Expand this Topic clickable element to expand a topic
Skip to content
Optica Publishing Group

Simple models for laser-induced damage and conditioning of potassium dihydrogen phosphate crystals by nanosecond pulses

Open Access Open Access

Abstract

When potassium dihydrogen phosphate crystals (KH2PO4 or KDP) are illuminated by multi-gigawatt nanosecond pulses, damages may appear in the crystal bulk. One can increase damage resistance through a conditioning that consists in carrying out a laser pre-exposure of the crystal. The present paper addresses the modeling of laser-induced damage and conditioning of KDP crystals. The method is based on heating a distribution of defects, the cooperation of which may lead to a dramatic temperature rise. In a previous investigation [Opt. Express 15, 4557-4576 (2007)], calculations were performed for cases where the heat diffusion was permitted in one and three spatial dimensions, corresponding respectively to planar and point defects. For the sake of completeness, the present study involves the 2D heat diffusion that is associated with linear defects. A comparison to experimental data leads to the conclusion that 1D calculations are the most appropriate for describing the laser-induced damage in KDP. Within this framework, the evolution of the damage density is given as a function of the laser energy density and an in-depth analysis of the results is provided based on simple analytical expressions that can be used for experimental design. Regarding the conditioning, assuming that it is due to a decrease in the defect absorption efficiency, two scenarios associated with various defect natures are proposed and these account for certain of the observed experimental facts. For instance, in order to improve the crystal resistance to damage, one needs to use a conditioning pulse duration shorter than the testing pulse. Also, a conditioning scenario based on the migration of point (atomic-size) defects allows the reproduction of a logarithmic-like evolution of the conditioning gain with respect to the number of laser pre-exposures. Moreover, this study aims at refining the knowledge regarding the precursor defects responsible for the laser-induced damage in KDP crystals. Within the presented modeling, the best candidate permitting the reproduction of major experimental facts is comprised of a collection of one-hundred-nanometer structural defects associated with point defects as for instance cracks and couples of oxygen interstitials and vacancies.

©2009 Optical Society of America

1. Introduction

Dihydrogen phosphate crystals (KH2PO4 or KDP) and their deuterated analogs (KD2PO4 or DKDP) are widely used to perform frequency conversions of laser pulses. For instance, such crystals can be employed to produce intense nanosecond laser pulses at 3ω (a wavelength of 351 nm) in order to initiate fusion reaction [1]. However, under current operating conditions of the National Ignition Facility in the U.S.A. or the Laser MegaJoule in France, for which the laser energy density and the pulse duration are close to 10 J.cm -2 and 3 ns respectively, defects damaging the optical properties appear in the crystal bulk. It is now admitted that this damage is induced by precursor defects that efficiently absorb the laser energy, inducing a fast temperature rise and a subsequent shock wave [2]. In order to ensure a good working order of these large laser aperture facilities, it is crucial to manage the creation of damage. To do so, one needs to identify the nature of the precursor defects and to understand the physical mechanisms leading to damage. In order to approach more satisfactory operating conditions, i.e. to increase the LID threshold, one can employ a process consisting in pre-illuminating the crystal by a laser pulse for which the fluence is lower than that of the operating conditions [3, 4]. This process is commonly referred in the litterature to as conditioning. However, the physical mechanism induced by the conditioning remains unknown. Despite significant findings, a good comprehension of these mechanisms remains to be achieved. Furthermore, the nature of the precursor defects is still unknown. By providing simple models, this paper aims at improving the knowledge of damage and conditioning mechanisms in KDP crystals, and at providing some additional information regarding the nature of the precursor defects.

These precursor defects may be of varying nature. Since the crystal growth requires certain additives, the precursor defects may be constituted of atomic impurities, such as Fe, Cr or Si [5]. However, several experimental studies based on the correlation between Laser-Induced Damage (LID) and the concentration of impurities seem to show that this kind of defect is not involved in LID [6, 7, 8, 9]. Consequently, defects created during the crystal growth may be invoked. First, one can consider point defects, corresponding to hydrogen or oxygen atoms as interstitial atoms or vacancies in the crystalline lattice [10, 11, 12]. These can be associated with (PO3)- [13] or (HPO4)- groups [8, 9, 14]. Since simple considerations based on heat transfer [15, 5] have demonstrated that the precursor defect size ranges between roughly 10 nm and 100 nm in order to produce a sufficiently high temperature, only a cluster of point defects can satisfy the precursor defect size requirement. One can also envisage that, under laser exposure, the latter intrinsic defects give rise to others, for which the absorption properties are larger than the original defect. An example includes (PO3)- units that transform into (PO3)2- [13]. Another class of defects satisfying this size requirement corresponds to structural defects, such as mother liquid inclusions, dislocations or cracks. One can also regard LID to be due to a couple of defects for which only a cooperative mechanism permits an efficient laser energy absorption, inducing a high local temperature. It can for instance be imagined that an inclusion or a cluster of point defects produces cracks in its vicinity. Indeed, they strongly deform the surrounding lattice and, since KDP crystals exhibit a low mechanical resistance, this might lead to the creation of stress [16] and, eventually, small cracks.

In order to understand the LID origin at 3ω, a model based on the coupling of statistics and heat transfer has been developed [17]. In this modeling, the LID results from the aggregation of defects where the critical temperature is reached because of the cooperation between them. Despite the fact that this model is somewhat speculative, it renders it possible to determine several experimental trends, such as mainly the S-shape of the damage probability curves and a particular scaling law characteristic of KDP crystals linking the critical fluence, Fc, to the pulse duration, τ, as roughly Fcτ x with x≃0.35 [5, 4] while a standard value is close to 0.5 [18]. In [17], 1D and 3D heat diffusion, corresponding to planar and point defects respectively, were considered. It was shown that only planar defects were able to provide results that were in a good agreement with experimental data. It could thus be concluded from this modeling that growth bands, cracks, array of dislocations or staking faults were good candidates for explaining LID in KDP crystals. For the sake of completeness, Section 2 describes the study of 2D heat diffusion associated with linear defects. Further, in its original form, the model provided only damage probabilities. The interest in the damage density by the experimentalists has grown in the last years since it renders it possible to obtain more physical information, see for example [19, 6]. As a result, damage densities are now considered instead of probabilities, as presented in Section 3. Within this framework, an in-depth analysis of the model implications based on analytical derivations is presented.

Regarding the conditioning process, preliminary explanation attempts have been provided by Feit et al [15] who suggested that the increase in LID threshold is due to the precursor defects decreasing in size. According to Chirila et al, the conditioning may be due to a ”passivation” mechanism for which the electronic structure of the precursor defects is altered by electrons or holes produced by the laser pre-exposure. Despite that these attempts give a preliminary insight regarding the manner in which the conditioning may work, this is merely a phenomenological description of the conditioning process. Further, a detailed comparison to experimental results in order to verify and report on the reliability of the assumptions has not been carried out. Section 4 aims at introducing a modeling of the conditioning process that is based on the damage modeling assumptions. Since there exist various types of precursor defects [20] for which conditioning may differ, two physical mechanisms, corresponding to different types of precursor defects, are proposed. The first model assumes the precursor planar defects to actually be composed of point defects for which the thermally-activated migration may lead to their annihilation. Another class of defects are structural ones, and the second model relies on the fact that a rise in the temperature induces phase transition and a subsequent crystalline rearrangement. As a consequence, the defect absorption vanishes [21]. The predictions of each model are compared to a list of experimental trends, including:

• The shape of the damage density with respect to the fluence is practically unchanged by laser conditioning, the influence of the conditioning consists in shifting the whole curve to higher fluences [6].

• The conditioning efficiency can be increased by utilizing short pulses [4].

• As long as no damage appears during the conditioning, its efficiency becomes increasingly improved as the conditioning fluence is augmented [20, 6]. Further, there exists a conditioning fluence threshold below which no conditioning effect whatsoever is observed [20].

• For given conditioning pulse parameters, the conditioning efficiency increases as a function of the number of conditioning pulses [20, 22].

Finally, Section 5 provides conclusions and outlooks of the present work. For the convenience of the reader, details of certain derivations and a table describing the symbols used throughout the paper are reported in the Appendices.

