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

Mie scattering from submicron-sized CO2 clusters formed in a supersonic expansion of a gas mixture

Open Access Open Access

Abstract

A detailed mathematical model is presented for a submicron-sized cluster formation in a binary gas mixture flowing through a three-staged conical nozzle. By measuring the angular distribution of light scattered from the clusters, the size of CO2 clusters, produced in a supersonic expansion of the mixture gas of CO2(30%)/H2(70%) or CO2(10%)/He(90%), has been evaluated using the Mie scattering method. The mean sizes of CO2 clusters are estimated to be 0.28 ± 0.03 μm for CO2/H2 and 0.26 ± 0.04 μm for CO2/He, respectively. In addition, total gas density profiles in radial direction of the gas jet, measuring the phase shift of the light passing through the target by utilizing an interferometer, are found to be agreed with the numerical modeling within a factor of two. The dryness (= monomer/(monomer + cluster) ratio) in the targets is found to support the numerical modeling. The apparatus developed to evaluate the cluster-gas targets proved that our mathematical model of cluster formation is reliable enough for the binary gas mixture.

© 2013 Optical Society of America

1. Introduction

At the present time, cluster-gas targets are widely used as one of the promising media for the femtosecond laser-matter interaction [1], including nuclear fusion [2], x-ray emission [3, 4], betatron x-ray radiation [5], laser-driven ion acceleration [6,7], and so on. Because the cluster-gas targets are produced using typically a gas jet nozzle, the design of the gas jet nozzle is the key to such experiments success.

Therefore, we have designed the gas jet nozzle based on a mathematical model of the clusters formation in the gas flows in the nozzle [8]. It is considered to be able to produce the clusters of about 1 μm size for pure rare gases (Ar, Kr, and Xe) according to our computations. The cluster-gas targets have been demonstrated to be useful as a clean x-ray source for x-ray diffraction experiments [4]. However, the x-ray emissions are limited due to the absorption by a significant quantity of residual gas around the laser interaction region.

To avoid this problem, we have used the cluster-gas target produced by the gas jet of the mixture gas of 90% helium and 10% CO2 instead of the pure CO2 gas through the nozzle [9]. Since the helium gas helps clusterization of molecular gases, a small quantity of CO2 gas is enough to produce large CO2 clusters, while the helium gas does not clusterize and remains as a residual gas. In this case, the absorption of soft x-ray emission in the residual gas is significantly reduced due to the transparency of helium to the soft x-rays. This results that the number of photons for the CO2/He mixture gas is four to seven times larger than that produced with the pure CO2 gas.

Furthermore, we have observed a long plasma channel such as promising laser-driven particle accelerations using the cluster-gas target produced from the CO2/He mixture gas. In fact, efficient generation of high energy ions up to 50 MeV per nucleon was achieved [6, 7] beyond traditional methods. Thus the targets consisting of clusters embedded in a background gas produced by the appropriate mixture gases have a potential as a variety of radiation sources.

In this manner, some experiments using the binary gas mixture have been already performed, whereas the mathematical model of the clusters formation for the binary gas mixture has not been reported in the past. Therefore, we introduce the mathematical model of the clusters formation in the supersonic expansion of the binary gas mixture with one inert component in vacuum and evaluate experimentally the model in this study. According to the numerical modeling, the existence of submicron-sized clusters in the target is predicted. Because the conventional Rayleigh scattering method cannot apply to this size range, we have developed an apparatus capable of measuring micron-sized particles using Mie scattering of the second-harmonics of Nd:YAG laser pulses [10]. Because the detailed knowledge about the initial target parameters is the key for complete understanding of laser-cluster interactions, furthermore, we have measured total gas densities and cluster mass fractions besides the cluster size in this study. There was no accepted method for measuring the cluster mass fraction with the exception of frequency domain holography (FDH) recently reported [11]. In contrast, we have introduced another approach extracting the cluster mass fraction taking advantage of the Mie scattering theory which is applicable to large clusters, i.e., it was derived in combination with the Mie scattering method, an attenuation measurement of the laser beam intensity, and an interferometry.

In this paper, we consider the mathematical model in Sec. 2. This model is the generalization of the clusterization theory [12] on the case of the mixture of several gases and give calculated values of the different cluster parameters for various initial gas pressures and the gas compositions. In Sec. 3, we describe measurement and analytical methods using Mie scattering in order to investigate the spatial structure of the cluster-gas targets. In addition, the interferometer and the laser attenuation are employed to spatially measure the total gas densities and cluster densities. We describe the comparison of the numerical modeling and the experimental results in Sec. 4 and discuss in Sec 5. Such comparison shows that considered clusterization theory can be used successfully to model submicron-sized cluster production. Finally, we conclude in Sec. 6.

2. Mathematical modeling of the clusters origination in nozzle flows of the binary gas mixture

The mathematical model of the clusters formation in a single gas flow is described in details in [8] and [12]. Here we present its generalization for the case of the binary mixture with one inert component. In fact, this new model is based on the same physical suppositions, but takes into account the presence of the inert component.

We consider a gas mixture with two components: inert (which we denote by index i) and clusterizing, or condensing one (denoted by index c). The condensing component, in turn, is subdivided into gas phase (index g) and liquid (or condensed) one, denoted by index l. So, we have, in fact, three components of our medium: inert component i, condensing gas cg and clusters cl.

Since we assume that the clusters are comparatively small and do not move with respect to the surrounding gas, we may consider the composition of our mixture (namely, mass or molar relation between i and c components) to be constant and known. So, we introduce the value ϕ as the mass fraction of the condensing component. So, we can write for the densities of the components

ρc=ϕρ,ρi=(1ϕ)ρ.

We introduce also the so-called dryness degree β which is the mass fraction of the cg subcomponent in the whole c component. So, we can also write

ρcg=βρc=βϕρ,ρcl=(1β)ρc=(1β)ϕρ,ρi=(1ϕ)ρ.

In these expressions ρcg, ρcl, and ρi are densities of the components of the mixture, i.e., mass of a component per single volume of the whole mixture. Since the volume of the clusters is negligibly small, for the gas components these are also the densities of these components themselves, which appear in all thermodynamic relations for this component. For the clusters, of course, ρcl is much smaller than the density of the liquid phase itself, which we denote as ρl0.

