DOS assumes dif-
ferent values depending on how far from the emitter it is
measured, hence it is usually referred to as a local density
of states, LDOS, to take into account this spatial depen-
0 0.5 1 1.5 2
10-
10-²
10⁰
10
²
SiC (d=100 nm)
Blackbody
Greybody
Angular frequency ω [in ΩSiC]
Spectral energy density (J s/μm³)
FIG. 3. Spectral energy density calculated using equation 1
(we note that the yaxis is in logarithmic scale). In blue, the
spectral energy density is calculated in the NF of a SiC emit-
ter at a distance d= 100 nm. We see that the NF thermal
emission shows a highly enhanced energy density around a
narrowband peak. In red and orange, we show the FF spec-
tral energy densities evaluated for a black (EBB = 1) and a
greybody (E= 0.25), respectively. The xaxis is normalized
with respect to ΩSiC =qεω2
LO +ω2
TO
1+ε, which is the phonon
polariton resonance frequency for bulk SiC.
dence. The LDOS can be calculated by considering a
thermal emitter as a superposition of point dipoles and
using the dyadic Green’s function to estimate the elec-
tric and magnetic fields generated by their oscillation52.
To relate these oscillations to macroscopic quantities (for
instance, the temperature), we impose the fluctuation-
dissipation theorem on the current density generated by
these dipoles. This determines its statistical correla-
tion. We can therefore write the LDOS for a thermal
emitter50,57,58. Following Joulain et al.,59, we highlight
the main steps towards the derivation of the LDOS, al-
beit referring the reader to dedicated resources for further
details59–61. The electric field correlation function can be
written as:
Eij (r,r0, t t0) = 1
2πZ
−∞
Eij (r,r0, ω)e(tt0)=
=Ei(r, t)E
j(r0, t0).
The electric field can be written from the electric current
density as:
E(r, ω) = 0ωZGE(r,r0, ω)·j(r0)d3r0,
where GEis the electric field’s dyadic Green’s function,
jis the current density and µ0is the magnetic vacuum
permeability51. Via fluctuation-dissipation theorem, we
This is the author’s peer reviewed, accepted manuscript. However, the online version of record will be different from this version once it has been copyedited and typeset.
PLEASE CITE THIS ARTICLE AS DOI: 10.1063/5.0134951
5
can write:
Eij (r,r0, ω) = ¯
e
¯
kBT1
µ0ω
2πIm[GE
ij (r,r0, ω)].
A similar procedure can be applied to the magnetic field,
using its field’s correlation and dyadic Green’s function.
Adding the electric and magnetic contributions, and con-
sidering only the positive frequencies, it is possible to
obtain the local spectral energy density:
φ(r, ω) = ¯
e
¯
kBT1
ω
πc2Im{Tr[GE(r,r0, ω)+GH(r,r0, ω)]},
from which the local density of states can be written as:
LDOSNF =ω
πc2Im{Tr[GE(r,r0, ω)+GH(r,r0, ω)]}.(6)
The Green’s functions in Eq. 6 depend on the consid-
ered geometry. Importantly, in vacuum, Im[Tr(GE)] =
Im[Tr(GH)]. From the boundary conditions at infinity
r→ ∞, one retrieves the DOSFF which we wrote in
Eq. 5. There are also geometries for which the magnetic
Green’s function is negligible compared to the electric
one, leading to a simplified expression for the LDOS. For
example, if we consider a planar emitter, calculating the
fields at a distance dλ, in other words in the qua-
sistatic limit, for some materials it is possible to approx-
imate the LDOS as62:
LDOSNF =1
4π2ωd3
Im[ε(ω)]
|ε(ω)+1|2,(7)
where εis the dielectric function of the emitting material,
which is a function of frequency. This approximation is
useful for obtaining insight on two important concepts.
Firstly, we see that in the lossless limit (Im[(ω)] 0),
the LDOSNF vanishes and no thermal emission occurs,
as expected. This unveils clearly the tight link between
optical losses and thermal radiation. Secondly, Eq. 7 is
instructive to understand near field heat transfer in the
presence of polaritons.
Polaritons are quasiparticles which derive from the col-
lective oscillation of light and matter. For example, they
appear in materials which present a large enough carrier
density and mobility, such as metals, to allow the elec-
trons to undergo oscillations driven by the electric field
at given frequencies. In the following, unless specified,
