Sunteți pe pagina 1din 16

Appl. Phys.

B 65, 313–328 (1997)


C Springer-Verlag 1997

Invited paper

Spatial coherence of short wavelength high-order harmonics


T. Ditmire1 , J.W.G. Tisch1 , E.T. Gumbrell1 , R.A. Smith1 , D.D. Meyerhofer2 , M.H.R. Hutchinson1
1 Blackett Laboratory, Imperial College of Science Technology and Medicine, London SW7 2BZ, United Kingdom
2 Department of Mechanical Engineering, University of Rochester, Rochester, New York 14627, USA

Received: 12 May 1997

Abstract. Nonperturbative high-order harmonic conversion the potential of using these harmonics as a source of high-
of high-intensity, ultrafast laser pulses in gases represents brightness, coherent, soft X-ray radiation. A number of recent
a unique means to generate high-brightness, coherent XUV, experiments have explored the possibility of using the har-
and soft X-ray radiation. We review the physics controlling monics in various applications requiring short pulses of XUV
the spatial qualities of harmonic radiation and recent experi- or soft X-ray radiation. Haight et al. have used harmonics
mental measurements of high-order harmonic spatial profiles. in time resolved pump probe photoelectron spectroscopy of
We also examine the factors controlling the spatial coherence GaAs [9, 10]. Balcou et al. utilized harmonics in photoioniza-
of the harmonics. A detailed series of Young’s two-slit experi- tion spectroscopy of noble gases with photon energy as high
ments that measure the spatial coherence of the harmonics is as 100 eV [11] and Larsson et al. exploited the short pulse na-
presented. These measurements indicate that the harmonics ture of the harmonics to measure the lifetime of the 1s2 p 1P
exhibit good fringe visibility and high spatial coherence. state in helium [12]. XUV harmonics have also been used to
probe the dynamics of laser plasmas [13].
PACS: 42.65.Ky; 42.25.Kb Aside from these applications, the properties of these har-
monics have been very carefully studied by a number of
groups around the world in recent years [4, 5, 14, 15]. Probing
the shortest wavelengths attainable with harmonic generation
Since the first observation of nonperturbative high-order har- has been a topic of extensive research. In addition to the
monic generation of laser light in a gas by McPherson et al. 155th harmonic of a Ti:sapphire laser, the 141st harmonic
in 1985 [1], high-order harmonics have received considerable from a 1 ps, 1054 nm laser has been observed [16] and the
attention as a potential high brightness, coherent source of 37th harmonic of a 248 nm KrF laser has been reported [17].
X-rays [2]. High-order harmonics are produced when a high The harmonic spectra that result in most experiments typi-
intensity laser is focused into a gas, and the nonlinear os- cally have a number of important features in common. The
cillations of the atoms in the medium re-emit photons at harmonic yield drops exponentially over the first few orders
odd harmonics of the incident laser light. Though low-order (up to the 7th or 9th). This drop is followed by a long plateau
harmonics, such as those with harmonic order less then the of roughly constant yield between harmonics, which is in turn
seventh or ninth, have been studied for many years [3], the followed by an abrupt cutoff. Theoretical and experimental
experiment of [1] and many others that followed soon after studies of this behavior indicate that the location of the har-
illustrated that at high intensities (> 1013 W/cm2 ) the expo- monic cutoff, and hence the shortest wavelengths achievable,
nential drop in yield with increasing harmonic order usually is usually given by a universal scaling law [18–20]:
associated with a lowest-order perturbation theory treatment
does not continue with higher harmonic orders [4–7]. Instead, hωcut ≈ Ip + 3.2Up , (1)
above a low order the harmonic spectrum exhibits a plateau of
many harmonics with roughly equal yield, which can reach to where Ip is the atom’s ionization potential, Up is the pon-
very high orders. For example, harmonics as high as the 155th deromotive field of the laser, and hωcut is the cutoff harmonic
order have been observed from the harmonic conversion of photon energy. This law has been confirmed in many experi-
800 nm light in helium [8]. These very high-order harmon- ments though some deviations from this scaling have been
ics have wavelengths that stretch into the XUV and soft X-ray observed [21, 22].
region. Other properties of the high-order harmonics have been
The high-order harmonics may represent a very attractive studied as well. The energy yields and conversion efficiency
source of short wavelength radiation. The extensive research of the harmonics have been measured. Yields as high as 60 nJ
of high-order harmonic generation has been motivated by in the 200 Å wavelength range have been reported [23]. The
314

spectral bandwidth properties have also been examined, in- two-slit experiments to measure the spatial coherence of soft
dicating that the harmonics exhibit linewidths comparable X-ray, high-order harmonics in the 270 Å to 480 Å wave-
to or larger than perturbation theory estimates for the trans- length range. In general, we find that these harmonics exhibit
form limited bandwidth [21, 24]. Recently techniques have good fringe visibility and high spatial coherence.
also been developed to measure the harmonic pulse width
as well [25–28]. These studies have confirmed that the har-
monic pulse width is close to that of the driving laser and 1 Harmonic production in a nonlinear, ionizing medium
therefore using femtosecond laser pulses, soft X-ray pulses
of under 100 fs can be generated. Glover et al. measured the To investigate the physical mechanisms affecting the spatial
harmonic pulse width by cross correlating the harmonic with properties of the high-order harmonic production, it is useful
the laser pulse by using photoionization [26]. This measure- to examine the propagation equations governing the build-up
ment indicated that the spatially integrated harmonic pulse of the harmonic field in a nonlinear medium driven by the
width from a 100 fs laser varied from 80 to 300 fs, depending intense laser pulse. Because the efficiency of the harmonic
upon the conditions. Similar results have been obtained using process is generally low, we can usually ignore any pump de-
ionization gated absorption [27]. Tisch et al. have also meas- pletion on the fundamental laser beam. We also consider laser
ured the harmonic pulsewidths from 1 ps laser pulses by using pulses that have pulses widths much greater then the time
chirped pulse temporal spectroscopy and have found that if scale of a single wavelength. This fact allows us to use the
the pulses are not spatially integrated, the harmonic pulse can slowly varying envelope approximation in deriving the wave
be substantially shorter then the drive laser pulse [28]. In add- equation governing the harmonic’s growth [2].
ition to these studies of harmonics produced in gases of single In the slowly varying envelope approximation, the wave
atoms, the properties of harmonics produced in molecules equation for the qth harmonic’s field oscillating at a frequen-
[29, 30] and clusters [31] have also been examined in experi- cy ωq is
ments. Recently harmonics produced from laser interactions
with solid target plasmas are beginning to receive attention as n 2q ω2q 4πω2q
well [32]. ∇ 2 Aq + Aq = − Pq , (2)
Aside from these experiments, there is particular interest c2 c2
in the spatial properties of the high-order harmonics. Knowl- where Aq is the field strength of the qth harmonic, n q is the
edge of the spatial divergence of the harmonics is important spatially and temporally varying refractive index of the me-
if the harmonics are to be used in applications. It also serves dia at ωq , and Pq is the polarization induced by the laser field
further to illuminate the nature of the physics controlling the at a frequency ωq . Equation (2) ignores the group velocity
harmonic generation. In addition, characterization of the spa- dispersion of the harmonic pulse as well as the group veloc-
tial coherence is of particular importance if harmonics are ity walk-off of the harmonic pulse with respect to the laser
to be used in interferometric applications. The coherence of pulse. Both of these effects are negligible for pulses of 100 fs
high-order harmonics is expected to be high since the ra- or longer in a low density (∼ 1019 atoms/cm3 ) gas medium.
diation is created by the parametric conversion of coherent, To further simplify this equation, we can introduce the
single spatial mode, laser radiation. Measurements of the har- slowly varying spatial envelopes for the harmonic field and
monics’ far-field profiles have indicated that the harmonics the polarization into (2). These are given by
largely preserve the low divergence, Gaussian character of the
laser radiation [33–35]. However, a more fundamental ques-  Zz 
 
tion is whether the high-order harmonics preserve the high aq (x, t) = Aq (x, t) exp −i qk0 (x, t) + ∆k(x, t) dz 0 ,
spatial coherence of the fundamental laser radiation.
−∞
Until recently, the spatial coherence of the harmonics had
(3)
not been measured [36]. While a number of previous experi-
mental studies have characterized the far-field profiles of the  Zz 
high-order harmonics, knowledge of the far-field profile alone pq (x, t) = Pq (x, t) exp −i qk0(x, t) dz 0 , (4)
does not indicate the actual transverse spatial coherence of −∞
the radiation. Measuring the transverse coherence requires
performing some manner of interference experiment such as where k0 is the wave number of the fundamental laser field,
a Young’s two-slit experiment. Previously, such techniques and the phase mismatch of the harmonic with the laser field
have been applied to measure the spatial coherence of short is defined as ∆k ≡ kq − qk0. Finally, using the fact that
wavelength XUV [37] and soft X-ray lasers [38–40] as well
as laser-plasma X-ray sources [41] and low-order harmonics ∂ 2 aq ∂aq
from solid target plasmas [42].  kq , (5)
∂z 2 ∂z
In this article, we briefly review the equations governing
the propagation of the harmonic field in a nonlinear medium, we arrive at the paraxial wave equation for the qth harmonic:
and discuss the physics important in determining the har-
monics’ far-field profiles. This is followed by a brief review ∂aq 4πω2q
of the experiments that have examined the harmonics’ spa- ∇⊥
2
aq + 2ikq − 2kq ∆kaq + ikq Nσabs aq = − 2 pq .
∂z c
tial profiles. We then discuss the physics of spatial coherence (6)
in the context of harmonic generation. Finally, we present
a detailed experimental study of the harmonics’ spatial co- To derive this we have explicitly separated the real and imag-
herence. In particular, we have conducted a series of Young’s inary part of the harmonic wave number, where kq is the
315

