Hostname: page-component-745bb68f8f-d8cs5 Total loading time: 0 Render date: 2025-02-11T15:14:21.495Z Has data issue: false hasContentIssue false

Laser-induced migration and isotope separation of epi-thermal monomers and dimers in supercooled free jets

Published online by Cambridge University Press:  07 June 2005

JEFF W. EERKENS
Affiliation:
CRISLA Research Laboratory, Nuclear Science and Engineering Institute, University of Missouri, Columbia, Missouri
Rights & Permissions [Opens in a new window]

Abstract

Explicit relations are developed to estimate the outflux of migrating isotopomers iQF6 to the outskirts of a supersonic supercooled free jet whose core is irradiated by a co-axial laser beam and intercepted by a skimmer that separates core gas from peripheral gases. The QF6 target gas is diluted in carrier gas G (G = He, N2, Ar, Xe, SF6, etc.) which determines the jet's supersonic characteristics and forms QF6:G dimers at low temperatures. Under isotope-selective laser excitation, excited iQF6* convert their vibrational energy V into kinetic energy T after forming transient iQF6*:G dimers that dissociate in sub-microseconds. Three migrating groups with different transport parameters are created in the jet: thermal monomers, faster-moving epithermal monomers, and slower-moving dimers. Jet-core-fleeing QF6 is enriched in iQF6 due to enhanced outwards migration of iQF6! epithermals and reduced escape of jQF6:G dimers in the jet. Isotope enrichments in the rim gases are highest for heavier carrier gases such as G = Xe or G = SF6.

Type
Research Article
Copyright
2005 Cambridge University Press

1. INTRODUCTION

The present paper is a sequel to Eerkens (1998), providing more complete calculations of dimer-mediated isotope enrichment effects from vibrationally laser-excited iQF6 molecules in supercooled gas streams. The laser-irradiated QF6 vapor (Q = S, Se, Te, Mo, W, U, etc.) is dispersed through carrier gas G (e.g., G = H2, He, N2, Ar, Xe, SF6, SiBr4) with which it can form QF6:G dimers, and is supercooled by supersonic expansion in a free jet. In molecular laser isotope separation (MLIS) applications, iQF6 isotopomers are then selectively excited via their ν3 vibrational absorption bands that are isotope-shifted from those of jQF6 species.

In one MLIS scheme employing supersonic free jets, jet-fleeing iQF6 molecules are harvested after excited iQF6*:G dimer predissociations and VT conversions take place, and formation of non-excited iQF6:G is reduced relative to jQF6:G dimers (Lee, 1977; VandenBergh, 1985; Eerkens, 1998). After passage through an irradiation chamber, the jet enters a skimmer that separates the jet's core flow from peripheral gases which are pumped out separately as shown in Figure 1. This MLIS harvesting concept was first conceived by Y.T. Lee (1977). Lee originally proposed to directly excite and dissociate already dimerized species such as QF6:G or (QF6)2. However excitation of QF6 monomers, which subsequently dimerize and predissociate in the same manner is more effective (Eerkens, 2001a). An advantage of the latter scheme is that monomers have stronger absorption peaks than dimers. Below 20 K however long-lived dimerization dominates, and only direct laser excitation of dimers can induce isotope separation.

Schematic of CRISLA chamber with free jet and skimmer.

The kinetics of transient QF6*:G dimer bondings and predissociations with VT conversion was investigated experimentally by VandenBergh (1985). This mechanism has a much higher VT conversion probability at low temperatures and pressures than direct specular VT conversions, and provides the basis for MLIS processes employing dimer harvesting, named SILARC (Separation of Isotopes by Laser Assisted Repression of Condensation) or CRISLA (Condensation Repression by Isotope Selective Laser Activation) (Eerkens, 1998). Earlier experiments in the 1970s through 1980s had shown pre-dissociation lifetimes of laser-excited dimers to be τpd < 10−6 s for internal ν3 infrared vibrations in QF6*:G or QF6*:QF6 dimers. Measured dimer absorption bands of these internal vibrations showed center frequencies that were slightly red- (and blue-) shifted from the monomer ν3 frequency. With broadened widths inversely proportional to predissociation lifetimes, one deduced τpd < 10−6 s (Smalley et al., 1976; Beswick & Jortner, 1978; Bernstein & Kolb, 1979; Geraedts et al., 1981, 1982; Snells & Reuss, 1987; Liedenbaum et al., 1989; VanBladel & VanderAvoird, 1990; Beu & Takeuchi, 1995, Beu et al., 1997; Beu et al., 1997; McCafferey & Marsh, 2002).

Although most earlier measurements were made with laser excitation of initially unexcited dimers, if an already ν3-excited QF6* monomer forms a QF6*:G or QF6*:QF6 dimer, the latter should dissociate at essentially the same rate as a direct-dimer-excited species would. Internally excited QF63)* molecules move at the same translational speeds as unexcited QF6 and should dimerize at the same pace. Because of the large frequency difference of internal ν3 vibrations and dimer-bond vibrations να, there is initially no coupling between these vibrations during dimer formation. After dimer bonding and a few dimer rotations and bond vibrations, predissociation with VT conversion of ν3 ensues in a Lissajous excursion. The small frequency defect Δν3 between internal ν3 vibrations of monomers and dimers to be accommodated during dimer formation, may aid the predissociation process. Dimerization, previously thought to require three-body collisions (Hirschfelder et al., 1967), actually proceeds much faster by two-body collisions (Eerkens, 2001a). This coupled with the discovery of microsecond lifetimes for excited QF6*:G dimers, correctly explains and predicts CRISLA isotope separations.

VandenBergh (1985) observed isotope-selective effects from predissociation and VT conversion of SF63)*:Ar, after SF6 monomers were excited and subsequently formed SF6*:Ar dimers. He also investigated and compared such effects for directly laser-excited SF6:Ar dimers. Both excitation schemes gave similar results, except that monomer excitations gave higher yields, attributable to the fact that photon absorption resonances for monomers are sharper and stronger. Since τpd < 10−6 sec and dimer formation times τdf and jet transit times ttr exceed 10−6 sec, the time τdVT to produce a VT conversion after a QF63)*:G dimer forms, is essentially instant compared to other microscopic processes in the jet; that is τdVT = τdf + τpd ≈ τdf.