The process of clusters formation is described in kinetics terms and, of course, this process is essentially nonequilibrium. But we consider two gas components i and cg to be in thermodynamic equilibrium with each other. So, we suppose that the temperature of two components of the gas phase is the same Ti = Tcg = T (and may differ from the temperature of the clusters Tcl). Another supposition is that the pressure of the mixture, in accordance with Dalton law, is the sum of the partial pressures of the gaseous components, P = Pi + Pcg. The liquid phase does not render the pressure, since it is not continuous.

These suppositions allow us to consider the density and internal energy of the whole mixture, not of the individual components. So, the governing system of equations practically does not differ from the single-gas case:

ρt+div(ρv)=0,
ρvt+div(ρvv)=gradP,
t(ρε+ρv22)+div(ρv(ε+v22))=divPv,
ρΩnt+div(ρvΩn)=Ir*n+r˙nρΩn1,n=0,1,2,
ρβt+div(ρvβ)=43πρl0Ir*3ϕ4πρρl0r˙Ω2ϕ.
Here Ωn are the moments of the distribution function of the clusters with respect to their radii:
ρΩn=0rnf(r,x,t)dr.

So, we can write for the third moment

1β=43ρl0πΩ3ϕ.

We use the same model of kinetics of the phase transitions as in the single-gas case, but with taking into account the presence of the inert component which changes the gasdynamics parameters but does not take part in the clusters origination process itself. To close the system Eqs. (3)(7), we use the following relationships. The critical radius

r*=2σ(T)ρl0RTln(Pcg/Ps(T)).

The nucleation rate

I=1ρl02σ(T)μπNA(PcgkT)2exp(4πσ(T)r*23kT).

The rate of clusters growth

r˙=Pcgρl02πRT(1TTs(Pcg)).

The model includes also some thermodynamic relationships which express the properties of the working gas and close the system of equations. To close the system, we need expressions P(ρ, ε, β), T(ρ, ε, β) and the expressions for some values at the saturation line. In the single-gas model we used expressions

P=P(ρ,ε,β)=ρε+(1β)Lsγβ(γ1)1.T=PβZρR,
which are the consequences of the simple equation of state of the ideal gas and the relationships between the heat of phase transition and the enthalpies of the phases,
Ls=HgHi.

In the gas mixture model, we apply Eq. (13) to each of the components i and c. As a result, instead of direct expressions Eq. (13) we obtain a system of algebraic equations, which includes Eq. (13) for both components, equality of the temperatures of the components and the relationship between the specific internal energies of the components:

ε=(1ϕ)εi+ϕεc.

In Eq. (13), we assume that the gas can be described by the equation of state of the ideal gas, and introduce the compressibility coefficient Z to take into account the properties of the actual working gas. When the gas is in the condition close to the saturation line, this supposition becomes very inaccurate. So, the use of some realistic equation of state is desirable. In fact, the equations of state of various gases are known, for example, they are available in NIST database [13].

To close the gas dynamics system of equations (or its modifications such as Eqs. (3)(7)), the equation of state should be resolved in the form P = P(ρ, ε) (or, for our case, P = P(ρ, ε, β)). But, usually, the equations of state of the real gases are given in the form of functions of some other pair of parameters, for example, (ρ, T), since for this pair of parameters there is a thermodynamic potential — Helmholtz energy. So, in order to use such an equation of state, we must solve a system of algebraic equations with respect to the unknown variables. In addition, since we consider the multiphase medium (even in the single-gas case), some additional relationships should be added to this system to obtain the equation of state of the whole medium. In the case of the single working gas which obeys the equation of state of the ideal gas, such a system may be resolved analytically (resulting in Eq. (13)), If the gas obeys some complicated equation of state, the system cannot be resolved analytically, and some numerical procedure is needed to resolve the system each time when the access to the equation of state takes place in the computation.

In the case of two gases, the numerical solution of an algebraic system is needed even if we consider both gases to be ideal. So, the transition to the real equation of state is not so complicated as in the case of the single gas.

In the computations, usually, we know the values ρ, ε, and β characterizing the whole mixture. To define other thermodynamic parameters, we solve the system

Tcl=Ts(Pcg),
Pi=Pi((1φ)ρ,T),
Pcg=Pcg(φβρ,T),
Pi+Pcg=Pcl(ρl0,Tcl),
ε=(1φ)εi((1φ)ρ,T)+φβεcg(φβρ,T)+φ(1β)εcl(ρl0,Tcl)
with respect to the values T, Tcl, Pi, Pcg, ρl0.

With use the model Eqs. (3)(13), we explored the clusters formation in the mixtures He+CO2, and H2+CO2. A series of one-dimensional computations has been performed with different initial pressures and concentrations of the condensing component. We took the three-staged nozzle [8]. In Tables 1, 2, 3 the results of these computations are presented. Table 1 contains the values of dryness at the nozzle’s outlet section, Table 2 presents the densities of clusters, and Table 3 gives the mean value of clusters’ radius. Comparison of the results shows that the presence of the inert component significantly changes the parameters of the clusters. While for the pure CO2 the characteristic values of dryness is about 74–90%, in the mixture the mass fraction of clusterized CO2 reaches 65–75%, i.e., the main part of the clusterizing component of the mixture can be turned to clusters.

Tables Icon

Table 1. Dryness for various molar concentrations of the condensing component and initial pressures.

Tables Icon

Table 2. Clusters densities, 109 cm−3, for various molar concentrations of the condensing component and initial pressures.

Tables Icon

Table 3. Cluster radius, μm, for various molar concentrations of the condensing component and initial pressures.

Since the gas flow in the nozzle is practically adiabatic, one usually considers that this process is characterized by constant entropy. It is the case for the single-gas flow as well as for the binary gas mixture. But, in the mixture, only the total entropy of the mixture is constant, while there is sufficient entropy exchange between the inert and condensing components. That is why the cluster formation process progresses completely different in the single-gas case and in the binary mixture, even if the initial partial pressure of the clusterizing component is the same as in the single-gas case. In other terms, the condensation process results in some heat release. In the single-gas case, all this heat of phase transition is absorbed by the condensing gas, its temperature increases (the condensation shock), and this results in deceleration of the condensation. Meanwhile, in the binary mixture, some part of the heat of phase transition is absorbed by the inert component, so, the heating of the condensing component is less than in the single-gas case. Consequently, the presence of the inert component facilitates the clusterization process.