2. 2D heat diffusion: study of the heating of an ensemble of linear defects

A heterogeneity can be considered as composed of an ensemble of linear defects (that can be associated with dislocations) oriented in the same direction. Since a dislocation disturbs the crystalline lattice, its absorption efficiency is larger than the one of a perfect crystal and it can be seen as a source inducing a temperature rise. The dislocation length is assumed to be larger than the thermal diffusion length given by 2Dτ where D is the thermal diffusivity. It follows that the temperature field remains the same whatever the position along the dislocation direction. In the following, the mathematical formalism used to model the problem is briefly described, but more details are given in [17]. The temperature field in the plane perpendicular to this direction, hereafter referred to as the 𝓟-plane, and for which the characteristic dimension has been set to 1 µm which is comparable to experimental dimension [23, 24], can be obtained by solving the Fourier equation:

Tt=DΔT+AρCΣi=0nADNSΠ(rri)

Here, r⃗i refers to the position (x,y) of the dislocation i randomly distributed in the 𝓟-plane, nADNS is the number of dislocations in the 𝓟-plane (ADNS stands for Absorbing Defects of Nanometric Size) and A is the absorbed power per unit of volume that can be expressed as 104 F/τ in units of W.cm -3 from empirical considerations [17]. Material physical parameters such as thermal diffusivity, D, and conductivity, λ, density, ρ, or heat capacity, C, are assumed to remain constant during the course of interaction. The function Π is defined as:

{Π(rri)=1aifx[xia2;xi+a2]andy[yia2;yi+a2]Π(rri)=0elsewhere

where a is the source size in the 𝓟-plane, set to 1 nm in the calculations. A general solution of Eq. (1) can be obtained [25] by summing the solutions of the Fourier equation designed for only one point source. Now, since working conditions include aDτ (for DKDP=6.5×10-7 m 2.s -1 and τ=1 ns, Dτ25nm), in order to deal with simple formula allowing fast numerical calculations, the function Π(x) can be approximated by the Dirac delta function δ (r⃗). The dislocations may then be seen as a point heat source in the 𝓟-plane. The temperature rise induced by this source reads (see Appendix B for the analytical derivation):

θ2D(r,t)=Aa24λKDPr24Dtduexp(u)u

Then, when considering a set of sources, cooperative effects lead to a higher temperature as opposed to the one induced by a single source. Further, the larger the number of ADNS involved in a temperature rise, the higher is the temperature. This is illustrated in Fig. 1 where a temperature field is associated with an ADNS distribution. Figures 1(a) and 1(b) respectively show the ADNS distribution and the resulting temperature field in a case without a big cluster. Figures 1(c) and 1(d) correspond to a case in which a cluster leading to a significant temperature rise is present (located in the circle of Fig. 1(c)). Moreover, it can be clearly seen that only clusters composed of a large number of ADNS give rise to high temperatures. It should be noted that the temperature scales differ between the two distributions. In one case, despite the appearance of clusters that significantly increase the temperature, it remained below Tc. In the second case, a cluster composed of more ADNS permitted the temperature to exceed Tc, subsequently giving rise to damage.

 figure: Fig. 1.

Fig. 1. A 2D temperature field in the 𝓟-plane (see text) for an ADNS distribution not leading to damage (top) and another temperature field implying a damage (bottom). The cluster leading to damage is inside the dashed red circle.

Download Full Size | PDF

The introduction of a damage probability by utilizing a critical temperature criterion is straightforward. For a given number of random ADNS distribution, it suffices to, at the end of the pulse, count the number of times for which the temperature is larger than the critical temperature in at least one place in the heterogeneity [17]. Figure 2 illustrates the results that can be obtained with this modeling. It shows the evolution of the damage probability as a function of the laser fluence with the following parameters: τ=3 ns, nADNS=2000 and 5000. The value of the parameter ξ/l has been set to 105 cm -1 in order to find a critical fluence (defined in the modeling as the fluence giving a probability of 10% [17]) close to the experimental one. The curve shape is comparable to those provided by the experiments. Within the modeling framework, this can be understood as the temperature rise increasing with the cluster size. Since the clusters are more numerous when their sizes are small, it follows that the higher the fluence, the larger the probability. Now, in order to determine whether the linear defects may be responsible for LID in KDP, one must verify if the calculations are able to reproduce experimental data. The main parameters characterizing the LID are Fc and x. With the large value ξ/l=105 cm -1 (compared to 104 cm -1, obtained from empirical considerations [17]) and nADNS=5000 (i.e. 0.5% of defects in a heterogeneity of 1 µm), one obtains Fc=8.11 J.cm -2 and x=0.21 for τ=3 ns (fit as shown by the inset (a) of Fig. 2). Fc is reasonable but not the value of x. With the same value of ξ/l but with nADNS=2000, a reasonable value x=0.31 is obtained, however, in this case, Fc=16.51 J.cm -2 is too high (inset (b) of Fig. 2). Also, ξ/l has to be set to 2×105 cm -1 in order to reproduce the experimental values of both Fc and x (Fc is here found to be 8.15 J.cm -2). Beyond the fact that ξ/l has a much larger value than the empirical one, it corresponds to an almost unrealistic physical process. Indeed, the temperature rise is due to the energy transfer from the free electrons (that have been produced themselves by the laser pulse [2]) to the lattice. The electronic density cannot exceed the critical plasma density given by nc= 0 ω 2/e 2≃9.1×1021 cm -3 at 3ω. By using a Drude model appropriately characterizing the electronic plasma, ξ/l can be linked to ne as A=ξF/lτ=nedEc/dt, where

dEcdt=e2ε0m×vcollc(ω2+vcoll2)×Fτ

is the energy absorbed by one electron per unit of time, and νcoll is the inverse of the time elapsed between two collisions of an electron on ions, commonly in the femtosecond range. In these calculations, ν -1 coll was set to 3 f s [26]. It follows that ξ/l=104 cm -1 corresponds to ne≃8×1021 cm -3, i.e. roughly the critical plasma density. With ξ/l=2×105 cm -1, one has ne≃20nc, which is physically unrealistic. These 2D heat diffusion calculations cannot be completely turned down despite the fact that slightly different parameters (a higher νcoll for example) would provide a more realistic value of ne. This order of magnitude calculation demonstrates that planar defects seem to provide data that are more consistent with experimental results as opposed to linear defects.

 figure: Fig. 2.

Fig. 2. The damage probability as a function of the laser fluence in the 2D modeling framework. The inset represents the critical laser fluence as a function of the pulse duration for (a) nADNS=5000 and (b) nADNS=2000.

Download Full Size | PDF

To conclude this section, an ensemble of linear defects, inducing a 2D heat transfer (that can be identified with a group of dislocations) can hardly fulfill all the conditions imposed by experimental data and physical requirements. Therefore, they appear not to be responsible for the LID in KDP crystals. Also, since 3D heat transfer calculations have already been turned down [17], focus is now placed on the 1D heat transfer associated with planar defects as the best means of explaining the LID in KDP crystals.

3. In-depth analysis of the damage density within the 1D modeling framework

It was shown that a collection of planar defects seemed to best explain the LID of KDP crystals within the present modeling framework. A study based on the damage probability, P, has been carried out [17]. Nevertheless, as of a few years, a great interest has been devoted to the experimental study of the damage density since it provides more information and avoids any dependence on the irradiated volume. Within the proposed model, the damage density, ρd, is simply obtained with ρd=Pρh, where ρh is the density of heterogeneities that are likely to cause damage. Fc is defined as the fluence required to reach a given damage density, ρc. It is set to 10 mm -3 after verification that a different value would lead to the same conclusions. Fig. 3 shows the evolution of the damage density as a function of the fluence for several values of parameters ρh, nADNS and τ with a heterogeneity of length L=10 µm [17]. The first conclusion that can be drawn is that the general shape exhibits characteristics similar to experimental results: a rapid increase at the threshold and a saturation for the highest fluences. Subsequently, the influences of the various parameters on the LID threshold were investigated. The influence of ρh is shown in Fig. 3(a). It appears that the larger ρh, the lower Fc. Indeed, when ρh increases, less energy is required to obtain a given number of heterogeneities for which T>Tc. The same reasoning leads to the conclusion that the larger nADNS, the lower Fc, as confirmed by Fig. 3(b). Concerning the influence of the pulse duration as displayed in Fig. 3(c), one finds the same behavior as the damage probability with Fcτ 0.35. Actually, the scaling law exponent deviates from the expected 1/2 value (associated with only one ADNS) as a result of cooperative effects between ADNS. An analytical development based on certain assumptions provides a better insight of this fact and allows to determine an approximated expression of x as (see Appendix C):

x12y(nADNS)zlnτ

where y(nADNS) is an increasing function of nADNS and z is a constant. Moreover, expression (5) shows that x becomes smaller as nADNS increases. Indeed, a large value of nADNS favors the cooperative effects and subsequently increases the deviation from 1/2. Also, large pulse durations provide the ADNS with more time to cooperate, thus increasing the deviation from 1/2 as shown by Eq. (5). Furthermore, it is noteworthy that the data of [27] could be better fitted with a value of x depending on the pulse duration as x=α+βlnτ (α and β are fit parameters) as opposed to with a constant value of x; a fact supporting the proposed modeling.