The pathways by which an internal vibration in excited gaseous QF6* can relax are:

1. Spontaneous (Einstein) Photon Emission (lifetime

)

2. Direct collisional VT conversion (lifetime τVT = kVT−1)

3. Direct collisional (near-)resonant VV transfer (lifetime τVV = kVV−1)

4. Transient excited dimer formation followed rapidly by predissociation (lifetime τdf = kdf−1)

In MLIS one needs to know how fast mechanisms 1 through 4 are, compared to:

5. Dimer dissociation lifetimes of non-excited dimers (τdd)

6. Transit time of the jet through the laser irradiation cell (ttr)

The transit time by a jet traversing a laser beam equals:

where L is the distance traveled by the gas through the laser irradiation region and U is the average flow velocity of the jet in the absorption cell. Typically, 2 < L < 20 cm, and 0.1 < U < 2 km/s, so 10−5 < ttr < 2 × 10−3 s. In the gas phase, only options (1) through (4) are available to excited QF6* molecules for V-relaxation. For V-excited monomers that adsorb or condense on a surface, mechanisms (1)–(4) still apply with some modification, but in addition one must include interactions with phonons in the condensate which populate the latter's internal energy reservoirs. This situation (cold surface condensation) is examined in a separate paper (Eerkens, 2003). The analyses provided in that paper and in the present one, supersede and correct the theory given in Section 2 of (Eerkens, 1998).

We first examine probabilities and rates of processes (1) through (5), and then obtain enrichment parameters for the lightest and heaviest hexafluoride, SF6 and UF6. These molecules possess desirable isotopes 33S (0.74% ab.) used to generate 33P for nuclear medicine, and 235U (0.71% ab.) used as fuel for nuclear reactors. 33SF6 can be excited by the 10P-26 and 10P-28 lines of the CO2 laser, while 235UF6 is excitable with 16-micron rapidly pulsed Raman-converted CO2 laser photons or by 5-micron CO laser photons (VandenBergh, 1985, Eerkens, 1998).

2. INTERACTIONS OF QF6 WITH PHOTONS

2.1. Laser excitation of QF6

The laser excitation rate kA for iQF6 molecules to absorb resonant laser photons equals:

where σA is the laser photon absorption cross-section (cm2). QF6 absorption bands are hot-band-broadened, their peaks becoming sharper and higher towards lower temperatures. The laser flux φL is given by:

Here IL is the laser beam intensity in Watts/cm2 and εL = hνL the laser photon energy in cm−1. Typical vibrational absorption cross-sections have (temperature-dependent) values on the order of σA ≈ 10−18 cm2 at T < −150 K near a fundamental band peak (e.g., the ν3 vibration of QF6), σA ≈ 10−20 cm2 for binary combination bands (e.g., the ν23 vibration of QF6), and σA ≈ 10−22 cm2 for tertiary QF6(3ν3) absorptions. The laser photon energy εL must of course be equal to the vibrational excitation quantum εa (or εa + εb, 3εa) to effect resonant absorptions. For the ν3 vibration of 33SF6 one has εL ≈ ε3 ≈ 939 cm−1 and σA3) ≈ 2×10−18 cm2, while for 235UF6 it is εL ≈ ε3 ≈ 628.3 cm−1 and σA3) ≈ 10−18 cm2. Then Equation (2) yields:

With IL = 1000 W/cm2 for example, kA(SF63)) = 1.06 × 105 s−1 and kA(UF63)) = 8.02 × 104 s−1.

2.2. Spontaneous emission by QF6 (process 1)

The spontaneous photon emission or “Einstein” rate

for a vibrational v3 = 1 → 0 transition of the photon-active ν3 vibration of QF6 can be expressed by (Eerkens, 1973):

Here

is the dipole-derivative transition matrix for v3 = 0 [harr ] 1 transitions, given by:

The derivative dipole charge zb ≡ (zb)01 or derivative dipole moment (μb′)01 = ezb, and the vibrational mass Mb of vibration mode b have been calculated for some polyatomic molecules (Kraus, 1967), while emission wavelengths λb are usually known from spectral measurements (λ3 ≈ 10.5 μm for SF6 and λ3 ≈ 15.9 μm for UF6). More often

is measured from which values for zb2/Mb can be derived. Then with Mb calculated or estimated, zb can be deduced.

QF6 molecules have two photon-active vibrations, ν3 and ν4. For SF6 and UF6, values of

, or τse[SF63)] = 0.042 s and τse[UF63)] = 0.083 s have been measured (Burak et al., 1970; Kim & Person, 1981). This yields z32/M3 = 0.215 for SF6 and z32/M3 = 0.249 for UF6. Using estimated values of M3(SF6) = 18.77 amu and M3(UF6) = 119.88 amu, one deduces that z3(SF6) = 2.01 and z3(UF6) = 5.46. Lifetimes of other QF6 molecules for v3 = 1 → v3 = 0 transitions are of the same magnitude. That is 0.03 < τse[QF63)] < 0.3 s generally. Thus with transit times ttr < 2 × 10−3 s, there is little chance for laser-excited QF63)* monomers to loose vibrational energy by photon emission during its journey in the free jet. If the ν3 vibration is excited in QF6:G or QF6:QF6 dimers, much faster predissociations prevail over relaxation by spontaneous emission; again few photons are released.

3. COLLISION-INDUCED REACTIONS OF EXCITED AND UNEXCITED QF6 MOLECULES

3.1. Collision rates, mean free paths; overview of reactive processes

The contact collision rate kc of QF6 molecules with molecules M (M = G or QF6) is:

where nM is molecular density (cm−3) of M, σcQ/M = π(roQ + roM)2 is the collision cross-section (Å2) for QF6 + M collisions, and uQ/M the relative molecular velocity (cm/s) given by:

In Eq (8), nM(cm−3) = 0.97 × 1019pM(Torr)/T(K) is assumed, and MQ/M(amu) = MQMM/(MQ + MM), where QF6 is shortened to Q in subscripts. roQ, roG are molecular contact radii of QF6 and G (Eerkens, 2001a). For SF6 or UF6 diluted in 0.01 torr Ar at T = 50 K for example, kc(SF6/Ar) = 7.94 × 105 s−1 and kc(UF6/Ar) = 9.88 × 105 s−1. Table 1 lists additional rates kc for other gases.

Molecular parameters for SF6(ν3) and UF6(ν3) gas-phase collisional VT relaxations

Molecular parameters for SF6(ν3) and UF6(ν3) gas-phase collisional VT relaxations

The mean collision-free path [ell ]c for a QF6 molecule diluted in carrier gas G at temperature T is:

where σcQ/G is the elastic collision cross-section for QF6 + G encounters and pG = (1 − yQ)ptot is carrier gas pressure. QF6+QF6 collisions are neglected since mole fraction yQ << 1 is assumed.