real part of the harmonic wave number, σabs is the absorption with an intensity distribution given as
cross section for harmonic photons, and N is the gas density. " #
In deriving (6) we have ignored all terms that vary as ∇⊥ kq ; 1 2(x 2 + y2 )
this is equivalent to saying that ∂n q /∂x = ∂n q /∂y ≈ 0. In do- I(x, t) = I0 exp − 2
1 + 4z 2/k02 w40 w0 (1 + 4z 2/k02 w40 )
ing this we have ignored any refraction of the harmonic field 
by a spatially varying refractive index arising from the plas- × exp −4 ln 2t 2 /τFWHM
2
, (9)
ma formation. This is a very good approximation since short
wavelength radiation will be very resistant to refraction by (where w0 is the 1/e2 radius of the laser at focus), this phase
plasmas because of the large critical density associated with is
soft X-ray radiation. The refractive index of a low density
plasma is  2k0 (x 2 + y2 )z
ϕGaus(x) = − tan−1 2z/k0 w20 − 2 4 . (10)
k0 w0 + 4z 2
1 ne Finally we have included a third contribution to the phase,
nq ≈ 1 − , (7)
2 n crit ϕran (x, t), which accounts for randomly varying temporal
phase initially on the laser beam. It is this rapidly varying
phase that accounts for imperfect coherence of the laser beam
where n crit is the critical density (the density at which the and may subsequently degrade the coherence of the harmon-
plasma frequency equals the laser frequency) and n e is the ic.
electron density. For harmonics with wavelength < 1000 Å, In the tight focusing limit, namely, that case when the
n crit > 1023 cm−3 , while the average gas density is typically laser confocal parameter is smaller then the length of the
< 1019 cm−3 in most harmonic experiments. Thus n q ≈ 1 and medium, the geometric phase is very important. It is usually
the radiation is very insensitive to radial variations in elec- the limit on the conversion efficiency. The origin of this can
tron density; n crit for the laser, however, is ∼ 1021 cm−3 . Thus be seen in (10) because the phase of the harmonics produced
the radially varying phase imparted to the laser light may not on one side of the focus have a phase nearly 180◦ different
be negligible and refraction may have an effect on the spatial from those produced on the other side of the focus (due to
profile of the laser [43]. Because the phase of the laser is then the tan−1 term) and consequently destructively interfere. This
imparted to the harmonic through pq , the far-field profile and phase can give rise to rings in the harmonics’ far-field pro-
divergence of the harmonics can be affected. files from geometric phase interferences. However, because
The nonlinear polarization, pq , is the term that drives the of the much greater conversion efficiency when the laser is
harmonic field. The amplitude of the nonlinear polarization weakly focused (kw20  l), most experiments are conduct-
is proportional to the nonlinear dipole moment of the atoms ed in a regime where geometric phase effects are minimized
in the medium oscillating at the harmonic frequency times and it is the intensity dependent phase and the production
the density of the atoms in the medium. At low intensity, the of free electrons that are most important in controlling the
phase of the polarization will generally be q times the phase harmonics’ spatial properties.
of the laser. At higher intensity, the phase of the polariza- In addition to the phase contributions outlined above,
tion will be intensity dependent. If we explicitly separate the examination of (6) indicates that the value of the phase mis-
intensity dependence of the polarization from the traditional match ∆k contributes to the harmonics’ phase. Though there
geometric phase, however, we can write the following for the is some phase mismatch associated with the refractive in-
nonlinear polarization: dex of the neutral medium, a much greater contribution to
the phase mismatch is the production of free electrons by the
 laser field during the harmonic generation. Using (7) for the
pq (x, t) ∼
= 2N(x, t) dq [I(x, t)] exp iϕdip [I(x, t)]
 refractive index of the fields in a plasma, this phase mismatch
× exp iq[ϕGaus (x) + ϕran(x, t)] , (8) can be written (when q  1)

∆k ≈ (πq/λ0 )(n e /n crit ) , (11)


where |dq [I(x, t)]| is the amplitude of the intensity dependent
dipole moment oscillating at a frequency q. This dipole has where n crit is the plasma critical density at the laser wave-
been calculated numerically from the Schrödinger equation length. The importance of this term can be simply seen if we
by a number of groups [18, 44]. In a perturbative treatment it find the solution of (6) in the plane wave limit. (This limit is
is simply proportional to the qth power of the laser intensity, a good approximation if the laser is weakly focused through
but, in strong laser fields its dependence with laser intensity is the harmonic generating medium.) In this limit, ignoring the
usually slower than this [2]. geometric phase mismatch and the absorption cross section,
In (8), we have explicitly written the three different pos- the qth harmonic field exiting a uniform medium of length l
sible contributions to the phase of the nonlinear polarization, is
as it is these terms that have the greatest effects on the har-
monics’ profiles and coherence. ϕdip [I(x, t)] is the intensity Aq (t) ∼
= −2iπkq n 0l dq [A0 (x, t)] eiϕd sinc(∆kl/2)ei∆kl/2 .
dependence of the phase of the dipole, dq [I(x, t)]. This term (12)
has been shown to be very important in shaping the harmon-
ics’ far-field profiles and may be important in a determination From this expression it is clear that the production of free
of their coherence [45, 46]; ϕGaus(x) is the phase associated electrons not only limits the conversion efficiency through the
with a focused laser beam. For a focused Gaussian laser beam sinc term but also imparts phase on the harmonics. This phase
316

can alter the far-field profile [47]. Furthermore, if the time his- the focus. At any time there is a radial intensity profile, which
tory of the production of free electrons is not uniform across means that the harmonics are emitted with different phases at
the beam, a degradation of spatial coherence can result. different radial locations. The interference of the different ra-
dial locations causes the harmonic emission to be spread over
a larger angular region than predicted by perturbation theory.
2 Measurements of high-harmonic far field spatial
profiles

A number of experiments have been conducted to investi-


gate the spatial profiles of the harmonics [33–35]. In general
these experiments produced harmonics by focusing a laser
into the output of a gas jet or a flowing gas medium. Har-
monics produced in the medium are spectrally dispersed in
one dimension with a reflective or transmission grating and
the spatial profiles were measured in the direction perpen-
dicular to the spectral dispersion. This arrangement permits
simultaneous measurement of the spatial profiles on a range
of harmonics.
Tisch et al. reported a measurement of the profiles
from harmonics produced by a 1054 nm, 1 ps, Nd:glass
laser focused into low-density helium at intensities up to
3 × 1014 W/cm2 [33]. This experiment examined harmonics
as high as the 119th order (88 Å). Raw data from that ex-
periment are shown in Fig. 1a. Harmonics of increasing order
are produced toward the right side of these data. In general
this experiment found that the divergence of the harmon-
ic radiation decreased with increasing harmonic order. The
overall divergence of the harmonics tended to be larger then
the perturbation theory prediction for harmonics produced in
an unionized medium [i.e., when Iq (x) ∝ I0 (x)q ].
In addition, the harmonics in the plateau exhibited sub-
stantial structure. Figure 1b shows the lineout of the 71st har-
monic. The harmonic exhibits a smooth central feature that is
surrounded by sharp spikes. This behavior was attributed to
the phase imparted to the harmonic beam by the production
of free electrons. As the gas is ionized by the high intensi-
ty pulse, n e increases resulting in a commensurate increase in
∆k. Because the ionization history varies radially in the fo- a
cus, the exp{i∆kl/2} term imparts a rapidly varying radially
phase that translates to structure in the far field.
Experiments of Peatross et al. also observed structure in
the lower-order harmonics produced in gases of very low den-
sity [35, 45]. These experiments were conducted using a 1 ps
Nd:glass laser focused into a low-density tubular target con-
taining gas. The experiments were conducted in a density
regime in which free electron phase matching effects were
expected to be negligible.
Figure 2 shows the measured far-field pattern of the 11th
through 17th harmonics generated in 0.3 torr of Xe at an in-
tensity of 5 × 1013 W/cm2 [45]. The smooth, solid line shows
the angular pattern of the f/70 incident 1 µm laser. These
profiles are typical of those generated in that experiment, with
a narrow central feature and broad wings or shoulders. The
width of the central feature is consistent with that predicted
from a perturbative calculation where the order of the har-
monic production is of order 5. This is typical of calculations
of high-order harmonic generation [2]. Because of the low gas
density, good focal quality, and large f number (thin target) b
used in these experiments, the only apparent explanation of
Fig. 1. a Spatial profiles from harmonics produced by a 1054 nm,
the wing structure observed was the effect of the phase of the 1 ps, Nd:glass laser focused into low density helium at an intensity of
atomic dipole response to the driving field. For a thin target, 3 × 1014 W/cm2 (taken from [33]). b Lineout of the 71st harmonic from
the emission can be considered to come from a single plane at (a). The dashed line is the prediction of lowest-order perturbation theory
317