The behavior of Fc with respect to ρh and nADNS can be derived from analytical considerations based on probability calculations. In Appendix C, the damage density is approximated by:

ρd=ρh(NnADNS1)(nADNSN)f(τ)TcF

where f (τ) is a function of the pulse duration and N is related to the domain length as L=Na (with a=1 nm). By solving ρd=ρc where ρc=ρd(Fc), an approximated expression of the LIDT can be derived:

Fc=f(τ)Tcln(nADNSN)ln(ρcNρh)

Also, Fc is significantly reduced when nADNS or ρh increases. Furthermore, as part of the search for the nature of the precursor defects in KDP crystals, the LIDT is predicted to evolve as roughly the inverse of the logarithm of their concentration. In addition, experiments designed for measuring the LIDT variations with respect to a controlled concentration of defects should provide an answer to the question of whether the defects under investigation are responsible or not for the LID.

 figure: Fig. 3.

Fig. 3. The damage density as a function of the laser fluence in the 1D modeling framework. (a) The influence of the density of heterogeneities Nh with nADNS=100 and τ=1 ns (b) The influence of nADNS with Nh=106 cm -3 and τ=1 ns (c) The influence of the pulse duration with nADNS=100 and Nh=106 cm -3.

Download Full Size | PDF

Other means of extracting information from the experimental variations of the damage density involve considering a simple form of the theoretical expression of the damage density which only accounts for macroscopic measurable parameters and employing a fitting procedure. Since both the experimental and theoretical damage densities may suffer from artifacts at high fluences (coalescence [28] and saturation respectively), it would thus seem preferable to focus the attention on the threshold. For fluences close to Fc, the damage density can be approximated by (see derivation in Appendix C):

ρd(F)ρc(Nρhρc)(FFc)Fc

In Eq. (8), apart from Nρh, all terms can be determined from experiments. Also, the value of Nρh can be obtained from experimental data by using an adapted fitting procedure. It was verified that Eq. (8) provides a good reproduction of pure numerical calculations at the threshold.

4. Modeling of the laser conditioning

For a long time, it was experimentally demonstrated that one can increase the LIDT of KDP crystals by carrying out a laser pre-exposure. However, no clear proposal of how conditioning works has been put forward. Beyond the fact that it is of interest to understand the fundamental mechanisms underlying conditioning, from an applicative point of view, it is also important to control the influence of the laser parameters in order to optimize the conditioning protocol. Thus, in this section, two scenarios of conditioning are proposed.

4.1. Principle of the conditioning modeling

Despite that the nature of defects remains unknown, it has been determined that the LID is due to a considerable temperature rise. It is thus reasonable to expect that conditioning effects are also due to temperature rises. An increase in temperature influences a material such that it may activate the migration of defects or induce phase transition. Further, it has previously been demonstrated that planar defects seem to best explain LID in KDP crystals. For this reason, the following two scenarios of conditioning are proposed:

i. The above-mentioned planar defects are assumed to actually be composed of atomic-size defects for which displacements in the plane are governed by an Arrhenius law. It is supposed that these point defects can annihilate during their migration. Also, within this scenario, the conditioning consists in decreasing the absorption of each planar defect by decreasing the number of absorbing point defects. This scenario is hereafter referred to as Conditioning Model 1 (CM1).

ii. Here, planar defects are considered as a structural defect, and it is assumed that the conditioning laser pulse induces a phase transition (boiling) in the vicinity of the defect. Subsequently, when cooling to the room-temperature, rearrangements of the crystalline lattice remove the defect and its absorption efficiency vanishes. We here deal with a “zero-one” model. This scenario is referred to as Conditioning Model 2 (CM2).

The following section presents the details of the modeling. Within the CM1 framework, each planar defects is composed of an ensemble of point (atomic size) defects that may strongly absorb the laser energy. This results in a significant temperature rise due to cooperative heating effects. The temperature along the direction perpendicular to the plane decreases faster than exponentially [17]. Since a point defect is only able to move in a region where the temperature rise is significant (see Eq. (9) below), it can be reasonably stated that the migration takes place essentially in the plane. In the course of migration, these defects are assumed to be able to annihilate [9, 14, 13], i.e. a recombination of a pair of particles mediated by thermally activated diffusion may occur. For instance, interstitial-vacancy pairs satisfy this requirement but, atomic defects can in general migrate to a more fundamental state with a lower absorption efficiency. For the sake of simplicity, we henceforth deal with couples of interstitial and vacancy atoms. In order to numerically evaluate the absorption variation due to a laser conditioning, the following procedure was adopted. The absorption was assumed to be proportional to the number of defects pairs per unit of surface. For each time step, the probability Pm of a defect displacement from a cell (with a characteristic size of 1 nm) to an adjacent one is given by:

Pm(t)=exp(EakT(t))

where Ea is the activation energy and k is Boltzmann’s constant. A jump is assumed (depending on the value of Pm) to occur every 1 ps [9], with the same probability in each direction of the plane. After a time step, it is verified whether the interstitial atom has met or not a vacancy. If so, the number of pairs decreases by one. If no meeting occurs, the migration can go on. This cycle is performed as long as the temperature is such that Pm(t) is larger than one percent of Pm(t=τ), which is the maximum value of the probability (for nanosecond conditioning pulses, a significative migration takes place up to 10 ns after the pulse has been switched off). After the migration has stopped, the remaining (non-annihilated) defect pairs are counted and the new absorption of a plane is given by αcond=nrα 0/n 0, where nr is the density of the remaining pairs, n 0 is the density of the initial pairs, and α 0 is the plane absorption before the conditioning. Various details concerning the numerical implementation of the algorithm permitting to optimize the calculations are given in Appendix D.

For CM1, in order to speed up the calculations, other reasonable assumptions have been made. Also, periodic conditions have been used to carry out the migration process. The planar defect absorption efficiency was assumed constant in the course of interaction whereas a few recombinations may occur. Nevertheless, the migration becomes really effective at the end of the laser pulse and after a certain time when the temperature is the largest and the probability, accordingly, is the highest. Furthermore, a homogeneous heating of the plane was assumed, i.e. T(x,y, t)=T(t) in the calculations.

Concerning CM2, since it was shown that the defects might have a planar geometry and, because one often deals with rapidly grown crystals, structural defects represent good candidates for explaining LID. When illuminated by a laser pre-exposure, the absorption properties of a structural defect may lead to a significant local temperature rise and subsequently to a phase transition in its surroundings. The phase transition of interest is the boiling, as it is the most suited for removing structural defects [29]. Since this phase transition takes place in a confined medium where a pressure of tens of GPa can be reached [2], a higher temperature than ambient is required for the boiling point, Tbp [30, 31]. Under ambient conditions, the boiling temperature is close to 673 K, however, in the calculations, Tbp is set to roughly 2000 K in order to take into account the latter fact. Now, depending on the structural defect thickness, e, a minimum quantity of matter has to be heated up to Tbp in order to remove this defect. Since a 1D heat diffusion is considered, this quantity of matter is proportional to the distance, d, from the structural defect for which TTbp; at least for an instant (since one has a propagating heat front). Therefore, in the modeling, the conditioning succeeds if dβe, where β is a constant set to 5 and e is assumed to range randomly between 0.1 nm and 2 nm [32]. For a given structural defect, provided that the latter inequality is satisfied, the absorption then is assumed to be zero. It is worth noting that any reasonable variation of these parameters does not lead to a change in the main results presented hereafter.

At this point, on the basis of both conditioning models, it is possible to predict the influence of a pre-exposure on the damage density produced by a testing pulse. The numerical simulation mimics the experimental protocol: for a given ADNS distribution with the absorption efficiency as defined previously, a conditioning is simulated by first calculating the laser-induced temperature field and, then, evaluating the new absorption efficiency of each ADNS. It is noteworthy that each ADNS is subjected to a specific temperature (depending on whether cooperative effects take place or not), and as a result, each ADNS has its own new absorption efficiency. With these absorption efficiencies, it is possible to carry out the standard damage testing in order to obtain the damage density. Since the influence of the conditioning is to decrease the absorption efficiency, it is clear that the LIDT will increase with such a treatment. Once the values of the LIDT before and after conditioning have been evaluated, hereafter referred to as F (bc) c and F (ac) c, respectively, one can define the conditioning gain as:

g=Fc(ac)Fc(bc)

for which the values are larger than unity. Here a ratio is deliberately considered because it is less sensitive to the value of the modeling parameters (compared to an absolute value of the fluence or the damage density after conditioning). Moreover, g depends on the testing pulse duration, τ, the conditioning pulse duration, τcond, as well as the conditioning laser fluence.