In order to investigate the spatial structure of the cluster target, some of these variants were also computed in the 2D formulation (see below Sec. 2.1).

2.1. Spatial structure of cluster targets formed by gas jets

Although the processes of clusters origination and growth take place mostly inside the nozzle, some gasdynamical processes take place in the free jet exiting the nozzle, for instance, the lateral expansion of the jet into the surrounding vacuum. That, in turn, may result in some additional phase transitions and, as a result of these processes, the cluster target is not completely spatially homogeneous but has some spatial structure.

Usually, the nozzles with low expansion angles are used, and the flow in the outlet section is practically homogeneous. Nevertheless, since outside the nozzle the vacuum is maintained, downstream the nozzle outlet section the expansion of the jet is observed (Fig. 1). At the periphery of the jet, the gasdynamic rarefaction wave is formed. Thus, the gas density profiles at different distances from the nozzle’s outlet have different width: more far from the nozzle, the higher is the width of the jet. Besides, there is essential difference between the near and far sections. If we consider the sections more close to the nozzle than the point C (intersection of the leading edge of the rarefaction wave with the axis), there is a domain of unperturbed jet near the axis. So, the density profiles for different distances from the nozzle have a plateau near the r = 0 point. When we increase the distance from the nozzle, the plateau becomes narrower, but the maximum value of density (e.g., the height of this plateau) practically remains the same, since the unperturbed jet exiting the nozzle is practically homogeneous. On the contrary, for the sections more far from the nozzle than the point C we observe the decrease of the maximum density value with the moving off from the nozzle. At Fig. 2 we present the results of numerical modeling which confirm these positions.

 figure: Fig. 1

Fig. 1 Scheme of spatial structure of a supersonic jet expanding to vacuum and density profiles at different distances from the nozzle.

Download Full Size | PDF

 figure: Fig. 2

Fig. 2 Spatial structure of the supersonic CO2 jet (numerical results, initial pressure 20 bar).

Download Full Size | PDF

The above discussed effects concern with the gas dynamics and do not depend on the presence of the clusters in the jet. Of course, the clusters which are contained in the jet expand along with the jet. Figure 2 also shows the profiles of the clusters concentration, and one can see that in the most typical case the clusters concentration profiles are like the gas density profiles.

But since in the rarefaction wave the isentropic decrease of the gas pressure and temperature goes on, sometimes the effect of repeated clusterization takes place (Fig. 3). In that case, the clusters concentration profile is not monotonous, but there is a considerable peak at the jet’s periphery, namely in the domain of the rarefaction wave. Also it is observed the decrease of the mean cluster size in that place. It is due to the origination of a large amount of new clusters with very small sizes (in comparison with the old clusters originated inside the nozzle).

 figure: Fig. 3

Fig. 3 Molecules and clusters concentration profiles for the mixture of 10% CO2 and 90% He, initial pressure 50 bar (numerical results).

Download Full Size | PDF

Since our mathematical model operates with the distribution function moments Ωn the standard deviation (S.D.) of the cluster size can be easily obtained, as well as mean radius of cluster:

<r>=Ω1/Ω0,S.D.<r>=(Ω2/Ω0Ω12/Ω02)1/2.

Figure 4 shows the spatial distribution of the mean cluster size and its standard deviation for the case of pure CO2 (when no off-nozzle clusterization was observed, Fig. 2) and for the case of CO2 + He mixture, when some off-nozzle clusterization took place, Fig. 3. One can see that the repeated clusterization causes higher variation of the S.D., because the spatial structure of the flow is more complex in that case.

 figure: Fig. 4

Fig. 4 Spatial distribution of the mean cluster size and the standard deviation for the 20 bar pure CO2 (upper pair of graphs) and for the 50 bar 10% CO2 + 90% He mixture.

Download Full Size | PDF

Our mathematical model of the cluster targets formation allows us to explore the spatial structure of the target as well as the target formation processes inside the nozzle.

The spatial extent of the gas-cluster medium suitable for the cluster target is higher than the diameter of the nozzle’s opening, especially at large distances from the nozzle. This fact is also directly observed in the experiment.

Calculation results obtained for mixtures of CO2 with H2 and He are presented in Figs. 5,6 for different gas pressures and distances from the nozzle exit. The calculations were made for 3-stage nozzle [8]. As it was shown earlier [8] such nozzle is optimized for creation of large clusters.

 figure: Fig. 5

Fig. 5 Calculation results obtained for mixtures of 30% CO2 with 70% H2 for different distances from the nozzle exit. Initial gas pressure is 60 bar.

Download Full Size | PDF

 figure: Fig. 6

Fig. 6 Calculation results obtained for mixtures of 10% CO2 with 90% He for different distances from the nozzle exit. Initial gas pressure is 60 bar.

Download Full Size | PDF

The numerical modeling obtained in the above manner is verified experimentally in the next sections.

3. Measurement and analytical methods

3.1. Derivation of cluster size distribution

We have already reported on the development of an apparatus capable of measuring micron-sized particles using Mie scattering theory, and the size measurements of submicron-sized CO2 clusters in the gas mixtures have been presented [10]. Here, in addition to the Mie scattering method, the interferometry and the laser attenuation are newly explained. Figure 7 shows the experimental setup for the measurements of the particle sizes using Mie scattering and the total gas density using the Michelson interferometer with a roof prism. The second harmonic (532 nm) of Nd:YAG laser (Spectra Physics, Quanta-ray) is split into the lines for the size measurement and the density measurement of the cluster-gas target.

 figure: Fig. 7

Fig. 7 Optical design for measurements of particle sizes using Mie scattering and total gas densities using a Michelson-type interferometer.