Fig. 2. Spatial profiles from harmonics produced by a 1054 nm, 1 ps,


Nd:glass laser focused into 0.3 torr of xenon at an intensity of 5 × 1013 W/
cm2 (taken from [35])

The spatial profiles from a 100 fs laser have been stud-


ied by Salières et al. [34, 48]. These experiments utilized
a 825 nm, 140 fs Cr:LiSrAlF6 laser focused with an f/25
lens into an argon or neon gas jet. These studies were con-
ducted in gas at higher density then the studies of [33]
and [35]. These experiments found that at low intensity, b
namely, that below the ionization saturation intensity (Isat ),
Fig. 3. a Spatial profiles from harmonics produced by a 825 nm,
the harmonics exhibited very smooth, single mode Gaus- 140 fs, Cr:LiSrAlF6 laser focused into a Ne gas jet at an intensity of
sian profiles. (The saturation intensity is the peak inten- 5 × 1014 W/cm2 (taken from [34]). b Lineouts of the 29th, 35th, and 41st
sity at which the ionization rate integrated over the laser harmonics from (a)
pulse is approximately two, i.e., when nearly 90% of the
atoms have been ionized by the end of the laser pulse.)
Figure 3a shows data taken in Ne under these conditions. the laser focus. At the lower intensity the profiles are very
Once again, the harmonics are spectrally dispersed in one smooth and nearly Gaussian. However, at the higher intensi-
axis and spatially resolved in the other. This data illustrates ty, the profiles exhibit some breakup from the refraction of the
that all the harmonics exhibit reasonably smooth profiles laser.
and that the divergence decreases with increasing order. Fig- Even more dramatic is a consideration of the two-
ure 3b shows lineouts of selected harmonics. These harmon- dimensional images of the harmonics from a refracted laser
ics exhibit a divergence of roughly 10–20 mrad. This is ap- beam. Figure 5 shows the images of harmonics produced
proximately the divergence predicted by perturbation theo- from a frequency doubled Nd:glass (526 nm) laser in a heli-
ry. um jet focused with an f/50 lens. These images were taken
At higher density and peak laser intensity above the ion- on the laser axis with an X-ray CCD detector. The detector
ization threshold of the gas, the smooth single mode spatial was located 250 cm from the laser focus. A 7500 Å thick alu-
profiles of the harmonics is no longer preserved [49]. This is minum filter blocked the laser light and passed the soft X-ray
largely due to the refraction of the fundamental beam. This re- radiation. The images are essentially a sum of the harmon-
fraction imparts phase distortion on the fundamental, which ic profiles of the harmonics that fall between the roll-over of
is transferred to the harmonic through (8). This effect is illus- the CCD detector on the long wavelength side at ∼ 350 Å
trated in Fig. 4 in which the profiles of harmonics produced and the cutoff of the aluminum filter at 173 Å on the short
in argon at intensities corresponding to ∼ Isat (Fig. 4a) and wavelength side (i.e., the 15th through the 31st harmonics.)
an intensity ∼ 2 × Isat (Fig. 4b) are compared. In these im- The first image is of harmonics produced at 2 × 1015 W/cm2
ages, the spectra are dispersed horizontally and the spatial (i.e., ∼ Isat ) the second and third images are at intensities
distribution is in the vertical direction. The images are of the of 5 × 1015 W/cm2 and 8 × 1015 W/cm2 , respectively (i.e.,
19th through the 29th harmonic and were taken 75 cm from > Isat ). The harmonics exhibit a clean, single mode Gaussian
318

3 Spatial coherence of harmonic radiation

Though the spatial profile studies presented in the previous


section have been very important in unraveling much of the
physics controlling the harmonic generation process, as men-
tioned earlier, they do not contain all pertinent information
about the spatial properties. Of equal importance is the spa-
tial coherence of the harmonic beam, or in other words how
well two parts of the harmonic profile interfere with each
other. The smooth, single mode character of the harmonics
implies that the good coherence of the laser beam is probably
a preserved; however, an interferometric based measurement is
required to quantify it. Because of the more difficult nature of
such experiments, until recently no data existed to help quan-
tify the spatial coherence of the harmonics in the XUV and
soft X-ray region [36].
The simplest interferometric measurement possible is the
well-known Young’s two slit experiment. Before we discuss
the physics affecting the coherence of the harmonics, it is use-
ful to briefly review the mathematics of spatial coherence and
place that in the context of a two-slit experiment.

3.1 Mathematical treatment of coherence


b The coherence of any light field can usually be quantified by
Fig. 4. a Spatial profiles from harmonics produced by a 825 nm, 140 fs, the mutual coherence function [50]. This function is defined
Cr:LiSrAlF6 laser focused into an Ar gas jet of density ∼ 5 × 1018 cm−3 at as
an intensity of 4 × 1014 W/cm2 . b Spatial profiles of harmonics produced
in Ar when the peak intensity is increased to 1 × 1015 W/cm2

Γ12 = a(P1 , t + τ)a∗ (P2 , t) , (13)

where a is the complex electric field amplitude of the field


at points P1 and P2 . The brackets denote an ensemble aver-
age. For all light fields we are concerned with, the ensemble
average can be taken to mean a time average over some ap-
propriate time interval. For an ultrashort pulse, such as a har-
monic, this time interval will almost always be the entire
pulse width [51]. The mutual coherence function is essential-
ly the correlation function between two points in the field of
the light source delayed with respect to each other by a time t.
a b c Note that Γii (0) is the intensity of the light source at point Pi .
Fig. 5a–c. Images of harmonics in the 173 Å to 350 Å range produced from It is usually convenient to work with the normalized quantity
a frequency doubled Nd:glass (526 nm) laser in a helium jet focused with usually called the complex degree of coherence, which is
an f/50 lens at an intensity of (a) 2 × 1015 W/cm2 , (b) 4 × 1015 W/cm2 ,
and (c) 8 × 1015 W/cm2 Γ12 (τ)
γ12 (τ) = √ , (14)
Γ11 (0)Γ22 (0)
where γ satisfies the relation
shape at the lower intensity but exhibit severe distortion at
higher intensities. 0 ≤ |γ12(τ)| ≤ 1 . (15)
It is this refraction that sets an upper limit on the use-
ful conversion efficiency that is achievable with high-order If γ is one, the field exhibits complete correlation between the
harmonic radiation [49]. Though the harmonic yield scales points 1 and 2 at time delay of t and is said to be fully co-
as the square of the gas density, implying that increasing the herent. We are most concerned with the coherence of the field
density is advantageous in achieving high yield, increasing with no delay between the two points. We will consequently
the density and the laser intensity begins to cause a degra- refer to the complex coherence factor (CCF)
dation in the harmonics’ profiles. This may be of concern
for some applications. Equation (7) suggests that higher elec- µ12 = γ12 (0) . (16)
tron densities can be tolerated for a short wavelength laser.
Therefore, the refraction effects will be more severe for the Next, we need to relate these quantities to the experimen-
fundamental of a near IR laser at 800 or 1000 nm than for tal setup, namely, the two-slit interference. The experimental
their second harmonic. This plasma induced phase can also configuration is generically shown in Fig. 6. We will assume
cause a degradation of the harmonics coherence. that the slits are very narrow compared to the harmonic beam
319

results from the introduction of a slit pair in the harmonics


beam.
From a practical standpoint, no slits used in any experi-
ment are infinitely thin. They must simply be thin enough to
assure that the diffraction from each slit is sufficient to assure
complete overlap of both diffracted patterns at the detector. In
practice this requires that the xz slit width, D, is D  λz/∆x.
For partially coherent light, it will be useful to compare
the measured dependence of the complex coherence factor
with a known case. It is well known that an extended source of
light, emitting incoherently over its surface, will exhibit some
level of partial spatial coherence when the coherence is meas-
ured a distance z from the source. The complex coherence
factor of an incoherent source can be quite easily calculated
using the well-known van Cittert–Zernike theorem [51]. This
Fig. 6. Generic experimental configuration of the Young’s two-slit experi- theorem states that the CCF of an incoherent source with an
ment to measure spatial coherence intensity function I(x) is given, far from the source as
RR  
I(ξ, η) exp i 2π
λz (∆x ξ + ∆y η) dξ dη
so that effectively only one point on the beam is sampled. If µ(x 1 ,y1 ; x 2,y2 ) = RR ,
I(ξ, η) dξ dη
a(P) is the harmonic field at the slit mask, then the harmonic (23)
field at the detector plane, Q, will be
where ξ and η are the spatial coordinates at the source. This
a(Q, t) = K 1 a(P1 , t − r1 /c) + K 2 a(P2 , t − r2 /c) , (17)
result, very similar to the result for Fraunhofer diffraction, in-
where the K i are constant amplitude factors, which depend on dicates that the CCF is simply the Fourier transform of the
the diffraction of the field from each slit. We note that the har- source function. For a uniform, circular disk with radius a, the
monic intensity detected if light passes through only one slit CCF measured between two points separated by ∆x and ∆y
will be is
D 2 E  p 
I j (Q) = |K j |2 a (Pj , t − r j /c) . (18) J1 2πaλz ∆x 2 + ∆y2
µ(∆x, ∆y) = 2 p , (24)
2πa
λz ∆x 2 + ∆y2
Consequently, the intensity detected when both slits are illu-
minated is where J1 (ζ) is the Bessel function of the first kind, order 1.
  