4.2. Damage densities

By applying the above-mentioned procedure, one can evaluate the damage density after conditioning. An illustration of the results that can be obtained for certain values of the CM parameters is given by Fig. 4. Calculations have been performed with constant values of the laser pulse parameters, i.e. τtest=3 ns, Fcond=3 J.cm -2 and τcond=2 ns. A value of τcond shorter than τtest was deliberately chosen due to the conditioning being all the more efficient when τcond is short. This is further demonstrated in the following section.

Figure 4(a) shows the predictions of CM1 with Ea=0.9 eV and 1.2 eV, and n 0=6.×1012 cm -2 and 1.2×1013 cm -2 (corresponding to a few hundreds of point defects in 100 nm×100 nm in size plane). As expected, regardless of the CM1 parameter values, the LIDT could be increased by performing a conditioning procedure. Further, the larger the Ea, the lower the LIDT since the probability (9) is a decreasing function of Ea. Indeed, as this probability is lowered, less annihilations occur and the absorption thus decreases to a lesser extent. Concerning the influence of n 0, it appears that the LIDT increases along with n 0. For the largest value of n2D, the probability of a moving interstitial to meet a vacancy is at a maximum. This results in a large number of annihilations and subsequently to a low absorption efficiency. The results of Fig. 4(a) exhibit a particular trend with regard to the shape of the damage densities with respect to the fluence: after conditioning, the damage density slope displays a significant increase. This is due to the fact that the biggest clusters are the most sensitive to the conditioning (they exhibit the highest temperatures due to the cooperative effects [17]). Experimentally, this fact has not been observed at 3ω.

 figure: Fig. 4.

Fig. 4. The damage density as a function of the laser fluence for (a) an unconditioned and a conditioned crystal within the CM1 modeling framework and (b) a conditioned crystal within the CM2 modeling framework.

Download Full Size | PDF

Figure 4(b) presents the results of CM2 obtained with c=5 and 15, and Tbp=1600 K and 2000 K. As CM1, CM2 predicts an increase of the LIDT. For c=5, the damage density slope is significantly increased by applying CM2 (compared to an unconditioned crystal). This, again, is due to the conditioning being more efficient for larger clusters. In the boiling temperature range of interest (i.e. Tbp around 2000 K), for c=5, the influence of Tbp principally consists in shifting the damage density curve. For a given Tbp, in addition to the displacement of the curve to lower fluences, the increase of c leads to a decrease of the slope. In that case, for the biggest clusters, the matter has to be distributed among the ADNS (forming the cluster), which results in an effective distance that may be shorter than the smaller clusters. In fact, numerical calculations (that provide the average number of ADNS related to every value of d) shows that all cluster sizes are conditioned for c=5, whereas the biggest clusters are less conditioned for c=15. There are thus two influences that offset each other, which results in a slope that does not evolve much between an unconditioned and a conditioned crystal. Also, with CM2, one can find a set of parameters that almost allows the conservation of the damage density slope after conditioning.

4.3. Influence of the laser pulse parameters

From a general point of view, the conditioning gain depends on the pulse duration of both the testing and the conditioning laser pulses, as well as the fluence of the conditioning pulse. Let us first focus on the influence of the pulses duration. The conditioning gain with respect to τtest and τcond provided by CM1 and CM2 is given in Figs. 5(a) and 5(b) respectively. For both models, Fcond=5 J.cm -2. The CM1 parameters are given by Ea=1.2 eV and n 0=1.2×1013 cm -2, and those of CM2 are c=6 and Tbp=2000 K. These parameters have been chosen in order to obtain comparable gains for both models. From both graphs, due to the iso-gain curves being mainly horizontal, it can be deduced that the gain depends more on τcond than on τtest. In order to better appreciate the evolution of the gain as a function of τcond, a cut of Figs. 5(a) and 5(b) has been plotted in Fig. 5(c) for τtest=3 ns. This is the pulse duration of interest for the laser facilities presented in the introduction. For both CM’s, it appears clear that the conditioning efficiency is better for shorter values of τcond. Moreover, Fig. 5(c) shows that CM2 gives rise to the largest gain variations with respect to τcond. For both CM’s, it is worth noting that for the shortest conditioning pulses (sub-nanosecond duration), the temperature of the biggest clusters may exceed Tc. It follows that the conditioning may produce damage.

Next, it is investigated how the gain evolves with respect to the conditioning pulse duration and fluence. Again for τtest=3 ns, Figs. 6(a) and 6(b) portray this gain evolution predicted by the CM1 and the CM2, respectively. In order to obtain a better insight of the gain evolution as a function of Fcond, a cut of Fig. 6(a) and Fig. 6(b) has been reported to Fig. 6(c), with τcond=500 ps and τcond=3 ns. Both models exhibit the same trend: the conditioning gain depends more on Fcond than on τcond. Indeed, the conditioning efficiency is directly related to the temperature reached by each cluster during the conditioning. This temperature is proportional to Fcond whereas it evolves slowly with respect to τcond (as τ-0.3 [17, 26]). The gain shape with respect to Fcond is given in Fig. 6(c). The general behavior of all curves implies that the gain remains close to unity up to a certain fluence, after which there is a linear increase. The curves thus exhibit a conditioning fluence threshold above of which the conditioning becomes efficient. The existence of such a threshold is linked to the fact that a minimum temperature has to be reached in order to modify the absorption properties of the ADNS. Within CM1, the probability Pm of defect displacement, and the subsequent possibility of annihilation, become non-negligible for kTEa. Within CM2, the absorption efficiency of an ADNS vanishes only if TTbp in the area surrounding the structural defect. Concerning the behavior of g for fluences larger than the above-mentioned threshold, it can be easily understood by the statement that the temperature is proportional to the fluence. Further, from Fig. 6(c), one notices that the longer τcond, the higher is the conditioning threshold. In addition, one finds again that CM2 is more sensitive to τcond than CM1.

 figure: Fig. 5.

Fig. 5. The LIDT gain obtained by conditioning with respect to τtest and τcond within (a) CM1 and (b) CM2. For both graphs Fcond=5 J.cm -2. (c) A cut of (a) and (b) for τtest=3 ns.

Download Full Size | PDF

4.4. Influence of the number of pre-exposures

In the previous section, it was seen that utilizing a conditioning procedure comprising several laser pulse pre-exposures is the best way to optimize the crystal performances. It is thus interesting to consider how the conditioning gain evolves as a function of the number, Np, of laser pre-exposures.

Within the CM1, for a given laser pre-exposure, the initial positions of defect couples correspond to the final positions of the previous laser pulse pre-exposure. The details of an optimized implementation are given in Appendix D. After each shot, annihilations occur. However they are fewer and fewer since the absorption and thus also the temperature are increasingly low. Therefore, the gain is expected to increase as a function of the number of pre-exposures, however with smaller and smaller variations. This is confirmed by Fig. 7, portraying this behavior as numerically provided by CM1 with various pulse parameters, i.e. Fcond=2 J.cm -2 and 3 J.cm -2, and τcond=500 ps and 3 ns. Between two sucessive laser exposures, the fluence and duration remain unchanged and it is assumed that the temperature has enough time to decrease back to room temperature. This behavior is further confirmed by a numerical benchmark and an analytical derivation specifying the gain evolution as close to a logarithmic increase (see Appendix D). The inset of Fig. 7 represents the gain with a logarithmic scale on the horizontal axis and allows to convince of this fact. More precisely, regarding the numerical benchmark, the migration of the couples of defects has been fully simulated (without optimization) in a single plane. This calculation also provides the same logarithmic behavior as the one obtained with CM1. Further, the point defects migration and annihilation in a 3D space has been simulated and again, the calculations show a logarithmic-like evolution of the absorption (assumed to be proportional to the density of non-annihilated point defects) as a function of the number of pre-exposures, closely related to the ones obtained previously. It follows that, in this case, the space dimensionality does not permit a discrimination of the defect geometry (plane or sphere).

 figure: Fig. 6.

Fig. 6. The LIDT gain obtained by conditioning with respect to τcond and Fcond within (a) CM1 and (b) CM2. For both graphs τtest=3 ns. (c) Cut of (a) and (b) for τcond=500 ps and 3 ns.

Download Full Size | PDF

Concerning CM2, since an ADNS has been conditioned (with the first pre-exposure), it no longer contributes to the temperature rise for the following pre-exposure. Also, the temperature rise induced by the second pre-exposure is lower than the rise of the first whatever the considered ADNS. As a result, the ADNS that have not been conditioned by the first pre-exposure cannot be conditioned by the second. It also turns out that CM2 does not provide any variation of the gain with respect to the number of pre-exposures for Np>1.