Download Full Size | PDF

In order to acquire angular distribution of the scattered light from the target, a photo-multiplier (PMT 1) in a light shielding box is mounted on a movable stage. Another PMT 2 is placed perpendicular to the laser axis to detect a reference signal so as to compensate intensity fluctuations caused by the target samples. The position dependence of the scattered signal intensity in this measurement can be obtained by shifting the nozzle position for the laser axis. Attenuation of the laser beam energy due to the presence of the cluster-gas target is also measured to obtain the cluster density.

The size distribution of the clusters is analyzed in a following procedure: the scattering co-efficient F(x, θ) strongly depends on the size parameter x = 2πrn/λ according to the Mie scattering theory, where θ, r, n, and λ are scattering angle, particle radius, refractive index of medium (= 1 in vacuum), and wavelength, respectively. The F(x, θ) is calculated applying an open Mie scattering code [14]. The size distribution of particle N(x) is obtained by least-square fit of the angular distribution of the scattered light intensity I(θ) from following equation,

I(θ)I0=αxF(x,θ)N(x),
where, α is the unknown apparatus constant, I0 is the incident laser energy. In present study, N(x) is assumed as a lognormal function [15], whose average and variance are obtained by fitting algorithm of the Nelder-Mead method. In present case, the N(x) has the dimension of the surface density projected along the laser axis.

In the above procedure, only relative cluster size distribution αN(x) can be obtained. Then, the absolute cluster density N(x) can be calculated by the following formula by measuring the transmitted light intensity It,

ItI0=exp(α1xCext(x)αN(x)),
where, the Cext(x) is an extinction cross-section which can be calculated according to the Mie scattering theory. The α−1 is determined to satisfy this formula, and then the N(x) can be found.

3.2. Spatial distribution of clusters

The nozzle can be scanned perpendicular to the laser axis in order to obtain the spatial distribution of clusters. As shown in Fig. 8, the angular distribution of the scattered light intensities I(θ, yi) are measured at the position yi on the axis perpendicular to the laser in the scattering plane. Because the PMT can accept the light scattered from the target in the overall laser pass, the scattering intensities I(θ, yi) should be transformed into the emission coefficients ε(θ, rj) in the radial direction rj of the target by employing the simple discretization of Abel inversion as following equation [16],

ε(θ,rj)=1πΔyi=jN1I(θ,yi+1)I(θ,yi)(i+12)2+j2,
where, Δy is the spatial separation between the data, which is the laser scan interval in this case. The radial separation Δr is expressed to be Δr = Δy. Using the ε(θ, rj) instead of I(θ) of the Eq. (22), the number densities of clusters at the radial position, which has the dimension of the volume density, are obtained by the similar fitting procedure.

 figure: Fig. 8

Fig. 8 Schematic view of Abel inversion.

Download Full Size | PDF

3.3. Interferometry

For measuring the total gas density, the Michelson interferometer with the roof prism was adopted on another laser line with collimated beam as shown in Fig. 7. A probe light traversing the cluster-gas target medium interferes with a reference light passing the side of the target. The image is magnified by a telescope and recorded by a CCD camera. The interferogram produces a planar phase shift image resulting from the passage of the light across the cylindrically symmetric medium. The radial profile of the refractive index of the target can be obtained via Abel inversion of this phase shift with a Hankel-Fourier method [16].

The refractive index is dominated by the density and the polarizability of the target particles. A cluster polarizability-volume αcl can be expressed as the product of the number of molecules per cluster N# and molecular polarizability-volume αcg as follows [17, 18],

αcl=N#αcg.
In the present case, assuming the homogeneous gas mixture of inert component i, condensing gas cg, and clusters cl, the Lorentz-Lorenz relation between the radial refractive index and gas density can be modified as follows,
34πn21n2+2=αclnclus+αcgρcg+αiρi
=N#αcgρclN#+αcgρcg+αiρi
=αcgρcl+αcgρcg+αiρi
=(1β)αcgρc+βαcgρc+αiρi
=αcgρc+αiρi,
where, nclus is the cluster concentration. In the cases of the present experiments, the polarizability-volumes of the inert components αi are 0.802 × 10−24 cm3 for H2 and 0.205 × 10−24 cm3 for helium [19], and the polarizability-volume of the condensing gas αcg is 2.61 × 10−24 cm3 for CO2[20]. Note that the polarizability-volumes of CO2 and H2 are averaged in terms of orientation, respectively. Also, total gas densities have the following relationship to the partial pressures of gas mixture,
ρc/ρc0:ρi/ρi0=pc:pi,
where, ρc0 and ρi0 are the densities at the standard temperature and pressure, and pc and pi are the partial gas pressures of the condensing and the inert components, respectively. Finally, we can calculate the profile of the total gas densities in radial direction from the Eqs. (30), (31), and the refractive index.

4. Results and comparison with numerical modeling

4.1. Characterization of 30% CO2 + 70% H2 mixture

We have characterized the CO2 clusters formed in the supersonic expansion of the 30% CO2 + 70% H2 mixture through the three-staged conical nozzle at stagnation pressure of 60 bar. We have acquired the angular distributions of the scattered light at the positions of y = 0, 0.2, 0.4, 0.6, 0.8, and 1.0 mm from the target center by shifting the nozzle perpendicular to the laser axis at the distance of 1 mm from the nozzle exit, respectively. As shown in Fig. 9(a), the angular distributions at the positions divided radially were obtained through the Abel inversion of those raw data according to the Eq. (24). The distributions are featured by the forward scattering caused by Mie scattering, indicating that CO2 clusters with the sizes of the laser wavelength scale exist in the cluster-gas target. Note that the contribution of Rayleigh scattering from monomer gases is usually negligible small because the scattering intensity is proportional to the sixth power of the particle size. As shown in Fig 9(b), the lognormal size distributions are obtained by the fitting procedure. The mean size of CO2 clusters is estimated to be 0.28 ± 0.03 μm with the standard deviation (S.D.) of 0.13 ± 0.02 μm. It is found that the variation of the mean cluster size inside the gas jet is constant within the experimental error. On the other hand, the numerical modeling predicts that the mean cluster size is 0.588 μm (S.D. 0.090 μm) independent of the radial positions. The reason for this discrepancy will be discussed in Sec. 5. We note that the conventional Rayleigh scattering method could not be applicable for the estimation of our submicron-sized clusters because the scattered intensity did not follow the simple sixth power law to the particle size.

 figure: Fig. 9

Fig. 9 (a) The angular distributions of light scattered from the CO2 clusters at the radial position of 0, 0.2, 0.4, 0.6, 0.8, and 1.0 mm from the target center at 1 mm distance from the nozzle exit for the 60 bar 30% CO2 + 70% H2 mixture, respectively. (b) The lognormal size distributions are estimated from the angular distributions of the scattered light.