(1) (2) r2 − r1 We can further quantify the coherence of a light source
I(Q) = I (Q) + I (Q) + 2K 1 K 2 Re Γ12 . through the definition of its coherence area. The coherence
c
(19) area is defined as [51]
ZZ

Examining (14), we note that the complex degree of coher- Ac = µ(∆x, ∆y) d∆x d∆y . (25)
ence can always be written as γ12 (τ) = |γ12 |(τ) exp[−iωq τ −
iϕ(τ)]. Thus we see that the intensity observed, given by (19),
of course oscillates as r1 −r2 varies. If the detector is far from An incoherent disk of diameter d will have a coherence area
the slits, then the fringes have a spatial period, L, of of Ac = 4λ2 z 2 /πd 2 measured at a distance z from the source.

λr
L= . (20) 3.2 Factors affecting the coherence of the harmonics
∆x
If we then define the fringe visibility on the basis of the max- With these mathematical preliminaries in mind, we are in
imum and the minimum intensities of this intensity pattern a position to consider the factors that can degrade the co-
as herence of the harmonics. We point out initially that if the
Imax − Imin fundamental laser is beam is completely coherent, in the ab-
V= (21) sence of any phase variations imparted on the harmonic, the
Imax + Imin harmonics will also exhibit full coherence. Their phase is
(where Imax and Imin the maximum and minimum intensities simply q times the phase of the laser; so, if the laser phase
of the fringe pattern), then the fringe visibility at the center of is completely correlated over all space, the harmonic’s phase
the pattern is simply will also exhibit complete correlation.
√ However, as we have seen there are a number of factors
2 I (1) I (2) that alter the phase of the harmonics during their generation.
V = (1) γ12 (0) . (22) If these factors impart a time-varying phase that differs from
I + I (2)
one point on the laser beam to the next, then the normalized
We can therefore obtain a direct measurement of the complex time averaged correlation of (14) is less than one. There are
coherence factor by a measurement of the fringe visibility that a couple of factors that can do this.
320

First, from (8), we see that the harmonic has a phase as- If the linearly increasing electron density between points x1
sociated with the intensity dependent dipole. If the intensity and x2 differs by the amount δn e at the end of the pulse, the
at two points on the laser beam are equal, then the time his- electron density can be written
tory of the dipole phase will be the same at these two points. (
If the CCF is then calculated between these two points the (n e0 + δn e ) τtp , i = 1
n e (t) =
i
(27)
phases will cancel and the coherence is not affected. On the n e0 τtp , i = 2
other hand, if the peak intensities differ slightly, the phase his-
tories at the two points on the beam in question will not be and the complex coherence factor is
exactly the same. This will give rise to interference in the time    
integral of (13) and may cause a decrease in coherence. πql t
µ12 = exp −i δn e . (28)
By this reasoning, we note that if the two-slit experiment 2λ0 n crit τp
is performed on the harmonic beam with the slits placed sym-
metrically about the beam center and the laser beam intensi- Time integration yields for the absolute value of the CCF:
ty is radially symmetric, we expect that the fringe visibility  
πql
will not be degraded. On the other hand, if the dipole phase |µ12 | = sinc δn e . (29)
does exhibit significant variation with intensity, as conjec- 4λ0 n crit
tured in [46] and [45], the CCF measured on two points not The maximum electron density at which harmonics will be
radially symmetric will exhibit a decrease in the CCF from produced is that at which ∆kl ∼ 2π, implying a maximum
unity as a result of the differing time histories of the phase. electron density of approximately 1–5 × 1017 cm−3 for the
Another contribution to the phase of the harmonics is the harmonics in the 500–100 Å range. It is reasonable, therefore,
∆k term as manifested in (12). As already discussed, this term to assume that the electron density can fluctuate by an amount
is largely due to the production of free electrons by ionization that is comparable to this value. Equation (29) implies that
when the laser intensity is sufficiently high. Thus, if there are electron density variations of this magnitude will degrade the
any density fluctuations or fluctuations in the ionization time coherence to µ ∼ 0.9–0.6, a significant decrease in harmonic
history between two points, the time history of n e , and conse- coherence. Equation (29) also implies that the coherence will
quently ∆k will differ. This will also result in a decrease in the be further degraded as q increases.
CCF upon time integration in (13). This physics is illustrat- This coherence degradation by the creation of free elec-
ed schematically in Fig. 7. Consider two points on the beam trons is not due to the phase fluctuations imparted directly on
generated by equal peak intensities. If there is a slight densi- the harmonic by refractive index changes, fluctuations that are
ty difference between the two points, the time history of the very small for short wavelength harmonic light in low densi-
electron density will differ slightly, as shown in Fig. 7. This ty plasma, but is due to the phase fluctuations placed on the
electron density time history translates into a phase time his- laser beam, which are then transferred to the harmonic. Be-
tory. The phases between the two points do not cancel and cause the phase of the harmonic is q times that of the laser’s,
will affect the time average correlation integral. for high orders, it is possible for a very small degradation in
We can further quantify this effect by making a estimation the laser’s coherence to result in a sizable degradation on the
for the harmonic pulse shape. We can derive a simple scal- harmonic. This is a general statement, independent of the ac-
ing for the complex coherence factor if we assume that the tual mechanism by which the laser’s coherence is originally
harmonic pulse is square in time and that the electron densi- altered.
ty ramps up linearly over the harmonic pulse. In this case, we This fact can be simply illustrated if we consider two
approximate the harmonic pulse as points on a laser beam with small phase fluctuations between
them. The laser field at the two points 1 and 2 can be written:
Aq (xi , t) = Aq0 exp{−i∆kl/2}  
  A1,2 = E 0 (t) exp iωt − ikx + iϕ1,2 (t) , (30)
πqn ie (t)
= Aq0 exp −i l , 0 ≤ t ≤ τp , (26)
2λ0 n crit which means that the CCF of the laser is
hA1 A∗2 i
µ1,2 (ω) = p
|A |2 |A2 |2

1 
= exp i(ϕ1 − ϕ2 ) , (31)
Zτp
 
= τp−1 exp i(ϕ1 − ϕ2) dt , (32)
0

for a square pulse, where we have simply averaged over the


laser pulse. (τp is the laser pulse width). Since we assume that
the phase fluctuations are small, we can say that the absolute
value of the CCF can be approximated as
Zτp
Fig. 7. General explanation of how a variation in density across the laser
beam can affect the harmonic’s spatial coherence. If the density between µ12 (ω) ≈ 1 − 1 τ −1 ∆ϕ dt . (33)
2 p
two points in the beam is slightly different, as the medium is ionized, the
electron density will follow the thick solid and dashed curves 0
321