4.5. Discussion

In order to put forward a conditioning model, this section first proposes a comparison of the results of CM1 and CM2 to available experimental data. The influence of the conditioning on the shape of the damage density with respect to the fluence, the influence of the conditioning pulse duration, the influence of the conditioning fluence, and the evolution of the gain as a function of the number of laser pre-exposures are also examined.

From previous works [6, 4], it seems that the shape of the damage density with respect to the fluence remains unchanged after a laser conditioning at 3ω. More precisely, the influence of the conditioning consists mainly in shifting the whole curve to larger fluence values. Section 4.2 showed that CM1 always leads to a modification of the damage density shape whereas, within CM2, it is possible to find parameters that allow the conservation of this shape, the conditioning simulation mainly leading to a shift of the curve under these conditions. Therefore, concerning the influence of the conditioning on the damage density shape, CM2 displays a better capacity for mimicking the experimental results.

 figure: Fig. 7.

Fig. 7. The conditioning gain as a function of the number of laser pre-exposures. The inset shows the case Fcond=2 J.cm -2 and τcond=3 ns with a logarithmic scale on the horizontal axis.

Download Full Size | PDF

Regarding the influence of the conditioning pulse duration on the gain, both models predict an increase in the conditioning efficiency as the pulse length decreases under conditions where Fcond is a constant not depending on τcond. This trends is in agreement with the experimental data [4]. Nevertheless, it is worth noting that a closer comparison to the experimental protocol should have used a ramp with increasing fluences; the highest fluence being an increasing function of τcond [4]. One should however keep in mind that the study was performed by varying physical parameters one by one in order to reach a good insight of the physical mechanisms. Further, the choice not to strictly follow the experimental protocol was based on the fact that it was not possible to accurately reproduce all the experimental conditions, such as for instance an installation change required to obtain large variations of τcond [33]. In the performed calculations, the introduction of Fcond increasing with respect to τcond would have been to decrease the slope of g. Going further in the comparison to experiments, the agreement is decent as long as τcond is not too short (no less than roughly 800 ps according to [4]). For shorter conditioning pulses, the trend is reversed: the shorter the pulse, the lower the gain. In such a case of very short pulses, one may expect that new physical effects are involved. For instance, this trend could be reproduced by CM1 by invoking the saturation of the migration probability. Indeed, this probability becomes close to unity and remains unchanged as the temperature goes up. Also, for increasingly short pulses, inducing higher and higher temperatures, the time allowed for migration vanishes whereas the probability ceases to increase. This results in shorter defect displacements and subsequently less annihilations, consequently leading to a lower gain. The only difference between CM1 and CM2 lies in the variations of the gain as a function of the conditioning pulse duration: CM2 exhibits larger variations than CM1. However, due to all the uncertainties (theoretical and experimental alike), the discrepancy is not enough to discriminate one model as opposed to the other by comparing to the experimental data.

Both conditioning models predict that a minimum conditioning fluence, F(m)cond, is required to initiate an improvement of the crystal resistance. For Fcond<F(m)cond, Fc remains unchanged (i.e. F (ac)c=F(bc)c). Again, this fact has been experimentally observed [20], showing that a certain temperature is required to initiate a conditioning process. This also validates the proposed models based on temperature-driven mechanisms. Further, above F(m)cond, for an increasing conditioning fluence, the gain is also augmented, which is in good agreement with the experimental observations. It follows that the higher the fluence, the better the gain. But it has been numerically observed that the critical temperature can be reached for a too high fluence, thus leading to damage. In that case, the beneficial effect of the conditioning is lost. This inconvenience can be eliminated by using several conditioning pulses of increasing fluence. Under such conditions, the first conditioning pulse with the lowest fluence passives only the biggest clusters. Subsequently, for an increasing fluence, smaller and smaller clusters are treated. This procedure also allows the use of a final high fluence, Fmaxcond, which gives rise to a decent gain, but without creating any damage. This protocol is also used experimentally and is called a conditioning ramp [4]. Moreover, the higher the Fmaxcond, the larger g, as also observed elsewhere [6]. Nevertheless, there exists an experimental limit for Fmaxcond [6] that may be due to the fact that another kind of defect begins to be excited.

The evolution of the gain as a function of the number of laser pre-exposures is the best way to test the CM’s since they predict very different trends. Within CM2, the gain does not evolve whereas CM1 predicts an increase. The latter phenomenology was observed in [20]: the damage density decreases as a function of the number of pre-exposures. Since the damage density after the conditioning and the conditioning gain are closely related, it can be concluded that CM1 accounts well for the influence of the number of pre-exposures. The comparison can be taken further by analyzing the variations of the conditioning efficiency with respect to the number of pre-exposures. According to [20], due to the damage density evolving almost linearly with a horizontal logarithmic scale, it was concluded that it exhibits a behavior close to a logarithmic evolution with respect to the number of laser pre-exposures. For fused silica, Suprasil and BK7, it is worth noting that a logarithmic behavior of F(ac) c, i.e. of the conditioning gain, has been also experimentally observed [34, 35, 36]. As shown in Section 4.4, CM1 appeared to also be able to reproduce this particular logarithmic-like behavior. Further, within this horizontal logarithmic scale, the slope of g depends on the conditioning fluence (see Appendix D), as it has been observed [22] for λ=532 nm and λ=355 nm. Furthermore, the higher Fcond, the larger the variations in g with respect to N (i.e. the larger the variations of the damage density). It is important to stress that this logarithmic-like behavior arises from the fact that an Arrhenius law was employed (see Appendix D). Another law would provide another type of behavior. Since this law is representative of the temperature-driven defect migration process, it follows that an agreement with the experimental data represents a strong evidence that the defects responsible for the LIDT in KDP are in fact clusters of absorbing point defects.

Nevertheless, a model based solely on this migration mechanism is not able to reproduce all the experimental trends, such as the evolution of the damage density shape after conditioning. It has been seen that the other model succeeds in this prediction. These results thus suggest that the precursor defects may have an extremely complex nature, combining the defect natures of both models. Also, only the association of a 2D structural defect with point defects could be able to reproduce all the experimental results that have been previously mentioned.

Now, one can wonder about the physical nature of the defects which can be associated with the above-mentioned point and structural defects. With regard to the point defects, couples of interstitials atoms and vacancies correspond well to the idea of them migrating and annihilating when an interstitial atom meets a vacancy. In that case, the absorption associated with these couples becomes zero after the annihilation. Concerning the nature of the considered atom, oxygen is believed to be a good candidate. Indeed, an oxygen vacancy introduces states located in the band gap [10, 11, 12] that enhance the absorption. In addition, these states allow an explanation for the particular experimental evolution of Fc as a function of the photon energy exhibiting successive plateaus [37]. The planar structural defects may be associated with cracks. Indeed, they exhibit the required absorption efficiency, since the breaking of periodic conditions first introduces states in the band gap in the vicinity of the crack surface [21], after which electric field enhancements can facilitate the free electron production [38, 39, 21] leading to the absorbing plasma. Furthermore, a crack may contain impurities that re-enhance the previous influences.

5. Conclusion

A model was developed based on statistics and heat transfer permitting to evaluate the temperature field induced by a distribution of defects in a heterogeneity. A damage is due to the presence of a cluster in which cooperative effects lead to a dramatic temperature rise. Within this modeling framework, the heat diffusion can be allowed in one, two or three spatial dimensions, and the corresponding defects are planes, lines or points respectively. The present work completed the study of the influence of the defect geometry by performing 2D calculations. The comparison of 1D, 2D and 3D calculations to experimental data such as the value of the scaling law exponent, led to the conclusion that the 1D framework was the best suited to explain laser-induced damage in KDP crystals. The defects responsible for laser-induced damage in KDP were thus believed to be a collection of planar defects. Within this framework, it was possible to predict the damage density as a function of the laser fluence. An in-depth analysis allowed the derivation of simple analytical expressions of the scaling law exponent, the critical fluence and the damage density with respect to several physical parameters. From this, additional information concerning the nature of the precursor defects, such as their density, could be obtained based on experimental data.

Two models of conditioning were developed based on temperature-driven mechanisms. Both models assume that the conditioning results from a decrease in the defects’ absorption. For this, the first model assumes that the absorption decrease is due to the annihilation of point defects that can migrate in the above-mentioned planar geometry. The second model assimilates the planar defects to structural ones, the heating of which might lead to crystalline rearrangements and to a subsequent significant decrease in absorption. Both models account for particular experimental trends. More precisely, these models reproduce the fact that the conditioning gain is better for shorter conditioning pulse durations. Moreover, there exists a minimum conditioning fluence required for initiating a conditioning. Furthermore, when a conditioning ramp is used, the higher the final fluence, the better the conditioning efficiency. Regarding the evolution of the gain with respect to the number of laser pre-exposures, the modeling based on the migration of point defects renders it possible to reproduce a logarithmic-like increase that is experimentally observed. Concerning the zero-one model based on the structural change, it fails completely in this prediction. Therefore, this particular result suggests that the planar defects are composed of at least point defects such as couples of interstitial atoms and vacancies.