Most QF6 + M encounters result in elastic collisions, but some result in the formation of transient QF6:M dimers. For vibrationally excited QF6* molecules, aside from elastic collisions, encounters with M can result in direct VT conversions or in VV transfers if M = QF6 with probabilities

. Also a dimer QF6*:M can form with probability

, which (pre-)dissociates within microseconds (Smalley et al., 1976; Beswick & Jortner, 1978; Bernstein & Kolb, 1979; Geraedts et al., 1981, 1982; Snells & Reuss, 1987; Liedenbaum et al., 1989; VanBladel & VanderAvoird, 1990; McCafferey & Marsh, 2002). The outcome of both

processes is that epithermal QF6 and M molecules recoil off each other after conversion of V to T energy, as discussed below. Direct resonant VV transfer between an excited QF6* and unexcited QF6 (i.e., M = QF6) has a high probability but does not effect overall V storage. However while the net energy distribution is not changed by VV transfers, for MLIS processes, near-resonant exchanges iQF6* + jQF6iQF6 + jQF6* generate undesirable isotopic losses (isotope scrambling). Such losses are minimized by a high dilution of QF6 in carrier G.

3.2. Direct specular VT relaxations (process 2)

In a gas mixture of QF6 diluted in carrier gas G, VT deexcitations of QF6* through contact collisions QF6* + G → QF6 + G + K.E. occur at rates:

Here kc is given by (8) and VT transition probability

is (Eerkens, 2001a):

with ε3 = hν3 and we assume ζ32FA ∼ 0.25. The collision function FC is given by:

with:

In (14), LR is the repulsive range parameter in Å, Δε = ε3 equals the energy quantum (in cm−1) that is VT-converted, MQ/G is the reduced mass of the collision partners in amu, and T is the gas temperature in K. For χ > ∼20, Eq. (13) reduces to the simpler SSH relation (Eerkens, 2001a; Schwartz et al., 1952; Schwartz & Herzfeld, 1954; Yardley, 1980):

VT deexcitations can also take place in collisions of QF6* with a like QF6 molecule, with MG replaced by MQ in the above expressions. However in QF6* + QF6 encounters, resonant VV transfers (process 3) have a much higher probability of occurring than the VT process 2.

Using molecular parameters listed in Table 1, calculated VT probabilities at different T for SF63)/G and UF63)/G with G = H2, He, N2, Ar, Xe, SF6, and UF6 are listed in Table 1. They show the enormous effect of mass MG and T on direct VT probabilities and rates.

3.3. Direct specular resonant VV transfers (process 3)

For a direct VV transfer of v3 in a QF63)/QF6 contact collision, the probability is (Eerkens, 2001a):

where:

and:

Here νb, xb, Mb, ζb2, are the fundamental frequency, anharmonic constant, mass constant, and steric collision factor for internal vibration b in molecule QF6 (in the present case b = 3), while in (17), we assumed ζ32FA ∼ 0.25 again, and we set FC = 1 since Δε = 0 for a VV transfer. For a two-quantum change in a QF6QF6 quantum box with v3 = 1 → 0 → 1, one can as an approximation set m − n = Σ|Δv3| = 2, with effective values m = 2 and n = 0 in Ωanh used in Eq. (18). Note also that MQ/Q = 0.5 MQ. The VV rate in the gas mix is then:

with:

Here yQ is the mole fraction of QF6 in the QF6/G gas mix, and kc is the QF6/G collision rate given by (8). The factor cQ/Q corrects for different collision cross-sections and masses of QF6/QF6 interactions. Table 4 of (Eerkens, 2001a) lists some values of collision radii roQ and roG needed in (20) for molecules QF6 and G, while the reduced mass MQ/M = 2MQMM/(MQ+MM).

Using values for M3, ε3, x3, σQ/Q, MQ/Q from Table 1, one finds from (16) for example at T = 50 K, that

. For 2% SF6 or UF6 diluted in 10−2 torr of Ar, one has from (19) that τVV(SF63)/SF6) = 8.2 × 10−3 s, and τVV(UF63)/UF6) = 2.36 × 10−3 s. These times are longer or comparable to ttr and τdf (< 10−3 s) at T < 50 K.

3.4. Dimer formation rates (process 4)

It was shown in (Eerkens, 2001a) that the dimer formation probability

can be expressed by:

where:

Here vd is the highest (dissociation) vibrational quantum level in the dimer potential well, while MQ, MM, yQ, yM are molecular masses and mole fractions of QF6 and M (M = G or Q ≡ QF6). Energy εα = hνα is the fundamental energy (frequency) quantum and Dα is the well-depth of the dimer bond. We use V′ to designate dimer-bond vibration as opposed to V (without a prime) for internal high-energy molecular vibrations. The dimer formation rate kdf is:

with kc given by (8) and

by (21).

Dα values for various collision partners can be obtained from Table 4 in (Eerkens, 2001a), assuming (Dα)Q/M = (DαQDαM)1/2, while να can be calculated from Eq. (19) in (Eerkens, 2001a), with values of ro = ½(roQ + roM) taken from the same table. For 2% SF6 or 2% UF6 diluted in 0.01 torr of H2 for example, one finds

at T = 50 K. With Ar instead of H2, one obtains

at T = 50 K. In 0.01 torr of carrier gas, the dimer formation rates are then kdf(SF6/H2) = 1.67 × 104 s−1 and kdf(SF6/Ar) = 501 s−1, while kdf(UF6/H2) = 1.59 × 104 s−1 and kdf(UF6/Ar) = 579 s−1. Other calculated values of

for selected gas mixtures at various temperatures are listed in Table 2. The table illustrates the diversity of dimerization rates due to the strong effect of mass MG and temperature T in the dimer formation process.

Dimerization parameters for SF6/G and UF6/G encounters

Figures 2 and 3 show plots of

versus gas temperature T for different SF6/G and UF6/G gas mixtures. The figures clearly show that towards lower temperatures, dimer formation/predissociation VT conversions are more probable than direct collisional VT relaxations. In some early (1970s) measurements of VT rates, researchers observed that vibrational relaxation rates, after initially dropping as expected, increased again towards lower temperatures (Yardley 1980). Except for light elements, this was in conflict with the collisional VT theory of Schwartz et al. (1952) and Schwartz & Herzfeld (1954), and often ascribed to possible VR (Vibration → Rotation) transfers, even though collisional VR conversions are rather improbable quantum-mechanically. By including low-temperature two-body dimer formation and pre-dissociation of V-excited species, agreement between theory and experiment is restored.