Download Full Size | PDF

4.2. Cluster density

Measuring the transmission of the laser in the target as described in Sec. 3.1, we can obtain the surface density of the clusters projected along the laser axis. Figure 10 shows the normalized laser intensity taken with no cluster-gas target present and with the cluster-gas target source on. Attenuation of the laser energy due to the presence of the cluster-gas target is clear. The mean transmission is found to be 0.787 ± 0.008 with the target, which results that the CO2 cluster density is estimated to be 5.5 × 108 clusters/cm2 or the CO2 density of the cluster phase is corresponding to 1.1 × 1017 molecules/cm2. Because the relative cluster concentration in the radial direction per unit length is found from the previous section, the surface density can be modified to the volume density of the clusters as a function of the radial direction as shown in Fig. 11.

 figure: Fig. 10

Fig. 10 Normalized laser intensity taken with no cluster-gas target present and with the cluster-gas target source on.

Download Full Size | PDF

 figure: Fig. 11

Fig. 11 CO2 density of the cluster phase as a function of the radius from the target center axis at the distance z = 1 mm from the nozzle exit.

Download Full Size | PDF

4.3. Total gas density profile

Figure 12(a) shows the fringe shift image near the nozzle exit using the Michelson interferometer. As shown in Fig 12(c), the gas density profile with cylindrical symmetry is obtained via Abel inversion of the planar phase shift image (Fig 12(b)), which is derived from the fringe shift image. Where, the radius is the distances from the center axis of the nozzle and the z is the distance from the nozzle exit. As can be seen from Fig 12(c), the gases diffuse in the radial direction with an increase in the distance of the z axis. Figures 12(d)–(g) show the number densities of CO2 and H2 molecules as a function of the radius at z = 0.75, 1.0, 1.5, and 2.0 mm, defined by interferometry (markers) and by numerical modeling (lines), respectively. The error bars are derived from the difference between the reconstructed phase shift and the original phase shift. The error at a position near to the center generally becomes large due to accumulation of the propagation of the errors at its outer positions in principle of Abel inversion. As typically shown in Figs 12(d)–(g), the numerical modeling predicts an approximately collimated gas flow at 7.59 ∼ 6.74 × 1018 cm−3 for CO2 and 1.77 ∼ 1.57 × 1019 cm−3 for H2 in region of z = 0.75 ∼ 2.0 mm, respectively. By contrast, it is found that the plateau regions in the experimental results are narrower than those of the numerical modeling. Namely, it means the leading edge of the rarefaction wave discussed in Sec. 2.1 extends wider than numerical modeling. Although the gas expansion profiles obtained from the experiment and the model are different, the heights of the densities are consistent within a factor of two. The reason for the different of the gas density profile will be discussed in Sec. 5.

 figure: Fig. 12

Fig. 12 (a) The fringe shift image at the nozzle exit obtained by using the interferometer. The CO2/H2 mixture at 60 bar was used. The z axis means the distance from the nozzle orifice. The y axis refers to a direction transverse to the jet. (b) The planar phase shift derived from the fringe shift image. The maximum phase shift is 2.15π rad. (c) The gas density profile obtained as a function of the radius via Abel inversion. The radius means the distance from the center axis of the nozzle. (d),(e),(f),(g) The number densities of CO2 and H2 molecules at z = 0.75, 1.0, 1.5, and 2.0 mm, defined by interferometry (markers) and by numerical modeling (lines), respectively.

Download Full Size | PDF

4.4. Dryness

The dryness β, i.e., the ratio of the CO2 density of the gas phase to the total density of CO2 molecules, is described again,

β=1ρcl/ρc.
Substituting the CO2 density of the cluster phase shown in Fig. 11 into the numerator, and the total density of CO2 molecules shown in Fig. 12(e) into the denominator of the right hand, the variation of the dryness for the radius from the center axis of the nozzle at z = 1 mm is shown in Fig. 13. The dryness at z = 1 mm in the numerical modeling stably lies at 0.57, which is lower than the experimental results, and increases at the region larger than 4 mm in the radius as shown in Fig 5. By contrast, the dryness in the experiment gradually increases with the radius of the target. Because the plateau region in the density profile is narrow and the boundary of the target is fuzzy as compared to the model, such behavior was observed. The reason for this discrepancy will be discussed in Sec. 5.

 figure: Fig. 13

Fig. 13 The variation of the dryness β for the radius from the center axis of the nozzle at z = 1 mm. The broken line is the values by the numerical modeling.

Download Full Size | PDF

4.5. Characterization of 10% CO2 + 90% He mixture

The same measurements were also performed for CO2 clusters formed in the supersonic expansion of 60 bar 10% CO2 + 90% He mixture. Figure 14(a) shows the angular distributions of scattered light intensity from CO2 clusters and (b) the cluster size distribution at the positions of y = 0, 0.2, 0.4, 0.6, 0.8, and 1.0 mm from the target center, respectively. The mean diameter of CO2 cluster is estimated to be 0.26 ± 0.04 μm (S.D. 0.08 ± 0.01 μm) by the fitting procedure. We found that the clusters grow a certain size regardless of the position of the gas expansion aside from its concentration,. The numerical modeling predicts that the size of CO2 clusters is 0.360 μm (S.D. 0.072 μm) for the case of CO2/He mixture at 60 bar.

 figure: Fig. 14

Fig. 14 (a) The angular distributions of the light scattered from the CO2 clusters in 10% CO2 + 90% He mixture targets for the stagnation pressure of 60 bar. (b) The lognormal size distribution estimated from the angular distribution.