The harmonic field generated by this laser beam at points 1 the pulses to their original pulse width of 2 ps. The result-
and 2 will be approximately ing pulse energy produced is up to 0.5 J. The beam profile
  after recompression is near Gaussian with a 1/ e2 diameter of
A1,2(q) = E q (t) exp iqωt − iqkx + iqϕ1,2 (t) , (34) 35 mm.
The laser pulses were then frequency doubled in a 1 cm
which means, by the same reasoning, that the magnitude of thick KDP crystal to a wavelength of 526 nm with a con-
the harmonic’s CCF is version efficiency of ∼ 45%. The high-order harmonics were
Zτp produced with 526 nm light instead of 1053 nm light because
the shorter wavelength drive has been shown to yield high-
µ12 (qω) ≈ 1 − q τ −1 ∆ϕ dt . (35)
2 p er conversion efficiency into harmonics in the 200–500 Å
0 range [23]. As discussed above, it is also expected to be less
susceptible to free electron phase fluctuations. Furthermore,
(now assuming that q∆ϕ is small, which is equivalent to say- the shorter wavelength laser light is also less susceptible to
ing that the laser’s CCF is very close to one). So we see that ionization induced refraction effects within the gas medium.
the CCF of the harmonic and the laser are related by The Gaussian laser beam spatial profile was apertured to
 a diameter of 1.5 cm immediately prior to the focusing lens
µ12 (qω) ≈ 1 − q 1 − µ12 (ω) . (36) to produce a uniform, near flattop profile. This beam was
then focused with a 75 cm focal length, plano-convex lens.
Thus the deviation of the harmonics CCF from unity will be q
The focal spot of the laser light was measured to be a di-
times larger than that of the laser. This reasoning implies that
ameter of 70 µm, indicating that it is very near the diffrac-
from the standpoint of preserving the inherent high coherence
tion limit. The laser pulses were focused into a gas plume
of the laser in harmonic generation, it may be advantageous
produced by a pulsed gas jet that could be backed with up
to use short wavelength laser light to reach short wavelength
to 50 bar of pressure (yielding an estimated gas density of
harmonics since q can be kept small.
5 × 1018 cm−3 [52]). The high-order harmonics produced in
the 200 Å to 500 Å range were detected with a flat-field soft
4 Measurement of high-order harmonic spatial X-ray spectrometer located on the laser axis. This spectrom-
coherence eter utilized a horizontal 50 µm entrance slit. The harmonic
radiation was dispersed with a flat field, grazing incidence,
4.1 Experimental Details gold-coated 1200 line/mm grating. The soft X-rays were then
detected with a Cs:I coated, two-stage microchannel plate that
To measure the coherence of XUV radiation produced by was coupled to a phosphor readout. The signal was then col-
high-order harmonic generation, we executed a Young’s two- lected with a lens and a CCD camera. The total distance from
slit experiment. The experimental configuration is shown in harmonic source (laser focus) to the entrance slit was 130 cm
Fig. 8. The majority of the data described in the sections that and the distance from source to detector was 180 cm. Because
follow were of harmonics generated in helium gas. This gas the grating focuses only in the dispersing direction and does
was chosen because the harmonics produced reach to higher not alter the divergence of the radiation in the transverse di-
order and shorter wavelength then the other gases, a feature rection, this configuration permits simultaneous measure of
essentially due to the high ionization potential of helium. Ex- spectral information on one axis, with a measure of the spatial
periments were also performed in other gases with similar profile of each harmonic on the transverse axis.
results.
In our experiment, harmonics were generated with a Nd: 4.2 Raw data
glass laser based on the well-known technique of chirped
pulse amplification. In brief, the laser consists of a diode A typical spectrum of harmonic radiation collected with
pumped, additive pulse mode-locked Nd:YLF oscillator, this configuration is shown in Fig. 9. This spectrum shows
whose 1053 nm, 2 ps pulses are stretched to 500 ps with the spatially integrated spectrum of high-order harmon-
a dispersive grating/lens stretcher. These pulses are ampli- ics produced with 526 nm light in He at an intensity of
fied in a Nd:glass regenerative amplifier and three additional 4 × 1015 W/cm2 . The spectrum exhibits the classic high-
amplifiers to an energy of 1 J. A grating pair recompresses order harmonic behavior, the production of a plateau of

Fig. 8. Experimental configuration used to


measure the high-order harmonics’ spatial co-
herence
322

slits in 20 µm thick Ti foil. Each slit had a width of 8 ± 1 µm


and slit pairs with spacings of 28, 50, 75 and 100 µm were
used for the measurements. Because of the proximity of the
slit pairs to the laser focus (the laser was roughly 1 mm in di-
ameter at the slits), the slits began to show laser damage after
10 to 20 full-power laser shots. To monitor this degradation,
the slits were backlit at a slight angle with a HeNe laser and
the image of the slits was imaged outside the vacuum cham-

Fig. 9. Spatially integrated spectrum of high-order harmonics produced with


526 nm light in He at an intensity of 4 × 1015 W/cm2

Fig. 11. Measured profile of the 13th harmonic at the location of the dou-
ble slits (4 cm from focus) generated at an intensity of 4 × 1015 W/cm2 .
This was measured by scanning a single slit across the harmonic profile.
The solid line is a Gaussian fit to the data

Fig. 10. Angularly resolved spectrum of high-order harmonics produced


with 526 nm light in He at an intensity of 4 × 1015 W/cm2

harmonics to a high order with a sharp cutoff below a cer-


tain wavelength (in this case occurring at the 23rd harmonic
λ ≈ 229 Å). The harmonics produced under these condi-
tions exhibit smooth, low divergence spatial profiles. A two-
dimensional raw image of harmonics produced under the
same conditions as Fig. 9 is shown in Fig. 10. The harmon-
ics are dispersed in the vertical direction; their far-field spatial
profile is manifested in the horizontal profile. These harmon-
ics clearly exhibit smooth, single mode Gaussian far-field
profiles, similar to the results reported in [34].
To measure the spatial coherence, we introduced slit pairs
into the harmonic’s beam before the spectrometer (see Fig. 8)
to generate interference fringe patterns on each harmonic. Be-
cause of the short wavelength of the harmonics, it is necessary
to use open slits to produce the interference fringe pattern.
These slits need to be sufficiently narrow to ensure that only
a small portion of the harmonic beam is sampled by each
slit. Furthermore, we require that each slit be narrow enough
to assure that the diffraction of the beam from each slit is
fast enough to yield complete overlap of the radiation pass-
ing both slits at the detector. This implies that the slits need to
be no wider than 10 µm. On the other hand, the slits must be
wide enough to pass enough radiation for the fringe pattern to
be detectable on each laser shot.
With these considerations, we found that the optimal con-
figuration was to use slit pairs of roughly 8 µm width placed
4 cm from the gas jet plume and laser focus. The slit pairs Fig. 12. CCD image of fringes generated on the 11th to the 19th harmonic
used in the experiment were produced by laser drilling the in helium with a peak intensity of 6 × 1015 W/cm2
323

ber with magnification. During the experiment when the slit


exhibited some damage, it was replaced by an unused slit foil
in the position of the original slits. The slit pairs were placed
perpendicular to the 50 µm entrance slit of the spectrome-
ter. The total distance from the slit pair to the detector was
176 cm. This distance assured that the harmonic fringe spac-
ing (> 600–1000 µm) was larger than the estimated 100 µm
spatial resolution of the detector.
The profile of the harmonics at the position of the slit pairs
was measured by placing a single 8 µm slit at the same posi-
tion as the slit pairs and scanning the slit across the beam. The
measured profile of the 13th harmonic generated at an intensi-
ty of 4 × 1015 W/cm2 is shown in Fig. 11. This measurement
is characteristic of all of the harmonics, which, in general,
have a diameter of roughly 350 µm.
A typical raw image of a fringe pattern generated in the Fig. 14. Comparison of fringes produced by the laser with a 50 µm slit pair
configuration of Fig. 8 by a slit pair with a 50 µm spacing when the laser propagates only through vacuum at its focus and when the
laser propagates through the ionizing gas jet at focus. Both fringe patterns
placed in a beam of harmonics is shown in Fig. 12. This is da- were produced when the laser peak intensity at focus was 5 × 1015 W/cm2 .
ta of harmonics generated at an intensity of 6 × 1015 W/cm2 (The gas jet backing pressure was 50 bar)
in helium. In this image, fringes on all the harmonics (the
11th to the 19th) are clearly visible. All the harmonics exhibit
some coherence under these conditions. From an analysis of patterns were generated with the same slits used for the har-
such images it is possible to determine the harmonic fringe monics; however the laser beam was reflected out of the
visibility under a variety of conditions. chamber after the slit pairs and detected with a CCD cam-
It is also important to note that the coherence of the laser era. Figure 13 shows the laser fringes in vacuum (no gas jet)
used to generate these harmonics is very good. Fig. 13 shows with three slit spacings. The fringe visibility is > 0.9 for all
lineouts of fringes generated by the laser itself. These fringe three slit separation. We also note that propagation of the laser
through the gas jet does not significantly degrade the coher-
ence of the laser. Figure 14 compares the fringe pattern of
the laser generated with 50 µm slit separation when the laser
propagates only in vacuum inside the chamber and the fringe
pattern at the same laser intensity with the gas jet pulsed, so
that the laser has propagated through the harmonic generating
medium. As can be seen in this figure, the laser’s fringes do
not change dramatically upon propagation through the gas jet.

4.3 Coherence measurement results

To derive information about the harmonics mutual coher-


ence function and how it changes with changing parameters,
we have examined the fringe visibility of the harmonics as
a function of a number of parameters. Lineouts of typical in-
terference patterns obtained by using slits with a 50 µm spac-
ing centered on the harmonic beam are shown in Fig. 15. Here
fringe patterns on the 11th to the 19th harmonic are shown
(covering the wavelength range of 277 to 479 Å). These data
were taken with a peak laser intensity of ∼ 4 × 1015 W/cm2 ,
an intensity above the ionization saturation intensity in heli-
um (∼ 1 × 1015 W/cm2 ). As shown above in Fig. 12, all the
harmonics exhibit interference fringes. However, the fringe
visibility is < 1.0, indicating that all the harmonics exhibit
some degree of partial coherence over a separation of 50 µm.
In fact, the fringe visibility decreases with increasing
harmonic order. Figure 16 shows the measured visibility
with a 50 µm slit spacing as a function of harmonic order
from the 11th to the 19th harmonic for two peak intensities:
1 × 1015 W/cm2 and 4 × 1015 W/cm2 . Each point in this fig-
ure represents the average of 6 laser shots within a ±10%
energy bin. The error bars in the measurement were deter-
Fig. 13. Lineout of fringes generated by the 526 nm laser beam with three
different slit separations. The fringes were generated with the same slit mined by the extent of the shot to shot variation of the fringe
pairs used for the harmonics in vacuum, 4 cm from the laser focus with the visibility. At the lowest intensity, the fringe visibility is virtu-
detector located 200 cm from the slit pairs ally constant over the harmonic order considered here, though
324