VT conversion probabilities for SF6*(ν3)/G gas mixtures.

VT conversion probabilities for UF6*(ν3)/G gas mixtures.

3.5. Collisional dissociation rates of dimers (process 5)

At monomer partial pressures pQ = yQptot with pQ less than the homogeneous cluster nucleation pressure pd (see Section 7.2), the dissociation rate of dimers is (Eerkens, 2001a):

with:

Here

is the probability that a dimer QF6:G dissociates to QF6 + G in a collision with molecule G. The parameter ad corrects the collision rate kc for QF6/G monomer collisions to the rate kcd = adkc for QF6:G/G dimer collisions which has a larger collision cross-section and higher reduced mass. Contact radii roQ, roG, and reduced masses MQ/M in (33) were discussed in Sections 3.3 and 3.4. Calculated values of

are listed in Table 2. At T ∼ 50 K and pG ∼ 0.01 torr for example, Tables 1 and 2 show that 104 < kdd < 105 s−1, or 10−5 < τdd < 10−4 s. This compares with transit times of 10−5 < ttr < 2 × 10−3 s.

4. MOLECULAR MIGRATIONS AND THERMALIZATION OF EPITHERMALS

4.1. Isotope harvesting from laser-irradiated free jets

In isotope separation techniques that utilize laser irradiated free jets of supersonically cooled gas mixtures, advantage is taken of the fact that QF6:G dimers, QF6(*) monomers, and QF6! epithermals, migrate at different rates due to different collision cross-sections and average molecular speeds. After isotope-selective excitation of iQF6, the iQF6* will dimerize briefly as iQF6*:G which rapidly dissociates with VT conversion, creating above-thermal or “epi-thermal” iQF6!. We label epithermal iQF6! with superscript ! to distinguish them from internal vibrationally excited iQF6* molecules marked by superscript * which move at thermal speeds. At low temperatures, under continuous laser pumping, the iQF6 are distributed over iQF6*, iQF6!, iQF6:G, and non-excited iQF6 populations, while jQF6 populate jQF6:G dimer and jQF6 monomer groups. Because of the different population fractions and different migration rates, isotopomers iQF6 flee the jet core at higher rates than the jQF6, resulting in isotope separation. To summarize, the jet contains iQF6 of four types, iQF6*, iQF6!, iQF6, and iQF6:G, which have three different average transport constants, while the jQF6 are divided over two classes, thermal monomers jQF6 and slower moving dimers jQF6:G, each with different transport parameters (see below).

The peripheral or “rim” gases are pumped out separately from the supersonic jet core gases which exit through a skimmer at the end of the jet chamber (see Fig. 1). As they migrate through the jet core while in transit to the skimmer, laser-excited iQF6 cycle repeatedly through excitation → dimerization → dissociation → VT-conversion/epithermalization → re-thermalization. If iQF6:G dimers are laser-excited instead of iQF6, the sequence is the same except it starts with dimerization → excitation → dissociation…. Non-excited jQF6 isotopomers experience only collisional dimerizations and dissociations, but at low temperatures most travel as jQF6:G dimers which migrate more slowly than iQF6 monomers. Thus iQF6 escape at higher rates from the jet core than jQF6, thereby enriching the rim gases with iQF6.

The mean free path between collisions of a QF6 molecule diluted in gas G was given by Eq. (10). Table 3 lists pG[ell ]c values for several gases G at different temperatures. For example, at pG = 0.01 torr, T = 50 K, and G = Ar, a SF6 molecule experiences one collision in every 0.22 mm of travel. Thus although “thin,” the gas in a jet with 10 mm radius, still follows Maxwell-Boltzmann kinetics and obeys the diffusion laws, as assumed in what follows.

Thermalization parameters for epithermal SF6! and UF6! molecules

4.2. Radial molecular migrations and jet core escapes

To estimate the radial outward diffusion of QF6 molecules and QF6:G dimers from a supersonic free jet, we approximate the actual jet contour by an equivalent cylinder with radius R equal to the skimmer entrance radius (see Section 7). All molecules in the cylinder striking the “wall” at R are then assumed to leave the cylinder, as if the wall were a total molecular absorber. The non-excited and vibrationally excited monomers will be shown to escape the jet at thermal rates kW, while epithermals escape at rates kW1 and dimers leave the jet with a rate constant kWd, such that kW1 > kW > kWd s−1 per molecule.

In (Eerkens, 2001b) it was shown that a batch of N QF6 molecules diffusing through carrier gas G in a long cylinder with volume πR2L and surface 2πRL, strike the wall at rates:

if the wall is 100% absorbing. Here [ell ]c is the mean-free-path and R is cylinder radius in cm, σcQ/G is the Q/G collision cross-section in Å2, pG is in torr, T is in K, and MQ/G = MQMG/(MQ + MG) is reduced mass in amu. These units will be used in all final expressions that follow. Table 1 lists typical values of kWpG. The inverse rate kW−1 equals the average time τW for a QF6 molecule in the cylinder to reach the wall, provided it is not terminated in the gas phase.

In our mathematical model, the “wall absorption” rate equals the jet core escape rate of QF6 molecules. Because the jet moves at supersonic speed, molecules who cross the “wall” at radius R, can not return to the jet core and can be considered “absorbed” by the “peripheral” or “background” gases in the jet chamber. For a batch of QF6 molecules entering the jet chamber, which looses QF6 to the “wall” by radial diffusion as it moves towards the skimmer, the number left in the jet core as a function of time is:

Here NQo is the tagged QF6 population entering the jet chamber at t = 0. After travel time t = ttr, the number of QF6 molecules remaining in the jet core is:

and the number NQesc or fraction Θ that escaped equals:

Here the transit time t = ttr = L/U, where L is the travel distance and U the average bulk velocity during the jet's travel through the irradiation chamber.

For a batch of outward diffusing epithermal molecules QF6!, the migration rate is similar to thermals, except transport parameters must be adjusted and thermalization must be considered. Using subscript 1 to label epithermal transport parameters, the epithermal migration rate kW1 is:

Here we used (34) and replaced T3/2 by T11/2T since the epithermal velocity u1 is proportional to T11/2. The mean free path [ell ]c due to scattering of QF6! monomers by molecules G is virtually the same as for thermals and is proportional to T/pG, if QF6 is highly diluted in the QF6/G gas mix.