In order to reproduce all the experimental facts, the use of both conditioning models is required. It follows that both the associated defect natures have to be taken into consideration. This suggests that the precursor defects actually have a complex nature in the sense that they associate planar structural defects and point defects. This study thus leads to the conclusion that a one-hundred-nanometer crack associated with couples of interstitial oxygen and vacancies may be the precursor defect responsible for laser-induced damage in KDP crystals.

Appendix

Tables Icon

A. List of parameters and variables

B. Analytical solution of the 2D diffusion equation with a point source

In the following, we solve the Fourier equation with a point source in 2D space:

Tt=DΔT+Eδ(r)F(t)ρC

where D, ρ and C are the thermal diffusivity, the density and the specific heat capacity respectively. E has the dimension of an energy per unit surface and F(t) is the temporal profile of the laser pulse. In the case where F(t)=δ (t), by using Fourier and Laplace transformations, it can be shown that [25]:

θ2D(r,s)=TT0=E2πλKDPK0(rsD)

where K 0 is the modified Bessel function (see e.g. [40]). Now, instead of a temporal Dirac source, we assume a time-dependent rectangular laser pulse shape such that:

{F(t)=γif0tτF(t)=0elsewhere

Using convolution properties of Laplace transformation, the new temperature rise reads:

θ2D(r,s)=E2πλKDPK0(rsD)1eτss

where (1–eτs)/s is simply the Laplace transformation of the rectangular laser pulse given by Eq. (13). The latter expression can be written in the following form:

θ2D(r,s)=ζ(r,s)eτsζ(r,s)

with

ζ(x,s)=E2πλKDPτK0(rsD)s

We are only interested in the temperature at t=τ, thus implying that the last term of Eq. (15) can be omitted (since it contributes only for t>τ). Finally, the temperature induced by a 2D point source reads:

θ2D(r,t)=Aa24λKDPr24Dtduexp(u)u

where E has been set to πa 2. The integral in Eq. (17) is evaluated numerically.

C. Analytical derivations in the 1D framework

C.1. Evaluation of the scaling law exponent

In order to analytically evaluate the scaling law exponent x linking the critical fluence Fc to the laser pulse duration τ as Fcτ x, one considers two critical fluences F c1 and F c2 associated with the pulse durations τ 1 and τ 2 respectively. Consequently, x reads:

x=ln(Fc1Fc2)ln(τ1τ2)

In order to simplify the problem, we assume that a compact cluster of planar defects has the same temperature evolution as a single plane, i.e. T=T 0 +αF/τ 1/2 [25] where α is a constant. We further assume that the pulse of duration τ 2 probes essentially this compact cluster (located at x=0) whereas the pulse of duration τ 1 also probes all the ADNS close to the cluster which cooperate for the temperature increase at x=0. We thus suppose that τ 2≤a few ns [41] and τ 1τ 2. Under such conditions, we have:

T2(x=0,t=τ2)=T0+αF2τ212

and

T1(x=0,t=τ1)=T0+αF1τ112+ncoop4Dτ12Dτ12Dτ1dxθ1D(x,τ1)

Since we integrate up to a distance 2Dτ1 from the cluster, we approximate θ 1D(x,τ 1) by αFτ 1/2 exp(−x 2/4) (the term that principally contributes to the integral). Further, , where n 0 is a constant proportional to nADNS. Simple algebra leads to:

T1(x=0,t=τ1)=T0+αF1τ112+βF1n0

where β is a constant. We then obtain:

x=ln(τ112τ212(α+βn0τ112))ln(τ1τ2)=12ln(α+βn0τ112)ln(τ1τ2)

For a given ratio r=τ 1/τ 2 and under the assumption of a sufficiently large value of τ 1 in order to obtain the inequality β n 0 τ 1/2 1α, we finally have:

x12ln(βn0)lnrlnτ12lnr

Despite that this expression has been established based on several assumptions, it allows us to shed light on the behavior of x with respect to nADNS and τ : the larger nADNS, the lower x and the longer the pulse duration, the lower x as we can simply obtain with a physical feeling.

C.2. Damage density and LIDT

Sections 2 and 3 demonstrated that the LID is due to clustering of planar defects distributed along a direction perpendicular to the planes. Also, the damage probability corresponds to the probability of the cluster having a spatial random distribution of the defects. This probability is thus given by the ratio between the number of configurations with a cluster and the total number of configurations (number of ways to distribute nADNS in N places). The latter number is simply given by CNnADNS where Cpn=n!/p!(n–p)! is the binomial coefficient. The number of configurations with a cluster of size nd (all ADNS are assumed to be adjacent) is given by (NnADNS)CNndnADNSnd, as long as nADNSN. Straightforward calculations then lead to:

P(NnADNS)[nADNS(nADNS1)(nADNSnd1)]N(N1)(Nnd1)

In the case where nADNSnd (which represents a correct assumption in our calculations [17]), we obtain:

P(NnADNS)(nADNSN)nd

Now, as long as the cluster size is small compared to the thermal diffusion length, the temperature is proportional to the number of ADNS composing the cluster multiplied by the fluence. We can therefore write nd=f (τ)Tc/F where f (τ) is a function of the pulselength. Finally, the damage density reads:

ρdρh(NnADNS)(nADNSN)f(τ)TcF

We define the LIDT Fc as ρd(Fc)=ρc where ρc can be set to any arbitrary value. By taking the logarithm of Eq. (26), simple algebra leads to:

Fc=f(τ)Tcln(nADNSN)ln(ρcNρh)

C.3. Damage density at the threshold

In order to approximate the damage density evolution with respect to the fluence at the threshold, we can develop ρd as a Taylor expansion:

ρd(F)]F=Fc=ρd(Fc)+ρdF]F=Fc(FFc)1!+
2ρdF2]F=Fc(FFc)22!++nρdFn]F=Fc(FFc)nn!+

Since nADNSN, we have ‖lnn(nADNS/N)‖≫‖lnn-1(nADNS/N)‖. By using Eq. (26), the derivative at the order n of ρd can thus be approximated by:

nρdFn(f(τ)TcF2ln(nADNSN))ρd

It follows that:

ρd(F)]F=FcρcΣn=0(f(τ)TcFc2ln(nADNSN))(FFc)nn!
ρcexp{β(τ)Tc(FFc)Fc2ln(nADNSN)}

Finally, by using the expression (27) of Fc, we obtain an approximated form for the evolution of ρd at the threshold:

ρd(F)ρc(Nρhρc)(FFc)Fc

Due to the exponential nature of ρd(F) and because of the particular form of its derivatives, expression (32) can be shown to be simply the tangent of lnρd(F) at F=Fc.

D. Conditioning modeling

D.1. Details regarding the numerical implementation of the conditioning

We define the normalized absorption efficiency, α, as:

α=1nani

where ni is the initial density of couples of defects and na stands for the number of annihilations per unit of surface of the plane. Also, if no annihilations occur, na=0 and α=1 remain unchanged. In a case where all the couples of defects annihilate, na=ni and subsequently α=0 subsequently. When setting αN to be the absorption efficiency after Np laser pre-exposures, it is straightforward that:

αNp=1na(1)+na(2)++na(Np)ni

Here, n(Np)a is the number of annihilations due to the Npth laser pre-exposure. If we define the absorption efficiency α(Np)=1-n(Np)a/ni, then, Eq. (34) can be rewritten as:

αNp=α(1)+(α(2)1)+(α(3)1)++(α(Np)1)

In practice, we obtain a new absorption efficiency with the relation of recurrence α Np+1=αNp+(α (Np)-1) where αNp is known from the previous calculation and α(Np) is determined as described below. α (Np) is a function of the number of displacements, dNp, of each interstitial atom induced by the N pth shot. We have tabulated the evolution of n(Np)a for any number of displacements. Then α(Np) is obtained from these tabulated values for a known value of dNp that is simply given by the number of time steps weighted by the displacement probability, i.e. ∑i P(t=ti). Here P(t=ti) is the probability given by Eq. (9) and ti is a multiple of 1 ps (see Section 4.1).