Fig. 17. Images and the resulting lineouts of fringes from the 11th harmonic
(λ = 479 Å) generated with slits of 50 um separation at two different peak
intensities

Fig. 15. Lineouts of typical interference patterns obtained using slits with Fig. 18. Fringe visibility of the 15th harmonic as a function of laser peak
a 50 µm spacing centered on the harmonic beam produced with an intensity intensity
of 4 × 1015 W/cm2 and a gas jet backing pressure of 50 bar

of 0.8. In contrast, at the higher intensity of 5 × 1015 W/cm2 ,


an intensity at which significant ionization has occurred in
the helium [53], the fringes are broadened, with a drop in
visibility to 0.45.
This trend is reflected over the intensity range from the
mid 1014 W/cm2 to the mid 1015 W/cm2 , the range over
which helium ionization becomes important. Figure 18 illus-
trates the visibility for fringes of the 15th harmonic (λ =
351 Å) with 50 µm spaced slits as a function of intensity. Be-
low the ionization saturation intensity the harmonic exhibits
good fringe visibility of between 0.7 and 0.8. This visibility
falls and levels off at about 0.4 at the highest intensity. The
Fig. 16. Fringe visibility as a function of harmonic order for two different
peak intensities
drop in fringe visibility illustrated here corresponds closely
to the onset of significant ionization [53]. This data also indi-
cates that at low intensity, the coherence of the harmonics can
there is a slight decrease in visibility as the order increases: be quite high, nearly that of the laser itself with the same slit
the visibility is 0.6 for the 11th harmonic while it drops to separation.
0.45 for the 19th harmonic. At the higher intensity, however, One interesting trend of the fringe visibility is illustrated
the variation of the fringe visibility with increasing harmonic in Fig. 19. Here is shown the fringe visibility of the 13th har-
order is more pronounced, falling faster with increasing order monic with a 50 µm slit spacing as a function of the position
than the lower intensity harmonics. Here the visibility on the of the center of the slit pair across the harmonic beam. The
11th harmonic is 0.45 and the visibility drops to 0.18 on the peak laser intensity is 5 × 1015 W/cm2 , an intensity above
19th harmonic. ionization saturation and a regime where the fringe visibili-
Figure 16 also indicates that there is a variation in the har- ty is degraded. In the center of the beam, the fringe visibility
monic fringe visibility with laser peak intensity. This effect is 0.4. However, as the slit pair is moved out to larger ra-
is illustrated in Fig. 17 where the images and the resulting dius, the visibility increases. At a radius of 200 µm, near the
lineouts of the fringes from the 11th harmonic (λ = 479 Å) edge of the harmonic beam (see Fig. 11), the fringe visibili-
generated with slits of 50 µm separation are shown at two dif- ty has increased to nearly 0.7. This trend is reproduced in all
ferent peak intensities. In both cases the slits were centered of the harmonics studied and is also evident when larger slit
on the beam. The first image shows the fringes generated spacings are used.
with a peak intensity of 8 × 1014 W/cm2 , below the onset Furthermore, we find that, at the high backing pressures
of significant ionization in the helium gas. Here the fringes used in our experiments, the fringe visibility does not sig-
are sharp and well defined with a corresponding visibility nificantly change with changing backing pressure. Figure 20
325

Fig. 19. Fringe visibility of the 13th harmonic with a 50 µm slit spacing as Fig. 21a,b. Fringe visibility of the 13th harmonic (a) and the 19th harmonic
a function of the position of the center of the slit pair across the harmonic (b) as a function of slit spacing when the slits are centered on the harmonic
beam. (The peak laser intensity is 5 × 1015 W/cm2 ) beam. The short dashed lines are the calculation of the CCF as a function
of separation of uniform incoherent disks of 15 and 16 µm diameter, respec-
tively. The long dashed line is the result of a calculation assuming a 65 µm
diameter incoherent disk

resolution prevented an accurate measurement of the visibil-


ity on the 19th harmonic with a 100 µm slit spacing.) From
this measurement (which essentially gives the complex co-
herence factor µ(∆x, ∆y) as a function of spacing ∆x), we
can use (25) to estimate the effective coherence area Ac . If
we assume that the CCF is azimuthally symmetric, we find
that the coherence area of the two beams are 1.5 × 10−4 cm2
and 6.1 × 10−5 cm2 for the 13th and 19th harmonics, respec-
tively. The coherence area can then be related to an effective
incoherent source size using the van Cittert–Zernike theo-
rem. Using the coherence area of a uniform disk, the effective
incoherent source size is
Fig. 20. Fringe visibility of the 15th harmonic through a 50 µm slit spacing
at an intensity of 6 × 1015 W/cm2 as a function of jet backing pressure with 2λz
ds = √ (37)
the slits centered πAc
Here z is the distance from source to slits. This quantity is,
shows the fringe visibility of the 15th harmonic through in other words, the diameter of a uniform, incoherent source,
a 50 µm slit spacing at an intensity of 6 × 1015 W/cm2 as which yields the same coherence area as that measured. Thus
a function of jet backing pressure with the slits centered. At a smaller effective source size indicates a higher degree of co-
these pressures the gas density in the gas jet scales linear- herence. Using this definition, we find that the 13th and 19th
ly with gas backing pressure, so Fig. 20 essentially shows harmonics exhibit roughly similar coherence source sizes.
the fringe visibility as a function of gas density. The fringe The 13th harmonic exhibits an effective incoherent source
visibility is roughly constant at a value of around 0.3 over diameter of 15 µm, while the 19th has a diameter of 16 µm.
the pressure range studied. (The decrease in signal at lower For comparison, the modulus of the complex coherence
backing pressures precluded an accurate measurement of the factor of a uniform incoherent disk emitting radiation at
visibility below 6 bar.) a wavelength of 405 Å located at the laser focus with a diame-
Despite the decrease of coherence with increasing laser ter of 15 µm calculated using the van Cittert–Zernike theorem
intensity, we find that the spatial coherence of the harmonics is plotted on the data for the 13th harmonic and a similar
is still high even at the high intensities. It is possible to recon- plot for a 277 Å source with a diameter of 16 µm is com-
struct the complex coherence factor and the effective coher- pared to the data for the 19th harmonic. Clearly this simple
ence area of the harmonics by measuring the fringe visibility model gives a reasonable estimation for the observed charac-
as a function of slit spacing. Such measurements at an inten- ter of the complex coherence factor. The calculated coherence
sity of 4 × 1015 W/cm2 on the 13th harmonic (λ = 405 Å) are factor of a 65 µm source is also shown on these data in this
shown in Fig. 21a. The same measurement on the 19th har- figure; 65 µm is the estimated size of the harmonic emission
monic (λ = 277 Å) is shown in Fig. 21b. (Inadequate spatial exiting the gas medium as calculated by a numerical solution
326