we will use the terms “metals” and “plasmonic mate-
rials” interchangeably. Analogously, in polar dielectric
materials, phonon polaritons can be excited from the
coupling between the electromagnetic field with the ma-
terial’s optical phonons63–65. In the following, this type
of material will be referred to as “polar material”. In
both cases, the excitation of a polariton is dependent on
the frequency of the incident light. Therefore, the re-
sponse function of a polaritonic material will be resonant
around a frequency, at which matter (charge or lattice
vibration in plasmonic and polar media, respectively) is
oscillating synchronously with the excitation, with min-
imal lag66. This response function is given by the lin-
ear polarizability of the material, which is proportional
to the dielectric function67. The two resonant models
most commonly used to describe the dielectric function
of polaritonic materials are the Drude model, suitable for
metals, and the Lorentz model, suitable for polar semi-
conductors and dielectrics68,69. According to the Drude
model, the dielectric function of a plasmonic material can
be written as:
εpl(ω) = ε 1ω2
p
ω2ω !,(8)
where εis the the dielectric constant evaluated at the
limit for very high frequencies, usually ε1 for metals,
ωpis the plasma frequency and γis the damping rate due
to the oscillating electrons colliding with each other. The
Lorentz model, on the other hand, predicts the dielectric
function for a polar material70:
εpo(ω) = ε1 + ω2
LO ω2
TO
ω2
TO ω2ω ,(9)
with ωLO and ωTO being the resonant frequencies cor-
responding to the longitudinal and transverse optical
phonon resonances in the polaritonic material71–73. The
frequencies ωTO and ωLO delimit the region in which the
dielectric function of the material is negative, which de-
fines the Reststrahlen band64,74.
In Fig. 4, the real part of the dielectric function of a
metal (Drude) and a polar dielectric (Lorentz) material
are plotted in orange and blue, respectively.
At the resonance frequency of a polaritonic
material64,65, when (ω) = 1, the LDOSNF res-
onates. Therefore, when polaritonic excitations are
available, the LDOS in the NF can dramatically exceed
the DOS of the FF, leading to a spectral energy density
which surpasses the blackbody spectrum in magni-
tude. In this case, it is often said that one can obtain
“super-Planckian” thermal emission76. Moving away
from the emitter, the magnitude of the fields decreases
exponentially, and both the LDOSNF and the thermal
emission spectrum ought to converge to their FF value,
which is independent of the distance.
Where does the difference between thermal NF and FF
come from? As discussed above, in contrast to the FF,
waves that occur in the NF have an in-plane wavevector
component kt=qk2
x+k2
ythat lies outside the vacuum
lightcone as shown in Fig. 2(d), in other words kt> k0.
Since the total wavevector satisfies k=|k|=pk2
t+k2
z,
this requires kz∈ I. This explains the exponential decay
of the evanescent excitation in the direction orthogonal
to ktin vacuum, which we can refer to as zwithout loss
of generality (following the reference system of Fig. 2
(a, b)). Hence, along z, the field does not propagate,
This is the author’s peer reviewed, accepted manuscript. However, the online version of record will be different from this version once it has been copyedited and typeset.
PLEASE CITE THIS ARTICLE AS DOI: 10.1063/5.0134951
摘要:

Dynamicmodulationofthermalemission-atutorialMichelaF.Picardi,1KartikaN.Nimje,1andGeorgiaT.Papadakis1,a)ICFO-InstitutdeCienciesFotoniques,TheBarcelonaInstituteofScienceandTechnology,Castelldefels(Barcelona)08860,Spain(Dated:17February2023)Thermalemissionistypicallyassociatedwithablackbodyatatemperatu...

展开>> 收起<<
Dynamic modulation of thermal emission - a tutorial Michela F. Picardi1Kartika N. Nimje1and Georgia T. Papadakis1a ICFO - Institut de Ciencies Fotoniques The Barcelona Institute of Science and Technology.pdf

共33页,预览5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!
分类:图书资源 价格:10玖币 属性:33 页 大小:3.18MB 格式:PDF 时间:2025-05-03

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
/ 33
客服
关注