Download Full Size | PDF

The mean transmission of the laser passing through the target is 0.707±0.009, which results that the CO2 cluster density is estimated to be 8.2 × 108 clusters/cm2 or the CO2 density of the cluster phase is corresponding to 2.0 × 1017 molecules/cm2.

By the relative abundance of clusters at the different radial positions shown in Fig. 14(b), the surface density is modified to the volume densities of the clusters as a function of the radial direction as shown in Fig. 15(a). Figure 15(b) shows the number densities of CO2 and He molecules as a function of the radius at z = 1.0 mm defined by interferometry (markers) and by numerical modeling (lines), respectively. The variation of the dryness for the radius from the center axis of the nozzle at z = 1 mm is shown in Fig. 15(c). The dryness by the numerical modeling stably lies at 0.32, which is about one-half of the experimental results, and increases at the region larger than 2 mm in the radius as shown in Fig 6. The reason why the dryness in the experiment slightly increases at the outer of the target is the same reason mentioned for the CO2/H2 target. Nevertheless the gas density of CO2 is as low as 10% of the total gas density, the absolute abundance of the clusters is greater than that for the CO2/H2 gas mixture, i.e., the dryness is small, which supports the numerical model. The difference between the experimental result and the model will be discussed in Sec. 5.

 figure: Fig. 15

Fig. 15 (a) CO2 density of the cluster phase as a function of the radius from the target center axis at the distance z = 1 mm from the nozzle orifice. (b) The number densities of CO2 and H2 molecules at z = 1.0 mm, defined by interferometry (markers) and by numerical modeling (lines), respectively. (c) The variation of the dryness β for the radius from the center axis of the nozzle at z = 1 mm. The broken line is the values by the numerical modeling.

Download Full Size | PDF

5. Discussion

The discrepancies between the experiment and the numerical modeling are recognized in the cluster size and its concentration. Its possible reasons are explained in this section. In the size measurement, the total error coming from the measurement and the fitting processes is 10% [10]. Therefore, we think that the possible errors which originate from the measurements and the data fitting processes are small and cannot be the main reason for the discrepancy.

Secondly, concerning the uncertainties arising from the mathematical model of the cluster formation, it should be noted that the model, of course, has some its own limitations and assumptions. For example, it applies the thermodynamical approach to very small clusters of the liquid phase. There are some problems with adequate thermodynamical description of the metastable gas state (the clusters are formed in the supercooled gas). Therefore, we think that the inaccuracy of about 2 times may be not too high for such a model.

On the other hand, the reasons may be in some geometrical inaccuracies in the nozzle shape. The gas expansion conditions are very sensitive to the fabrication accuracy of the nozzle. For instance, according to our computations, the clusters size is highly dependent on the nozzle’s expansion angle. It should be mentioned that the creation of the 3-stage nozzle with given geometrical parameters is the complex practical problem. It is not easy especially to make the last cone which has small solid angle (about 2°) and long size. Production accuracy was about 30% and it can entirely explain the difference between measurement and calculations. Moreover, we have found that the nozzle with very low expansion rate is needed for formation of large clusters. Therefore, any excess of the expansion rate, caused by some inaccuracies in the nozzle fabrication, may result in considerable diminishing of the cluster size.

In addition, the model set a boundary condition that is zero normal velocity to the nozzle’s wall. When there is the roughness of the nozzle inside, the ideal gas flow assumed in the model is not realized. The degradation of the plastic poppet inside the pulsed solenoid valve may also change the ideal gas flow. The possibility of these effects cannot be denied and may result in considerable diminishing of the cluster size.

As mentioned above, the assumption of the initial parameters in the model, the limitation of the thermodynamic approach, and the some inaccuracies in the nozzle fabrication are the main factors of the discrepancy between the measurements and the model.

6. Conclusion

We have presented the generalization of the clusterization theory on the case of the mixture of several gases and give calculated values of the different cluster parameters for various initial gas pressures and the gas compositions. The sizes of the submicron-sized clusters, which formed through the three-staged conical nozzle designed based on the 2D hydrodynamic calculations, were measured using the Mie scattering method.

The mean sizes of CO2 clusters for the cases of CO2/H2 and CO2/He gas mixtures are estimated to be 0.28 ± 0.03 and 0.26 ± 0.04 μm, respectively. The CO2 densities of cluster phase are estimated to be 1.1 × 1017 molecules/cm2 for CO2/H2 and 2.0 × 1017 molecules/cm2 for CO2/He by measuring the attenuation of the laser beam intensity. The total gas densities are measured by the Mickelson-type interferometer with the roof prism. The gas densities are verified to agree with the numerical modeling within a factor of two. The dryness can be spatially derived in combination with the Mie scattering method, the attenuation measurement of the laser beam intensity, and the interferometry. The distributions of the mean cluster size and the dryness inside the gas jet support the mathematical model, which claims the homogeneous cluster formation.

Thus, we have characterized the submicron-sized CO2 clusters and confirmed that our mathematical model is enough reliable for the production of submicron-sized clusters in supersonic gas expansion. This work gives the initial target state to simulate the physics of a variety of radiations generated in the laser-cluster interactions. Then, it contributes to the optimization for the x-ray emission, the laser-driven ion accelerations, and so on.

Acknowledgments

This work was supported by the Funding Program for Next Generation World-Leading Researchers (NEXT Program) from the Japan Society for the Promotion of Science (JSPS) and the grant RFBR 12-02-91169-GFEN-a.