The average epithermal temperature 〈T1〉 during slow-down can be approximated by the logarithmic mean between the initial kinetic energy ε0 and the final one εT, that is:

with:

Here a is defined as:

in which εQ is the original kinetic recoil energy given to QF6 in the VT conversion by dimer predissociation, εa is the released quantum of vibrational energy, and εT = kT is the thermal kinetic energy. From momentum and energy conservation of dissociating dimer partners one deduces that:

The original kinetic energy ε0 of a VT-energized QF6! molecule then equals:

In the case of laser excitation of ν3 in QF6 molecules, the quantum energy εa = ε3 = hν3.

Note from (42) that the heavier the dimer partner G is, the larger the fraction of the vibrational energy εa imparted to a recoiling QF6 molecule. Most useful carrier gases have masses MG that are less than those of SF6 (MQ = 146 amu) and UF6 (MQ = 352 amu). Thus for QF6 molecules one usually has ηQ/G ≤ 0.5. Heavy particles or solid surfaces might look attractive for maximizing εQ, but one finds that other problems are introduced. Laser isotope separations utilizing cold surfaces are discussed in a separate paper (Eerkens, 2005).

Returning to (38), the “wall” escape rate for epithermals is:

and the epithermal escape fraction Θ1 is:

For dimers which move thermally, the collision cross-section and reduced mass is higher than for monomers. That is in (34), one must replace σcQ/G and MQ/G by σcQG/G and MQG/G. Using subscripts d for dimers, the escape rate kWd for dimers then equals:

Here ψd, which is less than unity, is the migration rate correction factor for dimers compared to monomers. The QF6/G collision rate kc and mean-free-path [ell ]c were given by (8) and (10). The dimer escape fraction is finally:

4.3. Slowing-down parameters for epithermal molecules

The average number of collisions that an epithermal molecule experiences till it is thermalized, and the average distance it moves, can be obtained from “Fermi Age Theory” originally developed for epithermal neutrons instead of molecules (Glasstone & Edlund, 1952). This theory shows that the average number of collisions NT to thermalize an epithermal molecule equals:

Here εI, εII are kinetic energies of an epithermal molecule before and after a momentum-transfer collision, and the average logarithmic energy decrement ξ equals:

with:

Kinetic energy εQ was given by (42), and the ratio a = εQT by Eq. (41). In Fermi age theory, an epithermal neutron or molecule is considered thermalized when its kinetic energy has dropped to εT = kT, rather than (3/2)kT (Glasstone & Edlund, 1952). The average fractional energy loss for each epithermal QF6! + G momentum transfer collision is given by:

If energy losses of QF6! during thermalization are purely due to scattering and momentum transfers, the slowing-down length or migration distance “as the crow flies,” is (Glasstone & Edlund, 1952):

where [ell ]c and NT were given by (10) and (49), and the thermalization “age” ϑT equals:

Here we assumed D[ell ]c/ξ to be independent of energy and the “flux diffusion coefficient” is approximated by D ≈ 1/3[ell ]c. Table 3 lists ξ, εQ, NT, [ell ]c, and sth of SF6 and UF6, with various carrier gases at several temperatures T and kinetic energies ε0.

The thermalization rate of epithermals in the gas (= a feed term for the thermal group), is:

where units are in torr, Å2, K, and amu, as before. This rate is high and forces epithermal molecules to re-enter the thermal monomer group repeatedly and be re-excited, etc, several times before reaching the wall. The same is true for dimers which are continuously formed and dissociated in collisions. That is, in a unit volume of gas, rapid exchanges take place between populations of the four migrating groups, but under steady-state conditions, the fractional populations are constant (see Section 5). Individual molecules are continuously cycled through laser-excitation, dimerization/predissociation/epithermalization, and re-thermalization in tens of microseconds as they migrate to the wall. However as long as the population fractions for each migrant class are constant, the effective escape rate per molecule per second is kW for the thermal group, kW1 for the epithermal fraction, and kWd for those in the dimer state, even though individual molecules are continuously exchanged between these different migrant groups. That is, under continuous laser pumping, all thermalizing iQF6 leaving the epithermal group are replaced by newly excited iQF6* that dimerize, and epithermalize. Thus as long as molecules are not permanently removed in the gas and populations are steady, the migration of iQF6 can be considered taking place by three independently moving species: a fraction fit at rate kW for excited or unexcited thermal monomers; a fraction fi1 for epithermals at rate kW1; and a fraction fid at rate kWd for dimers. Steady-state fractional populations fit = fi* + fim, fi*, fi!, fim, fid of excited, epithermal, and non-excited iQF6 monomers and dimers are considered next.

5. POPULATION OF DIMERS, LASER-EXCITED MOLECULES, AND EPITHERMALS

5.1. Monomer laser excitation

As mentioned, under steady monomer laser-irradiation, the iQF6 population is distributed over four distinct groups: excited species (*), epithermals (!), non-excited monomer molecules (m), and non-excited dimers (d). The fraction fi* of laser-excited iQF6* is defined as fi* = ni*/ni = [iQF6*]/[iQF6], where the brackets (by convention) indicate concentrations. Here ninQi is the density (molecules/cm3) of a selected isotopomer iQF6, nQ = ΣjnQj is the total density of all QF6, while the isotopic abundance of iQF6 equals xi = ni/nQ. Other populations of interest are the fraction fi! of epithermal iQF6! molecules, the fraction fim of non-excited iQF6 monomers, and fraction fid of non-excited dimers. We shall assume that most dimers formed in the gas mixture will be heterodimers QF6:G if mole fraction yQ < ∼0.05. That is, we neglect homo-dimers QF6:QF6 and trimers, tetramers, etc. Neglect of oligomers other than dimers is reasonable because of the short residence time of the jet in the chamber (Sec. 7.2). A material balance then gives:

Neglecting excitations of in-flight epithermals and second-step excitations of once-excited species, the excited species production rate under laser irradiation is given by the product φLσAfimni = kAfimni, where φL is the laser flux, σA is the photon absorption cross-section, and fimni is the non-excited monomer population of iQF6. Then, since fim = 1 − fi! − fi*fid, one has:

The first term on the right-hand-side of the kinetics balance (57) is the laser excitation rate per unit volume of iQF6 monomers which produces iQF6*, while the second term represents losses of vibrational excited iQF6* due to dimerizations/predissociations and direct VT and VV transfers, spontaneous emissions, and “wall” escapes of iQF6*. To-the-wall survival probabilities e* for QF6* in the gas phase are included in (57) to balance all sources and sinks:

Here the average travel time to the wall τW = kW−1. Unless the gas is very thin, e* ∼ 0 usually. Parameters

were already reviewed, while kA was discussed in Section 2.1.