ence drop of 0.95, which is within the error of the measure-


ments of Fig. 14.
The electron density fluctuations may be the result of two
factors. First, intensity variations across the laser focus, per-
haps induced by refraction, could cause varying ionization
histories over the harmonic beam. Similarly, density fluctua-
tions in the initial gas density could also give rise to a time-
dependent electron density difference between two points (as
illustrated in Fig. 7).
To further illustrate this point, we have performed numeri-
cal calculations to derive the harmonic field in the presence of
initial gas density variations. This model involves solving the
Fig. 22. Fringe pattern on the 9th harmonic produced in Ar with 50 µm slits
with a peak intensity of 1 × 1015 W/cm2 wave equation for the harmonic field in the presence of ion-
ization and has been described previously in detail [54]. For
illustrative purposes, the calculated complex coherence fac-
of the wave equation for the 13th harmonic [54]. Our calcula- tor of the 15th harmonic as it exits the medium as a function
tions indicate that the actual harmonic source size is, in fact, of spacing is shown in Fig. 23 for two intensities. This cal-
roughly equal to the laser spot source size (∼ 70 µm, 1/e2 di- culation assumes a 1 ps, 527 nm pulse in 1 mm of helium gas
ameter) despite the nonlinearity of the harmonic generation with a mean density of 5 × 1018 cm−3 . A 30% initial density
process because the harmonic yield is strongly saturated in ramp is placed across the 70 µm Gaussian focal spot in one
the center of the Gaussian focus profile. The comparison in spatial direction. (This density profile is chosen for illustra-
Fig. 21 clearly indicates that the observed harmonic’s coher- tive purposes only, but it is not unreasonable to assume that
ence is significantly better than a purely incoherent 65 µm gas density variations of this order can occur across the profile
source of radiation produced at the laser focus. of the gas jet output at the high backing pressure used in our
Though all the studies presented thus far have been with experiment [55].) As shown in Fig. 23, with a peak intensity
harmonics produced in helium, we also note that good co- of 5 × 1014 W/cm2 , the harmonic exhibits near perfect coher-
herence has been observed on harmonics from other gases as ence across the beam. At an intensity of 3 × 1015 W/cm2 , the
well. Figure 22 shows the fringe pattern on the 9th harmon- coherence is clearly degraded.
ic produced in Ar with 50 µm slits under conditions similar to Though the calculated coherence factor does not exactly
those of the helium harmonics (Intensity = 1 × 1015 W/cm2 ). match the observed decrease in coherence, we emphasize that
The fringe visibility is very good with µ = 0.8. this calculation is meant only for qualitative estimates. For
example, our model does not include laser beam refraction
and filamentation in the plasma, effects which are, in gener-
5 Discussion of the coherence results al, difficult to model accurately. Nonetheless, it is clear from
this analysis that spatial nonuniformities in the free electron
The measured visibility of the harmonics indicates that the formation can have a dramatic effect on the spatial coherence
coherence of the harmonics is quite high. However, it is also of the harmonics.
clear that it is not perfect, particularly at high intensity. This Though the fringe visibility does not vary with gas jet
is interesting in light of the fact that the laser exhibits near- backing pressure, this does not necessarily contradict the hy-
ly perfect coherence, even upon passage through the gas jet. pothesis that free electrons are the cause of the coherence
Much of these observations can be explained in the context of degradation. Though the background gas density varies as
the discussion of Sect. 3. the backing pressure is changed, it should be noted that the
The most striking trend in the harmonics’ fringe visibil- maximum electron density at which harmonics are produced
ity is the decrease of the visibility with increasing intensity. is not necessarily determined by this maximum density. As
The observed partial degradation of coherence at high laser
intensity is, according to the data of Fig. 18, apparently cor-
related with the onset of ionization in the helium. This seems
to indicate that the formation of a plasma by ionization is the
primary cause of the observed decrease in coherence. This
seems to argues for the free electron mechanism discussed in
the context of (29). This model, and the simple scaling re-
lation, are also consistent with the observed decrease in the
visibility with increasing harmonic order. The observed fall of
the CCF from 0.8 to about 0.5 at the highest intensity for the
15th harmonic is completely consistent with the predictions
of (29).
This decrease in coherence is also consistent with the
measured coherence of the laser. As we discussed in Sect. 3, Fig. 23. Numerically calculated complex coherence factor of the 15th har-
the deviation of the harmonics CCF will be expected to be monic as it exits the medium as a function of spacing for two intensities
found by solution of the wave equation for the harmonic. This calculation
larger then that of the laser by roughly the harmonic order. assumes a 1 ps, 527 nm pulse in 1 mm of helium gas with a mean density
Thus the observed harmonic coherence of 0.5 at the highest of 5 × 1018 cm−3 . A 30% initial density ramp is placed across the 70 µm
intensity would imply that the laser need only have a coher- Gaussian focal spot in one spatial direction
327

illustrated by (12), the harmonic yield essentially drops to and larger radius. This ring is composed of an annulus that
zero when ∆kl ∼ 2π. This implies that the maximum elec- is much smaller then the total diameter of the beam. After
tron density at which harmonics will be produced is about the beam exits the medium, these annuli begin to spread by
1–5 × 1017 W/cm2 for the 11th to the 19th harmonics. At the diffraction.
high backing pressures used in our experiments, the average When the slits are placed symmetrically about the laser
background density was higher then 1018 cm−3 . As the gas focus, one slit is illuminated by one side of the ring, while
density is changed, the effective free electron density seen by the other is illuminated by the other side. When the slits are
the harmonics does not. Thus it is not inconsistent that the moved to one side, both slits are illuminated almost com-
fringe visibility does not change with backing pressure. pletely by the diffracted light from one side of the annulus.
As discussed in Sect. 3, the intensity-dependent dipole can Thus, when the slits are placed in the center, they essential-
also degrade the coherence. That this is not the dominant ly compare the coherence of the light produced over a large
physics is indicated by two factors. First, the decrease in co- distance, namely, the radius of the anulus. However, when
herence seems to be closely correlated with the onset of ion- the slits are placed to one side, they effectively sample only
ization in the helium. Any effects of the intensity-dependent light produced over a very narrow region. As a result, the ob-
dipole will be present in the absence of ionization. Further- served fringe visibility is high toward one side and lower in
more, if the intensity-dependent dipole were important, we the center.
would also expect that the observed visibility would decrease
with as the slit pair is moved to one side, away from the cen-
tral axis of the laser focus. In fact what is observed is the
opposite of this; the fringe visibility improves as the slits are 6 Conclusion
moved away from the center of the beam.
The reason for this trend in the data can probably be ex- In conclusion, we have examined recent results on the spatial
plained by the fact that our measurements were performed properties of short wavelength high-order harmonic radiation.
with the slit pairs placed a small distance down stream of the The harmonics represent a novel way of generating bright,
harmonic near field. This situation is illustrated schematically coherent soft X-ray radiation. They exhibit smooth, nearly
in Fig. 24. Though the slit pairs are placed within a confocal Gaussian spatial profiles, though they have some structure un-
parameter of the harmonics from the laser focus, some propa- der certain conditions. In general, there are a number of fac-
gation of the harmonic beam occurs before the slit mask. tors that can shape the harmonic’s spatial profile. The phase
When the laser peak intensity is above the ionization thresh- slip associated with a Gaussian beam can have some effect,
old, toward the temporal peak of the pulse, no harmonics are but with weakly focused beams this is usually not a prob-
produced in the center of the beam because of the production lem. More significant is the radially varying phase imparted
of free electrons. As a result, the harmonic emission at a giv- across the harmonic beam due to the intensity dependence
en time exits the medium as a ring whose thickness is much of the nonlinear polarizability. This has been shown to be
smaller than it radius (see [54] for a calculation of this ef- very important in the development of “wings” on the har-
fect). The radii of these rings increases later in the pulse as monic profile. Also important are the phase shifts associated
the intensity increases and ionization has occurred at larger with the free electron production by optical ionization. Not
only do free electrons impart phase to the harmonic beam
but they also cause refraction and breakup of the fundamen-
tal beam, which is manifested in the harmonics’ profiles. We
have also presented the first measurement of the spatial coher-
ence of high-order harmonic radiation in the soft X-ray region
of 270–480 Å. We find that the harmonics exhibit good coher-
ence, even at high intensity, though the coherence is degraded
by the onset of ionization in the medium. The magnitude of
the coherence degradation is consistent with some simple es-
timates and harmonic production calculations in the presence
Gas of density variations. We find that the harmonics exhibit an
effective incoherent source size that is approximately 15 µm
in diameter. The observed trends in the harmonics’ coherence
seem to be consistent with phase fluctuations imparted by
free electron production through ionization. However, there
may be some degradation effects from the phase differences
resulting from the intensity dependent dipole as well. Fu-
ture experiments will have to be conducted to determine the
details shaping the harmonics’ coherence.
Finally we note that the coherence characteristics of the
harmonics are substantially superior to those of previously
reported measurements of X-ray laser coherence, which typ-
ically exhibit an effective coherence source size of ∼ 100 µm
Fig. 24. Illustration of the physics giving rise to an increased fringe visibili- [38]. Thus the coherence area of the harmonics is nearly
ty as slit pairs are moved from the center of the harmonic beam to one side two orders of magnitude larger than soft X-ray lasers. These
of the beam results confirm that the harmonics are promising for inter-
328