References and links

1. T. Fennel, K.-H. Meiwes-Broer, J. Tiggesbäumker, P.-G. Reinhard, P. M. Dinh, and E. Suraud, “Laser-driven nonlinear cluster dynamics,” Rev. Mod. Phys. 82, 1793–1842 (2010). [CrossRef]  

2. T. Ditmire, J. Zweiback, V. P. Yanovsky, T. E. Cowan, G. Hays, and K. B. Wharton, “Nuclear fusion from explosions of femtosecond laser-heated deuterium clusters,” Nature 398, 489–492 (1999). [CrossRef]  

3. E. Parra, I. Alexeev, J. Fan, K. Y. Kim, S. J. McNaught, and H. M. Milchberg, “X-ray and extreme ultraviolet emission induced by variable pulse-width irradiation of Ar and Kr clusters and droplets,” Phys. Rev. E 62, R5931–R5934 (2000). [CrossRef]  

4. Y. Fukuda, Y. Akahane, M. Aoyama, N. Inoue, H. Ueda, Y. Nakai, K. Tsuji, K. Yamanaka, Y. Hironaka, H. Kishimura, H. Morishita, K. Kondo, and K. G. Nakamura, “Relativistic laser plasma from micron-sized argon clusters as a debris-free x-ray source for pulse x-ray diffraction,” Appl. Phys. Lett. 85, 5099–5101 (2004). [CrossRef]  

5. L. M. Chen, W. C. Yan, D. Z. Li, Z. D. Hu, L. Zhang, W. M. Wang, N. Hafz, J. Y. Mao, K. Huang, Y. Ma, J. R. Zhao, J. L. Ma, Y. T. Li, X. Lu, Z. M. Sheng, Z. Y. Wei, J. Gao, and J. Zhang, “Bright betatron X-ray radiation from a laser-driven-clustering gas target,” Sci. Rep. 3, 1912 (2013).

6. Y. Fukuda, A. Y. Faenov, M. Tampo, T. A. Pikuz, T. Nakamura, M. Kando, Y. Hayashi, A. Yogo, H. Sakaki, T. Kameshima, A. S. Pirozhkov, K. Ogura, M. Mori, T. Z. Esirkepov, J. Koga, A. S. Boldarev, V. A. Gasilov, A. I. Magunov, T. Yamauchi, R. Kodama, P. R. Bolton, Y. Kato, T. Tajima, H. Daido, and S. V. Bulanov, “Energy increase in multi-MeV ion acceleration in the interaction of a short pulse laser with a cluster-gas target,” Phys. Rev. Lett. 103, 165002 (2009). [CrossRef]   [PubMed]  

7. Y. Fukuda, H. Sakaki, M. Kanasaki, A. Yogoa, S. Jinno, M. Tampo, A. Faenov, T. Pikuz, Y. Hayashi, M. Kando, A. Pirozhkov, T. Shimomura, H. Kiriyama, S. Kurashima, T. Kamiya, K. Oda, T. Yamauchi, K. Kondo, and S. Bulanov, “Identification of high energy ions using backscattered particles in laser-driven ion acceleration with cluster-gas targets,” Radiat. Meas. 50, 92–96 (2013). [CrossRef]  

8. A. S. Boldarev, V. A. Gasilov, A. Y. Faenov, Y. Fukuda, and K. Yamakawa, “Gas-cluster targets for femtosecond laser interaction: Modeling and optimization,” Rev. Sci. Instrum. 77, 083112 (2006). [CrossRef]  

9. Y. Fukuda, A. Y. Faenov, T. Pikuz, M. Kando, H. Kotaki, I. Daito, J. Ma, L. M. Chen, T. Homma, K. Kawase, T. Kameshima, T. Kawachi, H. Daido, T. Kimura, T. Tajima, Y. Kato, and S. V. Bulanov, “Soft x-ray source for nanostructure imaging using femtosecond-laser-irradiated clusters,” Appl. Phys. Lett. 92, 121110 (2008). [CrossRef]  

10. S. Jinno, Y. Fukuda, H. Sakaki, A. Yogo, M. Kanasaki, K. Kondo, A. Y. Faenov, I. Y. Skebelv, T. A. Pikuz, A. S. Boldarev, and V. A. Gasilov, “Characterization of submicron-sized CO2clusters formed with a supersonic expansion of a mixed-gas using a three-staged nozzle,” Appl. Phys. Lett. 102, 164103 (2013). [CrossRef]  

11. X. Gao, X. Wang, B. Shim, A. V. Arefiev, R. Korzekwa, and M. C. Downer, “Characterization of cluster/monomer ratio in pulsed supersonic gas jets,” Appl. Phys. Lett. 100, 064101 (2012). [CrossRef]  

12. A. S. Boldarev, V. A. Gasilov, and A. Y. Faenov, “On the generation of large clusters in forming gas-jet targets for lasers,” Tech. Phys. 49, 388 (2004). [CrossRef]  

13. http://webbook.nist.gov/chemistry/fluid.

14. W. J. Wiscombe, “Mie scattering calculations: Advances in technique and fast, vector-speed computer codes,” Ncar technical note 140+STR 1979 (edited/revised 1996).

15. M. Kerker, The Scattering of Light and Other Electromagnetic Radiation (Academic Press, New York and London, 1969).

16. R. Álvarez, A. Rodero, and M. C. Quintero, “An abel inversion method for radially resolved measurements in the axial injection torch,” Spectrochim. Acta Part B 52, 1665 (2002). [CrossRef]  

17. H. M. Milchberg, S. J. McNaught, and E. Parra, “Plasma hydrodynamics of the intense laser-cluster interaction,” Phys. Rev. E 64, 056402 (2001). [CrossRef]  

18. H.-Y. Kim, J. O. Sofo, D. Velegol, M. W. Cole, and G. Mukhopadhyay, “Static polarizabilities of dielectric nanoclusters,” Phys. Rev. A 72, 053201 (2005). [CrossRef]  

19. D. R. Lide, ed., Handbook of Chemistry and Physics, 88th edition(CRC, New York, 1997).

20. G. Maroulis and A. J. Thakkar, “Polarizabilities and hyperpolarizabilities of carbon dioxide,” J. Chem. Phys. 93, 4164–4171 (1990). [CrossRef]  

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 (15)