The loss rate kV = kVV + kVT from direct VT and VV transfers of the ν3 vibration in SF6* and UF6* equals:

Here the collisional encounter rate kc was given by Eq. (8),

by (12),

by (16), and cQ/Q by (20). At enrichment-favoring temperatures T ∼ 30 K, and with pG ∼ 0.01 torr, yQ = 0.02, and masses MG > ∼10 amu, Table 1 shows that

, while Table 2 gives

. Both

decrease with decreasing T, but

increases with decreasing T. Under these conditions, with kc ∼ 106 s−1, one obtains kdf ∼ 103 s−1, kVT3 ∼ 10−4 s−1, kVV3 ∼ 102 s−1, while

. Thus (57) and (58) can be simplified to:

and:

The epithermal population fi!ni is balanced by the relation:

Here kdf was given by (30), kth by (55), and kW1 by (44). Since for T < 50 K, kVT << kdf, and usually e* << 1, e1 << 1, the approximation in (62) applies under enrichment-favoring conditions. The to-the-wall survival probability e1 for gas-phase epithermals in (62) equals:

Here μ1kc cancels out in the ratio kth/kW1 in the exponential of (63). Factors e1 and e* apply to “leaks” from internal microscopic processes per unit volume in the jet, insuring correct values for fi* when R < [ell ]c. Survival factors e* and e1 must not be confused with bulk losses Θ or Θ1.

The population of non-excited dimers fid is determined finally by the kinetic balance:

Under steady-state laser-irradiation and gas flow, d(fi*ni)/dt = d(fi!ni)/dt = d(fidni)/dt = 0. Then ignoring kVT since kVT << kdf, Eqs (62) and (64) yield:

with:

and:

with:

Finally, setting (57) equal to 0 and inserting (65) and (67) yields:

The dimer formation and dissociation rates kdf and kdd were given by Eqs (30) and (31), together with (21) through (27) and (32), while kth was expressed by (55) and kA by Eq (2).

Concentration fractions fi! and fid given by (70) and (71) are key factors in isotope separation. The higher fi!, the higher the enrichment factor β is. Note that at low temperatures with high levels of dimerization, one has kdd << kdf and kdd → 0. Then the third terms in the denominators of (69) and (70) approach infinity and both fi! → 0 and fi* → 0, while fid → 1 according to (71). Finally note that unless kdd → 0, in the limit that kA → ∞, the population fi! becomes:

The ratio

is clearly important, and the higher

is, the higher fi! and β.

A second source term +fj*(1 − xi)(ni/xi)kVV could have been added to the right-hand-side of (57) to represent VV re-transfers from other (non-i) excited jQF6* isotopomers which were produced by collisional VV transfers iQF6* + jQF6iQF6 + jQF6*. Here (1 − xi) = xj is the fraction of other isotopomers jQF6 (j ≠ i), ni/xi = nQ, kVV is the VV transfer rate, and fj* is the steady-state fraction of isotopomers jQF6 that are excited due to VV transfers from iQF6*. A second balance equation would have to be written for fj*, but if yQ = [QF6]/[G+QF6] < ∼0.05, the “VV recycle” term gives only a small correction and we shall (conservatively) neglect it.

We tacitly neglected laser excitation of epithermals or higher excitations of iQF6* molecules. If the kinetic energy εQ is high enough, it might be possible for an excited epithermal iQF6*! to relax its ν3 vibrational energy by direct VT conversion collisions, since

increases with T. However for ν3 vibrations of QF6 one finds εQ < 1000 K, and the probability

is still negligible unless MQ/G < ∼4 amu (i.e., for G = H2 or He). “In-flight” excited epithermals iQF6*! must first slow down before they can dimerize and experience predissociation with VT conversion again. After thermalization, still-excited iQF6* become part of the thermal population group fi*, whose kinetics balance was already considered. Double and higher step excitations iQF6* → iQF6** → iQF6*** can also occur, since absorption frequency differences for higher vibrational steps are usually small. However subsequent dimerization/VT conversion proceeds preponderantly in one-quantum steps, so we can lump fi** and fi*** populations with the fi* crowd.

5.2. Dimer laser excitation

If iQF6:G dimers are laser-excited instead of iQF6 monomers, one can write:

Here fid is the fraction of iQF6 that are dimerized as iQF6:G and kdd is the dimer dissociation rate given by (31). The laser rate kA applies here to resonant absorptions by iQF6:G dimers whose internal ν3 vibrations are slightly shifted compared to the monomer ν3 vibration of iQF6.

Under steady-state conditions, setting time derivatives equal to 0, one obtains:

with:

Usually kA >> kdd, kth >> e1kW1, and e1 << 1 if pG ∼ 10−3 torr, in which case the approximations in (76) and (78) apply. Combining (76) and (77) yields:

and:

Here we added subscripts d to fi! and fid to distinguish dimer laser-excitation fractions fi!d and fidd from expressions for monomer laser-excitation fractions fi! and fid given in Section 5.1.

Comparing (79) with (70), we see that the expressions for fi!d and fi! are almost the same except for the last term in the bracket. For dimer-excitation the last term kdd/kdf → 0 if kdd → 0, and:

On the other hand, for monomer pumping we found fi! → 0 when kdd → 0 since the last term kdf/kdd → ∞. Thus at very low temperatures where kdd → 0, only laser excitation of iQF6 dimers can promote isotope separation. However because monomers have higher and sharper peak absorption cross-sections (σA) than dimers, it is more advantageous to pump monomers instead of dimers, as long as kdd is not entirely 0. That is, with the same laser intensity or flux φL, one usually finds (kA)monomer ∼ 10(kA)dimer. Temperatures and pressures for optimum enrichments using either monomer or dimer laser pumping will be analyzed in the next section.