ferometric applications requiring good coherence or high- 26. T.E. Glover, R.W. Schoenlein, A.H. Chin, C.V. Shank: Phys. Rev. Lett.
intensity applications demanding good, single mode beams. 76, 2468 (1996)
27. Y. Kobayashi, O. Yoshihara, Y. Nabekawa, K. Kondo, S. Watanabe:
Opt. Lett. 21, 417 (1996)
28. J.W.G. Tisch, T. Ditmire, D.D. Meyerhofer, N. Hay, M. Mason, M.H.R.
Acknowledgements. We thank P. Ruthven and A. Gregory for technical as- Hutchinson: Phys. Rev. Lett. submitted (1997)
sistance. This work was funded by the EPSRC and the Ministry of Defence. 29. Y. Liang, S. Augst, S.L. Chin, Y. Beaudoin, M. Chaker: J. Phys. B: At.
R.A. Smith is supported by an EPSRC fellowship, and D.D. Meyerhofer’s Mol. Opt. Phys. 27, 5119 (1994)
work is supported by the National Science Foundation and a NATO Collab- 30. D.J. Fraser, M.H.R. Hutchinson, J.P. Marangos, Y.L. Shao, J.W.G. Tisch,
orative Research Grant. M. Castillejo: J. Phys. B: At. Mol. Opt. Phys. 28, L739 (1995)
31. T.D. Donnelly, T. Ditmire, K. Neuman, M.D. Perry, R.W. Falcone:
Phys. Rev. Lett. 76, 2472 (1996)
32. S. Kohlweyer, G.D. Tsakiris, C.G. Wahlström, C. Tillman, I. Mercer:
Opt. Comm. 117, 431 (1995); D. von der Linde, T. Engers, G. Jenke,
References P. Agostini, G. Grillon, E. Nibbering, A. Mysyrowicz, A. Antonetti:
Phys. Rev. A 52, R25 (1995); P.A. Norreys, M. Zepf, S. Moustaizis,
1. A. McPherson, G. Gibson, H. Jara, U. Johann, T.S. Luk, I.A. McIntyre, A.P. Fews, J. Zhang, P. Lee, M. Bakarezos, C.N. Danson, A. Dyson,
K. Boyer, C.K. Rhodes: J. Opt. Soc. B 4, 595 (1987) P. Gibbon, P. Loukakos, D. Neely, F.N. Walsh, J.S. Wark, A.E. Dangor:
2. A. L’Huillier, L. Lompré, G. Mainfray, C. Manus: High-Order Har- Phys. Rev. Lett. 76, 1832 (1996)
monic Generation in Rare Gases, In Atoms in Intense Laser Fields, ed. 33. J.W.G. Tisch, R.A. Smith, J.E. Muffett, M. Ciarrocca, J.P. Marangos,
M. Gavrila (Academic Press, Boston 1992), pp. 139–202 M.H.R. Hutchinson: Phys. Rev. A 49, R28 (1994)
3. J.F. Reintjes: Nonlinear Optical Parametric Processes in Liquids and 34. P. Salières, T. Ditmire, K.S. Budil, M.D. Perry, A. L’Huillier: J. Phys.
Gases, (Academic Press, Orlando 1984) B: At. Mol. Opt. Phys. 27, L217 (1994)
4. A. L’Huillier, P. Balcou: Phys. Rev. Lett. 70, 774 (1993) 35. J. Peatross, D.D. Meyerhofer: Phys. Rev. A 51, R946 (1995)
5. S.J.J. Macklin, J.D. Kmetec, I.C.L. Gordon: Phys. Rev. Lett. 70, 766 36. T. Ditmire, E.T. Gumbrell, R.A. Smith, J.W.G. Tisch, D.D. Meyerhofer,
(1993) M.H.R. Hutchinson: Phys. Rev. Lett. 77, 4756 (1996)
6. J.K. Crane, M.D. Perry, S. Herman, R.W. Falcone: Opt. Lett. 17, 1256 37. M.H. Sher, S.J. Benerofe, J.F. Young, S.E. Harris: J. Opt. Soc. Am. B
(1992) 8, 114 (1991)
7. N. Sarukura, K. Hata, T. Adachi, R. Nodomi, M. Watanabe, S. Watan- 38. J.E. Trebes, K.A. Nugent, S. Mrowka, R.A. London, T.W. Barbee,
abe: Phys. Rev. A 43, 1669 (1991) M.R. Carter, J.A. Koch, B.J. MacGowan, D.L. Matthews, L.B. DaSilva,
8. Z. Chang, A. Rundquist, H. Wang, H. Kapteyn, M. Murnane, X. Liu, G.F. Stone, M.D. Feit: Phys. Rev. Lett. 68, 588 (1992)
B. Shan: Coherent, Tunable, X-Ray Emission at 5 nm Using High- 39. P. Celliers, F. Weber, L.B. DaSilva, J.T.W. Barbee, R. Cauble, A.S. Wan,
Harmonic Generation, In Applications of High Field and Short Wave- J.C. Moreno: Opt. Lett. 20, 1907 (1995)
length Sources VIl, Vol. 7, 1997 OSA Technical Digest Series (Optical 40. G. Cairns, C.L.S. Lewis, A.G. MacPhee, D. Neely, M. Holden, J. Kr-
Society of America, Washington, DC, 1997) ishnan, G.J. Tallents, M.H. Key, P.N. Norreys, C.G. Smith, J. Zhang,
9. R. Haight, D.R. Peale: Phys. Rev. Lett. 70, 3979 (1993) P.B. Holden, G.J. Pert, J. Plowes, S.A. Ramsden: Appl. Phys. B 58,
10. R. Haight, D.R. Peale: Rev. Sci. Instr. 65, 1853 (1994) 51(1994)
11. P. Balcou, P. Salières, K.S. Budil, T. Ditmire, M.D. Perry, A. L’Huillier: 41. I.C.E. Turcu, I.N. Ross, M.S. Schulz, H. Daido, G.J. Tallents, J. Krish-
Z. Phys. D, to be published (1995) nan, L. Dwivedi, A. Hening: J. Appl. Phys. 73, 8081 (1993)
12. J. Larsson, E. Mevel, R. Zerne, A. L’Huillier, C.-G. Wahlström, 42. J. Zhang, M. Zepf, P.A. Norreys, A.E. Dangor, M. Bakerezos, C.N. Dan-
S. Svanberg: J. Phys. B: At. Mol. Opt. Phys. 28, L53 (1995) son, A. Dyson, A.P. Fews, P. Gibbon, M.H. Key, P. Lee, P. Loukakos,
13. W. Theobald, R. Häßner, C. Wülker, R. Sauerbrey: Phys. Rev. Lett. 77, S. Moustaizis, D. Neely, F.N. Walsh, J.S. Wark: Phys. Rev. A 54, 1597
298 (1996) (1996)
14. P. Balcou, C. Cornaggia, A.S.L. Gomes, L.A. Lompré, A. L’Huillier: 43. R. Rankin, C.E. Capjack, N.H. Burnett, P.B. Corkum: Opt. Lett. 16,
(1992) 835 (1991)
15. K. Kondo, T. Tamida, Y. Nabekawa, S. Watanabe: Phys. Rev. A 49, 44. K.C. Kulander, B.W. Shore: Phys. Rev. Lett. 62, 524 (1989)
3881 (1994) 45. J. Peatross, D.D. Meyerhofer: Phys. Rev. A 52, 3976 (1995)
16. M.D. Perry, G. Mourou: Science 264, 917 (1994) 46. P. Salieres, A. L’Huillier, M. Lewenstein: Phys. Rev. Lett. 74, 3776
17. S.G. Preston, A. Sanpera, M. Zepf, W.J. Blyth, C.G. Smith, J.S. Wark, (1995)
M.H. Key, K. Burnett, M. Nakai, D. Neely, A.A. Offenberger: Phys. 47. C. Altucci, N.A. Ansari, R. Bruzzese, C.d. Lisio, S. Solimeno, R.B. Kay:
Rev. A 53, R31 (1996) J. Phys. B: At. Mol. Opt. Phys. 26, 1761 (1993)
18. J.L. Krause, K.J. Schafer, K.C. Kulander: Phys. Rev. Lett. 68, 3535 48. P. Salières, T. Ditmire, M.D. Perry, A. L’Huillier, M. Lewenstein: J.
(1992) Phys. B: At. Mol. Opt. Phys. 29, 4771 (1996)
19. P.B. Corkum: Phys. Rev. Lett. 71, 1994 (1993) 49. T. Ditmire, J.K. Crane, H. Nguyen, M.D. Perry, J. of Nonlinear Opt.
20. K.C. Kulander, K.J. Schafer: Limits on High-Order Harmonic Gen- Phys. and Mat. 4, 737 (1995)
eration from Single-Atom Calculations, In Multiphoton Processes, 50. M. Born, E. Wolf: Principles of Optics: Electromagnetic Theory of
ed. D.K. Evans, S.L. Chin (World Scientific, Singapore 1994), Propagation Interference and Diffraction of Light (Pergamon, Oxford
pp. 391–396 1980)
21. C.-G. Wahlström, J. Larsson, A. Persson, T. Starczewski, S. Svanberg, 51. J.W. Goodman: Statistical Optics (John Wiley, New York 1985)
P. Salieres, P. Balcou, A. L’Huillier: Phys. Rev. A 48, 4709 (1993) 52. M.D. Perry, C. Darrow, C. Coverdale, J.K. Crane: Opt. Lett. 17, 523
22. J. Zhou, J. Peatross, M.M. Murnane, H.C. Kapteyn, I.P. Christov: Phys. (1992)
Rev. Lett. 76, 752 (1996) 53. S. Augst, D.D. Meyerhofer, D. Strickland, S.L. Chin: J. Opt. Soc. B 8,
23. T. Ditmire, J.K. Crane, H. Nguyen, L.B. DaSilva, M.D. Perry: Phys. 858 (1991)
Rev. A 51, R902 (1995) 54. T. Ditmire, K. Kulander, J.K. Crane, H. Nguyen, M.D. Perry: J. Opt.
24. T. Ditmire, J.K. Crane, K.S. Budil, M.D. Perry, P. Salieres, A. L’Huil- Soc. Am. B 13, 406 (1996)
lier: Coherence Properties of High Order Harmonic Radiation, In 55. C. Altucci, C. Beneduce, R. Bruzzese, C. Delisio, G.S. Sorrentino,
Conference on Ultrafast Phenomena IX, Vol. 60, ed. P.F. Barbara et al. T. Starczewski, F. Vigilante: J. Phys. D: App. Phys. 29, 68 (1996)
(Springer, Berlin 1994) p. 228–232
25. M.E. Faldon, M.H.R. Hutchinson, J.P. Marangos, J.E. Muffett, R.A.
Smith, J.W.G. Tisch, C.G. Wahlström: J. Opt. Soc. Am. B 9, 2094
(1992)

S-ar putea să vă placă și