D.2. Analytical analysis of the evolution of the gain as a function of the number of pre-exposures

From the last paragraph, it is straightforward that:

αNp+1=αNpna(Np+1)ni

Since n(Np)a is roughly proportional to the length covered by a migrating defect, provided that not too many annihilations have occurred, this length being itself proportional to the probability (9) of defect displacement, we deduce that n(Np) a is proportional to exp(-Ea/kT) where T is the temperature induced by the Npth laser pre-exposure. Within our modeling, this temperature is proportional to αNp. Thus Eq. (36) transforms into:

αNp+1=αNpβexp(γαNp)

where β is a constant and γEa/kFcond. The equivalent continuous form of Eq. (37) is:

dαdNp=βexp(γα)

Since Fcα -1 and using the definition of the gain (10), we deduce that g=α -1. It follows that the differential equation governing the variation of the gain with respect to Np reads:

dgdNp=βg2exp(γg)

To the best of our knowledge, this differential equation does not have a solution. Nevertheless, we can obtain some information concerning the behavior of g. First, the second term of Eq. (39) is positive, implying that g is an increasing function. Secondly, the asymptotic behavior of g when Np→∞ is imposed by the exponential function. Also, for the large values of Np, Eq. (39) can be approximated by g′=β exp(-γg), for which the solution is simply a logarithmic function. Also, we can expect that the solution of (39) is close to a logarithmic function. Indeed, a numerical resolution of Eq. (39) leads to the mentioned logarithmic behavior of g. Finally, since γ is inversely proportional to Fcond, a higher Fcond gives rise to larger variations of g with respect to Np.

Acknowledgments

Anthony Dyan is acknowledged for fruitful discussions, as are all members of the Groupe de Travail Endommagement Laser (GTEL), especially Pierre Grua, Jean-Pierre Morreeuw and Hervé Bercegol. Nicolas Mallejac has also contributed with helpful input. Many thanks are due to the MP2 group of the Institut Fresnel in Marseille. The AlphaScience company is acknowledged for its english corrections of the present manuscript.

References and links

1. J.J. De Yoreo, A.K. Burnham, and P.K. Whitman, “Developing KH2PO4 and KD2PO4 crystals for the world’s most powerful laser,” Int. Mater. Rev. 47, 113–152 (2002). [CrossRef]  

2. C.W. Carr, H.B. Radousky, A.M. Rubenchik, M.D. Feit, and S.G. Demos, “Localized dynamics during laser-induced damage in optical materials,” Phys. Rev. Lett. 92, 087401 (2004). [CrossRef]   [PubMed]  

3. J. Swain, S. Stokowski, D. Milam, and F. Rainer, “Improving the bulk laser damage resistance of potassium dihydrogen phosphate crystals by pulsed laser irradiation,” Appl. Phys. Lett. 40, 350–352 (1982). [CrossRef]  

4. J.J. Adams, J.R. Bruere, M. Bolourchi, C.W. Carr, M.D. Feit, R.P. Hackel, D.E. Hahn, J.A. Jarboe, L.A. Lane, R.L. Luthi, J.N. McElroy, A.M. Rubenchik, J.R. Stanley, W.D. Sell, J.L. Vickers, T.L. Weiland, and D.A. Willard, “Wavelength and pulselength dependence of laser conditioning and bulk damage in doubler-cut KH2PO4,” Proc. SPIE 5991, 59911R-1 (2005). [CrossRef]  

5. A.K. Burnham, M. Runkel, M.D. Feit, A.M. Rubenchik, R.L. Floyd, T.A. Land, W.J. Siekhaus, and R.A. Hawley-Fedder, “Laser-induced damage in deuterated potassium dihydrogen phosphate,” Appl. Opt. 42, 5483–5495 (2003). [CrossRef]   [PubMed]  

6. R. A. Negres, P. DeMange, and S. G. Demos, “Investigation of laser annealing parameters for optimal laser-damage performance in deuterated potassium dihydrogen phosphate,” Opt. Lett. 30, 2766–2768 (2005). [CrossRef]   [PubMed]  

7. M. Pommies, D. Damiani, B. Bertussi, J. Capoulade, H. Piombini, J.Y. Natoli, and H. Mathis, “Detection and characterization of absorption heterogeneities in KH2PO4 crystals,” Opt. Commun. 267, 154–161 (2006). [CrossRef]  

8. J.E. Davis, R.S. Hughes, and H.W. Lee., “Investigation of optically generated transient electronic defects and protonic transport in hydrogen-bonded molecular solids. Isomorphs of potassium dihydrogen phosphate,” Chem. Phys. Lett. 207, 540–545 (1993). [CrossRef]  

9. C.D. Marshall, S.A. Payne, M.A. Henesian, J.A. Speth, and H.T. Powell, “Ultraviolet-induced transient absorption in potassium dihydrogen phosphate and its influence on frequency conversion,” J. Opt. Soc. Am. B 11, 774–785 (1994). [CrossRef]  

10. C.S. Liu, N. Kioussis, S.G. Demos, and H.B. Radousky, “Electron or hole-assisted reactions of H defects in hydrogen-bonded KDP,” Phys. Rev. Lett. 91, 015505 (2005). [CrossRef]  

11. C.S. Liu, N. Kioussis, S.G. Demos, and H.B. Radousky, “Electronic structure calculations of an oxygen vacancy in KH2PO4,” Phys. Rev. B 72, 134110 (2005). [CrossRef]  

12. K. Wang, C. Fang, J. Zhang, C.S. Liu, R.I. Boughton, S. Wang, and X. Zhao, “First-principles study of interstitial oxygen in potassium dihydrogen phosphate crystals,” Phys. Rev. B 72, 184105 (2005). [CrossRef]  

13. N.Y. Garces, K.T. Stevens, L.E. Halliburton, S.G. Demos, H.B. Radousky, and N.P. Zaitseva, “Identification of electron and hole traps in KH2PO4 crystals,” J. Appl. Phys. 89, 47–52 (2001). [CrossRef]  

14. M.M. Chirila, N.Y. Garces, L.E. Halliburton, S.G. Demos, T.A. Land, and H.B. Radousky, “Production and thermal decay of radiation-induced point defects in KDPO crystals,” J. Appl. Phys 94, 6456–6462 (2003). [CrossRef]  

15. M.D. Feit and A.M. Rubenchik, “Implications of nanoabsorber initiators for damage probability curves, pulselength scaling and laser conditioning,” Proc. SPIE 5273, 74–82 (2004). [CrossRef]  

16. J.J. De Yoreo and B.W. Woods, “A study of residual stress and the stress-optic effect in mixed crystals of K(DxH1-x)2PO4,” J. Appl. Phys. 73, 7780–7789 (1993). [CrossRef]  

17. G. Duchateau and A. Dyan, “Coupling statistics and heat transfer to study laser-induced crystal damage by nanosecond pulses,” Opt. Express 15, 4557–4576 (2007). [CrossRef]   [PubMed]  

18. B.C. Stuart, M.D. Feit, A.M. Rubenchik, B.W. Shore, and M.D. Perry, “Laser-Induced Damage in Dielectrics with Nanosecond to Subpicosecond Pulses,” Phys. Rev. Lett. 74, 2248–2251 (1995). [CrossRef]   [PubMed]  

19. L. Lamaignère, S. Bouillet, R. Courchinoux, T. Donval, M. Josse, J.-C. Poncetta, and H. Bercegol, “An accurate, repeatable, and well characterized measurement of laser damage density of optical materials,” Rev. Sci. Instrum. 78, 103105 (2007). [CrossRef]   [PubMed]  

20. P. DeMange, R.A. Negres, C.W. Carr, H.B. Radousky, and S.G. Demos, “Laser-induced defect reactions governing damage initiation in DKDP crystals,” Opt. Express 14, 5313–5328 (2006). [CrossRef]   [PubMed]  

21. P. Grua, J-P. Morreeuw, and H. Bercegol, “Progress in the understanding of fracture related damage of fused silica,” Proc. SPIE 6720, 672003 (2007). [CrossRef]  

22. P. DeMange, C.W. Carr, R.A. Negres, H.B. Radousky, and S.G. Demos, “Laser annealing characteristics of multiple bulk defect populations within DKDP crystals,” J. Appl. Phys. 104, 103103 (2008). [CrossRef]  

23. S.G. Demos, M. Staggs, M. Yan, H.B. Radousky, and J.J. De Yoreo, “Investigation of optically active defect clusters in KH2PO4 under laser photoexcitation,” J. Appl. Phys. 85, 3988–3992 (1999). [CrossRef]  

24. S.G. Demos, M. Staggs, J.J. De Yoreo, and H.B. Radousky, “Imaging of laser-induced reactions of individual defect nanoclusters,” Opt. Lett. 26, 1975–1977 (2001). [CrossRef]  

25. H.S. Carslaw and J.C. Jaeger, Conduction of Heat in Solids (Oxford Science Publications, 1959).

26. A. Dyan, F. Enguehard, S. Lallich, H. Piombini, and G. Duchateau, “Scaling laws in laser-induced potassium dihydrogen phosphate crystal damage by nanosecond pulses at 3ω,” J. Opt. Soc. Am. B 25, 1087–1095 (2008). [CrossRef]  

27. J.J. Adams, T.L. Weiland, J.R. Stanley, W.D. Sell, R.L. Luthi, J.L. Vickers, C.W. Carr, M.D. Feit, A.M. Rubenchik, M.L. Spaeth, and R.P. Hackel, “Pulse length dependence of laser conditioning and bulk damage in KD2PO4,” Proc. SPIE 5647, 265 (2005). [CrossRef]  

28. M.J. Matthews and M.D. Feit, “Effect of random clustering on surface damage density estimates,” Proc. SPIE 6720, 67201J (2007). [CrossRef]  

29. By performing calculations considering the liquefaction as phase transition and only considering the volume expansion due to this phase transition, a minimum of 100 times the crack thickness is required to fill it, i.e. a length of a few hundreds of nanometers. Calculations show that with such a length, its value depends on the fluence but not on the pulse duration. since this fact differs from the experimental observations, the liquefaction should be turn down.