6. LASER-INDUCED ISOTOPE SEPARATIONS IN SKIMMER-SCOOPED JETS

6.1. Enrichment factor βm for monomer laser excitations

Assuming continuous laser irradiation of the jet and steady jet flow, the mole fractions of thermals (iQF6 + iQF6*), epithermals (iQF6!), and thermal dimers iQF6:G throughout the jet will be fitxiyQ = (1 − fi!fid)xiyQ, fi!xiyQ, and fidxiyQ, where yQ is the mole fraction of QF6 in the gas mix and xi = [iQF6]/[Σj(jQF6)] is the isotopic abundance of iQF6. Here the fraction of excited and non-excited thermal monomers is fit = fi* + fim = 1 − fi!fid. Fractions fit, fi!, fid, fjt, and fjd are constant but the number of iQF6 and jQF6 in the jet drops as the jet travels through the chamber to the skimmer due to escape fractions Θ, Θ1, and Θd.

Using a prime to label the isotopic abundance of iQF6 in core-escaped gas and no prime for incoming un-irradiated feed gas, the isotopic composition xi′ of QF6 in the escaped gas is:

or:

Here βm = xi′/xi is the enrichment factor for iQF6 in the jet-escaped rim gases, and φ1, φd equal:

Parameters μ1, ψd, κ1, ωd, and kW were given by (40), (47), (66), (68), and (34), while mole fractions fi*, fi!, and fid were given by (69), (70), and (71). Mole fractions fjm and fjd of other isotopic monomers jQF6 and dimers jQF6:G are:

with escape fractions Θ and Θd. Also of course xj = 1 − xi. Note that fjd given by (87) agrees with fid expressed by (71) for kA = 0. Also note if xi << 1, the enrichment factor becomes:

with:

The enrichment relation (83) or (88) has three important components, fi*, φ1, and cd. The laser-pumped excited fraction fi* increases with laser radiation intensity and the higher fi* is, the higher βm. However fi* also depends on the availability of monomers and, as mentioned, if kdd → 0 or ωd → 1, one finds fi* → 0. This happens at low temperatures when all monomers become dimerized during the jet flight and there are no monomers left for laser excitation. Since the factor fi* can never exceed 1, only the epithermal factor φ1 and dimer factor cd can make βm larger than 2. The larger φ1 and cd are, the higher βm is.

Figures 4 and 5 show plots of βm at different pressures and temperatures for several SF6/G and UF6/G gas mixtures, assuming a fixed kA = 105 s−1 and jet core with radius R = 1 cm and length L = 20 cm. The curves show maximum βm's around T = 25 K for SF6/G and T = 35 K for UF6/G at pressures between pG = 0.005 and 0.05 torr. High values of β are driven by high values for cd and φ1. The latter has high values when εQT = ηQ/Gεa/(kT) is high. Since εa is fixed for a given QF6, one expects low values of T and a high value for ηQ/G = MG/(MG+MQ),(i.e., high MG), to yield a high φ1 and β. Indeed Figures 4 and 5 show heavy SiBr4 (MG = 348) as carrier to give the highest enrichment, but its poor aerodynamics gives problems (see Sections 7 and 8). Low temperatures for good enrichments can only be attained temporarily in supersonic jets. Feasibility of attaining such low temperatures will be investigated in Section 7.

Enrichment factors βm(T) for SF6(ν3) monomer excitation at fixed pG.

Enrichment factors βm(pG) for SF6(ν3) monomer excitation at fixed T.

Enrichment factors βm(T) for UF6(ν3) monomer excitation at fixed pG.

Enrichment factors βm(pG) for UF6(ν3) monomer excitation at fixed T.

Inspection of (83) and (88) shows that the observed peaks of βm around T = Tmax ∼ 25K for SF6/G and T = Tmax ∼ 35 K for UF6/G, occur just before ωd → 1 or ωd ≈ 0.995, where:

For T much less than Tmax, almost total dimerization occurs as

. Under these strong dimerization conditions, there are no monomers left for laser excitation during the supersonic jet flight in the chamber, and thus β → 1. One finds that epithermals contribute very little to βmax, and the major isotopic effect is due to dimers.

6.2. Enrichment factor βd with dimer laser excitation

Isotope enrichment factors under direct selective laser excitation of dimers iQF6:G can be deduced in the same manner using the fractions fidd and fi!d obtained in Section 5.2. One has:

where λ1 (≈ kA/kth), φ1, φd, and cd = ωd(1 − φd) were given, respectively, by (78), (84), (85), and (89) with ωd = kdf/(kdf + kdd) defined by (68), and where fidd = [1 + kA/kth + kA/kdf + kdd/kdf]−1 was given by (80). For xi << 1, the expression for βd reduces to:

For low temperatures where kdd << kdf and ωd → 1, one has that cd → (1 − φd) and (92) becomes:

with:

Figures 6 and 7 show enrichment factors βd(T) for 33SF6/G and 235UF6/G for carriers Ar and Xe at various temperatures and pressures, assuming R = 0.4 cm and a laser excitation rate of kA = 104 s−1 (simulating (σA)dimer ∼ 0.1 (σA)monomer). The plots show peaks of βd in the same general region (20–40 K) where βm peaks, but in addition there are some high βd values at very low pressures (pG ∼ 0.001 torr) and temperatures (T < ∼ 0.5 K), where complete dimerization has set in. This region is unattractive however since such low jet expansion temperatures and pressures are difficult (costly) to achieve, and product cuts Θo are very low there (see below).

Enrichment factors βd(pG,T) for SF6:G dimer laser excitations.

Enrichment factors βd(pG,T) for UF6:G dimer laser excitations.

6.3. Overall product cuts Θo for monomer and dimer excitations and depletion factor γ

The enrichment factor β alone is insufficient to characterize the performance of an isotope separation process and a minimum of two performance parameters are needed. Besides β, the product “cut” Θ or tails isotope depletion factor γ must be specified to define a separation process (Benedict et al., 1981). Usually the product output cut Θo is specified which is defined by:

From a mass balance, one can show that the isotope depletion factor γ in the “Tails” stream (this γ not to be confused with γ = cp/cv) is related to β and Θo by (Benedict et al., 1981):

Here a single prime refers to product, a double prime to tails, and no prime to the feed stream.