Fig. 1
Fig. 1 Scheme of spatial structure of a supersonic jet expanding to vacuum and density profiles at different distances from the nozzle.
Fig. 2
Fig. 2 Spatial structure of the supersonic CO2 jet (numerical results, initial pressure 20 bar).
Fig. 3
Fig. 3 Molecules and clusters concentration profiles for the mixture of 10% CO2 and 90% He, initial pressure 50 bar (numerical results).
Fig. 4
Fig. 4 Spatial distribution of the mean cluster size and the standard deviation for the 20 bar pure CO2 (upper pair of graphs) and for the 50 bar 10% CO2 + 90% He mixture.
Fig. 5
Fig. 5 Calculation results obtained for mixtures of 30% CO2 with 70% H2 for different distances from the nozzle exit. Initial gas pressure is 60 bar.
Fig. 6
Fig. 6 Calculation results obtained for mixtures of 10% CO2 with 90% He for different distances from the nozzle exit. Initial gas pressure is 60 bar.
Fig. 7
Fig. 7 Optical design for measurements of particle sizes using Mie scattering and total gas densities using a Michelson-type interferometer.
Fig. 8
Fig. 8 Schematic view of Abel inversion.
Fig. 9
Fig. 9 (a) The angular distributions of light scattered from the CO2 clusters at the radial position of 0, 0.2, 0.4, 0.6, 0.8, and 1.0 mm from the target center at 1 mm distance from the nozzle exit for the 60 bar 30% CO2 + 70% H2 mixture, respectively. (b) The lognormal size distributions are estimated from the angular distributions of the scattered light.
Fig. 10
Fig. 10 Normalized laser intensity taken with no cluster-gas target present and with the cluster-gas target source on.
Fig. 11
Fig. 11 CO2 density of the cluster phase as a function of the radius from the target center axis at the distance z = 1 mm from the nozzle exit.
Fig. 12
Fig. 12 (a) The fringe shift image at the nozzle exit obtained by using the interferometer. The CO2/H2 mixture at 60 bar was used. The z axis means the distance from the nozzle orifice. The y axis refers to a direction transverse to the jet. (b) The planar phase shift derived from the fringe shift image. The maximum phase shift is 2.15π rad. (c) The gas density profile obtained as a function of the radius via Abel inversion. The radius means the distance from the center axis of the nozzle. (d),(e),(f),(g) The number densities of CO2 and H2 molecules at z = 0.75, 1.0, 1.5, and 2.0 mm, defined by interferometry (markers) and by numerical modeling (lines), respectively.
Fig. 13
Fig. 13 The variation of the dryness β for the radius from the center axis of the nozzle at z = 1 mm. The broken line is the values by the numerical modeling.
Fig. 14
Fig. 14 (a) The angular distributions of the light scattered from the CO2 clusters in 10% CO2 + 90% He mixture targets for the stagnation pressure of 60 bar. (b) The lognormal size distribution estimated from the angular distribution.
Fig. 15
Fig. 15 (a) CO2 density of the cluster phase as a function of the radius from the target center axis at the distance z = 1 mm from the nozzle orifice. (b) The number densities of CO2 and H2 molecules at z = 1.0 mm, defined by interferometry (markers) and by numerical modeling (lines), respectively. (c) The variation of the dryness β for the radius from the center axis of the nozzle at z = 1 mm. The broken line is the values by the numerical modeling.

Tables (3)

Tables Icon

Table 1 Dryness for various molar concentrations of the condensing component and initial pressures.

Tables Icon

Table 2 Clusters densities, 109 cm−3, for various molar concentrations of the condensing component and initial pressures.

Tables Icon

Table 3 Cluster radius, μm, for various molar concentrations of the condensing component and initial pressures.

Equations (32)

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

ρ c = ϕ ρ , ρ i = ( 1 ϕ ) ρ .
ρ c g = β ρ c = β ϕ ρ , ρ c l = ( 1 β ) ρ c = ( 1 β ) ϕ ρ , ρ i = ( 1 ϕ ) ρ .
ρ t + div ( ρ v ) = 0 ,
ρ v t + div ( ρ v v ) = grad P ,
t ( ρ ε + ρ v 2 2 ) + div ( ρ v ( ε + v 2 2 ) ) = div P v ,
ρ Ω n t + div ( ρ v Ω n ) = I r * n + r ˙ n ρ Ω n 1 , n = 0 , 1 , 2 ,
ρ β t + div ( ρ v β ) = 4 3 π ρ l 0 I r * 3 ϕ 4 π ρ ρ l 0 r ˙ Ω 2 ϕ .
ρ Ω n = 0 r n f ( r , x , t ) d r .
1 β = 4 3 ρ l 0 π Ω 3 ϕ .
r * = 2 σ ( T ) ρ l 0 R T ln ( P c g / P s ( T ) ) .
I = 1 ρ l 0 2 σ ( T ) μ π N A ( P c g k T ) 2 exp ( 4 π σ ( T ) r * 2 3 k T ) .
r ˙ = P c g ρ l 0 2 π R T ( 1 T T s ( P c g ) ) .
P = P ( ρ , ε , β ) = ρ ε + ( 1 β ) L s γ β ( γ 1 ) 1 . T = P β Z ρ R ,
L s = H g H i .
ε = ( 1 ϕ ) ε i + ϕ ε c .
T c l = T s ( P c g ) ,
P i = P i ( ( 1 φ ) ρ , T ) ,
P c g = P c g ( φ β ρ , T ) ,
P i + P c g = P c l ( ρ l 0 , T c l ) ,
ε = ( 1 φ ) ε i ( ( 1 φ ) ρ , T ) + φ β ε c g ( φ β ρ , T ) + φ ( 1 β ) ε c l ( ρ l 0 , T c l )
< r > = Ω 1 / Ω 0 , S . D . < r > = ( Ω 2 / Ω 0 Ω 1 2 / Ω 0 2 ) 1 / 2 .
I ( θ ) I 0 = α x F ( x , θ ) N ( x ) ,
I t I 0 = exp ( α 1 x C ext ( x ) α N ( x ) ) ,
ε ( θ , r j ) = 1 π Δ y i = j N 1 I ( θ , y i + 1 ) I ( θ , y i ) ( i + 1 2 ) 2 + j 2 ,
α c l = N # α c g .
3 4 π n 2 1 n 2 + 2 = α c l n clus + α c g ρ c g + α i ρ i
= N # α c g ρ c l N # + α c g ρ c g + α i ρ i
= α c g ρ c l + α c g ρ c g + α i ρ i
= ( 1 β ) α c g ρ c + β α c g ρ c + α i ρ i
= α c g ρ c + α i ρ i ,
ρ c / ρ c 0 : ρ i / ρ i 0 = p c : p i ,
β = 1 ρ c l / ρ c .
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.