30. T.J. Norman, J.M. Zaug, and C.W. Carr, “High-Pressure Decomposition of DKDP,” Chem. Mater. 18, 3074–3077 (2006). [CrossRef]  

31. Y. Kobayashi, S. Endo, K. Koto, T. Kikegawa, and O. Shimomura, “Phase transitions and amorphization of KH2PO4 at high pressure,” Phys. Rev. B 51, 9302–9305 (1995). [CrossRef]  

32. If several ADNS contribute to the determination of d, e.g. n, then d transforms into d/n in our calculations.

33. The variation of the gain with respect to τcond can be increased by setting the modeling parameters to values differing from the ones used.

34. S. Xu, X. Yuan, W. Yin, X. Xiang, and X. Zu, “Effect of UV laser conditioning on fused silica in vacuum,” Opt. Mater. 31, 1013–1016 (2009). [CrossRef]  

35. J.Y. Natoli, B. Bertussi, and M. Commandré, “Effect of multiple laser irradiations on silica at 1064 and 355 nm,” Opt. Lett. 30, 1315–1317 (2005). [CrossRef]   [PubMed]  

36. L. Gallais, J.Y. Natoli, and C. Amra, “Statistical study of single and multiple pulse laser-induced damage in glasses,” Opt. Express 10, 1465–1474 (2002). [PubMed]  

37. C.W. Carr, H.B. Radousky, and S.G. Demos, “Wavelength dependence of laser-induced damage: determining the damage initiation mechanisms,” Phys. Rev. Lett. 91, 127402 (2003). [CrossRef]   [PubMed]  

38. N. Bloembergen, “Role of cracks, pores, and absorbing inclusions on laser induced damage threshold at surface of transparent dielectrics,” Appl. Opt. 12, 661–664(1973). [CrossRef]   [PubMed]  

39. F.Y. Génin, A. Salleo, T.V. Pistor, and L.L. Chase, “Role of light intensification by cracks in optical breakdown on surfaces,” J. Opt. Soc. Am. A 18, 2607–2616 (2001). [CrossRef]  

40. M. Abramowitz and I.A. Stegun, Handbook of Mathematical Functions (Dover, 1965).

41. The equivalent distance of about 100 nm corresponding to the mean distance between two ADNS with nadns=100 and n=10000, which represent standard values in our calculations.

Cited By

Optica participates in Crossref's Cited-By Linking service. Citing articles from Optica Publishing Group journals and other participating publishers are listed here.

Alert me when this article is cited.


Figures (7)

Fig. 1.
Fig. 1. A 2D temperature field in the 𝓟-plane (see text) for an ADNS distribution not leading to damage (top) and another temperature field implying a damage (bottom). The cluster leading to damage is inside the dashed red circle.
Fig. 2.
Fig. 2. The damage probability as a function of the laser fluence in the 2D modeling framework. The inset represents the critical laser fluence as a function of the pulse duration for (a) nADNS =5000 and (b) nADNS =2000.
Fig. 3.
Fig. 3. The damage density as a function of the laser fluence in the 1D modeling framework. (a) The influence of the density of heterogeneities Nh with nADNS =100 and τ=1 ns (b) The influence of nADNS with Nh=106 cm -3 and τ=1 ns (c) The influence of the pulse duration with nADNS =100 and Nh=106 cm -3.
Fig. 4.
Fig. 4. The damage density as a function of the laser fluence for (a) an unconditioned and a conditioned crystal within the CM1 modeling framework and (b) a conditioned crystal within the CM2 modeling framework.
Fig. 5.
Fig. 5. The LIDT gain obtained by conditioning with respect to τtest and τcond within (a) CM1 and (b) CM2. For both graphs Fcond =5 J.cm -2. (c) A cut of (a) and (b) for τtest =3 ns.
Fig. 6.
Fig. 6. The LIDT gain obtained by conditioning with respect to τcond and Fcond within (a) CM1 and (b) CM2. For both graphs τtest =3 ns. (c) Cut of (a) and (b) for τcond =500 ps and 3 ns.
Fig. 7.
Fig. 7. The conditioning gain as a function of the number of laser pre-exposures. The inset shows the case Fcond =2 J.cm -2 and τcond =3 ns with a logarithmic scale on the horizontal axis.

Tables (1)

Tables Icon

Table 1 A. List of parameters and variables

Equations (40)

Equations on this page are rendered with MathJax. Learn more.

Tt=DΔT+AρCΣi=0nADNS Π (rri)
{Π(rri)=1aifx[xia2;xi+a2]andy[yia2;yi+a2]Π(rri)=0elsewhere
θ2D(r,t)=Aa24λKDP r24Dt du exp (u)u
dEcdt=e2ε0m× vcollc(ω2+vcoll2)×Fτ
x 12 y (nADNS)zlnτ
ρd=ρh(NnADNS1)(nADNSN)f(τ)TcF
Fc=f (τ) Tc ln(nADNSN)ln(ρcNρh)
ρd (F)ρc(Nρhρc)(FFc)Fc
Pm(t)=exp (EakT(t))
g=Fc(ac)Fc(bc)
Tt=DΔT+Eδ(r)F(t)ρC
θ2D(r,s)=TT0=E2πλKDPK0(rsD)
{F(t)=γif0tτF(t)=0elsewhere
θ2D(r,s)=E2πλKDPK0(rsD)1eτss
θ2D(r,s)=ζ (r,s) eτsζ(r,s)
ζ(x,s)=E2πλKDPτK0(rsD)s
θ2D(r,t)=Aa24λKDPr24Dtduexp(u)u
x=ln(Fc1Fc2)ln(τ1τ2)
T2(x=0,t=τ2)=T0+αF2τ212
T1(x=0,t=τ1)=T0+αF1τ112+ncoop4Dτ12Dτ12Dτ1dxθ1D(x,τ1)
T1(x=0,t=τ1)=T0+αF1τ112+βF1n0
x=ln(τ112τ212(α+βn0τ112))ln(τ1τ2)=12ln(α+βn0τ112)ln(τ1τ2)
x 12ln(βn0)lnrlnτ12lnr
P (NnADNS)[nADNS(nADNS1)(nADNSnd1)]N(N1)(Nnd1)
P (NnADNS)(nADNSN)nd
ρd ρh (NnADNS) (nADNSN)f(τ)TcF
Fc=f(τ)Tcln(nADNSN)ln(ρcNρh)
ρd(F)]F=Fc=ρd(Fc)+ρdF]F=Fc(FFc)1!+
2ρdF2]F=Fc (FFc)22! ++nρdFn]F=Fc(FFc)nn! +
nρdFn (f(τ)TcF2ln(nADNSN)) ρd
ρd (F)]F=Fc ρc Σn=0 (f(τ)TcFc2ln(nADNSN)) (FFc)nn!
ρc exp {β(τ)Tc(FFc)Fc2ln(nADNSN)}
ρd (F)ρc(Nρhρc)(FFc)Fc
α=1nani
αNp=1na(1)+na(2)++na(Np)ni
αNp=α(1)+(α(2)1)+(α(3)1)++(α(Np)1)
αNp+1=αNpna(Np+1)ni
αNp+1=αNpβexp(γαNp)
dαdNp=βexp(γα)
dgdNp=βg2exp(γg)
Select as filters


    Select Topics Cancel
    © Copyright 2024 | Optica Publishing Group. All rights reserved, including rights for text and data mining and training of artificial technologies or similar technologies.