In the case of monomer excitations, the product cut Θom according to (95) equals:

which reduces to:

where Θ = 1 − exp(−kWttr), Θ1 = 1 − exp(−μ1kWttr), Θd = 1 − exp(−ψdkWttr), φ1 = Θ1/Θ, φd = Θd/Θ (see Eqs (37), (45), (48), (84), (85)). The parameter cd = ωd(1 − φd) was previously defined by (89), ωd = kdf/(kdf + kdd) by (68), while fractions fit = fim + fi* = 1 − fi!fid, fi! = κ1fi*, and fid = ωd{1 − (κ1 + 1)fi*}, with fi* given by (69) and κ1 by (66). We also used the relations xj = 1 − xi, fjd = ωd, and fjm = 1 − fjd = 1 − ωd.

For the dimer excitation case, one deduces similarly:

where λ1kA/kth was given by (80) and fidd by (82).

If xi << 1, both Θom and Θod reduce to:

with Θ given by (37) and ωd by (89). Figures 8 and 9 give calculated curves of Θo for SF6/G and UF6/G in the pressure and temperature ranges where βm and βd are high. For Θo > ∼0.3, the cylindrical jet approximation may become less realistic, and the equations for calculating βm, βd, and Θo are less reliable. Note that at temperatures below T = 1 K, where βd has some secondary peaks, the product cuts Θo are very small (Θo < ∼10−3), and economically unattractive.

Isotope product cuts Θ(T) for 33SF6 at fixed pG.

Isotope product cuts Θ(pG) for 33SF6 at fixed T.

Isotope product cuts Θ(T) for 235UF6 at fixed pG.

Isotope product cuts Θ(pG) for 235UF6 at fixed T.

For a usable laser isotope separation process, one would like Θo > ∼0.08 and β > ∼1.8 or γ < ∼0.93 according to (96). From Figs. 4 through 9, it is clear that cuts Θo are higher the higher T and p are, contrary to the enrichment factor β which is higher the lower T and p are. Thus compromise temperatures and pressures must be chosen to optimize separations. For example for SF6/Xe, one might select T = 25 K and pXe = 0.007 torr, which gives β = 2.00 and Θo = 0.10. For UF6/SF6, a possible selection might be T = 35 K, pSF6 = 0.02 torr, yielding β = 1.95, Θo = 0.23. These are reasonable process parameters, and allow 33SF6 to be enriched from 0.75% to 96% in seven stages for example, or 235UF6 from 0.73% to 2.8% in two, or from 0.73% to 5.4% in three stages. This is much better than in Diffusion (β ∼ 1.002) or Ultracentrifuge (β ∼ 1.1) separation plants that enrich 235UF6 employing hundreds of stages.

For the MLIS/CRISLA scheme, staging is simple since Feed, Product, and Tails streams are all gaseous and chemically the same except for isotope concentrations. In MLIS schemes other than CRISLA, the product is chemically changed and must be reformed between stages.

7. SUPERSONIC FLOW AND SUPERSATURATION CONSIDERATIONS

7.1. Free jet flow characteristics

To ascertain that the free-jet low temperatures and pressures desirable for useful isotope separations are attainable without excessive cluster growth, we briefly examine some supersonic expansion relations obtained from the adiabatic expansion relations (Liepmann & Puckett, 1953):

Here subscripts o and t on T, p, A refer to upstream reservoir (stagnation) and throat conditions of the gas mixture which is expanded from a large reservoir through a small constricting circular opening or a slit (the “throat”) with cross-sectional area At. No subscripts indicate downstream conditions in the supersonic jet stream. A is the cross-sectional area through which the jet core flows and U is the supersonic bulk-flow velocity of the jet at a downstream station. Uso is the sound velocity of the gas in the reservoir, given by (Liepmann & Puckett, 1953):

with M being the atomic mass of the gas atoms or molecules in amu. Gas coefficients γ = cp/cv equal γ = 1.67 for monatomics He, Ar, Xe, γ = 1.4 for diatomic gases (H2 and N2), γ ≈ 1.3 for SF6, CH4, and Br2, and γ ≈ 1.07 for UF6 and SiBr4.

Low-pressure background gases in the jet chamber create a diffuse boundary layer around an expanding supersonic free-jet core, as the latter experiences oscillating expansions and contractions that form a series of Mach cells with normal and oblique shocks. Figures 10 and 11 show free-jet structures measured by (Eerkens, 1957, Eerkens et al., 1958) and (Chow, 1959) for reservoir pressures of po = 0.8 and 3.9 torr, and throat diameters of Dt = 2.22 and 1.27 cm, with jets exhausting into vacuum chambers pumped down to less than 0.01 torr. The latter conditions are close to the expansions needed for successful CRISLA operations. The jet contours of Figures 10 and 11 also show that it is reasonable to approximate the jet core by a gas-filled cylinder with a “vacuum wall” as assumed in Section 4.2 to simplify mathematics. As long as radial diffusion losses from the jet are not excessive (Θo ≤ ∼0.3), jet expansion parameters using (101)–(104) provide useful first-order estimates. Note that because the flow is supersonic, the downstream skimmer (see Fig. 1), can not perturb the upstream flow pattern.

Measured free-jet contour and pressure distribution for G = Air; po = 3.86; p∞ = 0.21 torr (after Eerkens et al., 1958; Chow, 1959).

Measured free-jet contour and pressure distribution for G = Air; po = 0.80; p∞ = 0.15 torr. (after Eerkens et al., 1958; Chow, 1959).

Theoretical calculations of two-dimensional jets, though available (Adamson & Nichols, 1959), are complex and involve graphical constructions. While the one-dimensional flow relations (101)–(104) can not predict the observed two-dimensional “Mach diamonds or cells” of supersonic free jets, they do provide average order-of-magnitude downstream values of the supersonic flow expansion cross-sections (A/At) and accompanying pressures and temperatures. To attain gas expansions to T ∼ 25 K and p ∼ 0.01 torr with a noble gas carrier for example, one calculates reservoir pressures of 5 torr and throat diameters of 0.532 cm, which are quite manageable. Table 4 lists calculated values of the required reservoir tank pressures po and required circular throat radii Rt or rectangular slit widths Wt, for various carrier gases containing a few percent QF6, which are expanded in one step to pressure p = 0.01 torr and various temperatures in a vacuum chamber. The reservoir temperature is assumed to be To = 300 K in these one-step-expansion calculations, while a final jet radius of Rj = 1 cm is assumed in all cases. Calculations with SiBr4 as carrier gas G shows it to be impractical due to a very low γ. Low-energy vibrations (hνakT) in this molecule retain too much energy at low temperatures.

Calculated one-step supersonic jet expansion parameters for QF6/G with G = He, Ar, Xe (γ = 1.67); G = H2, N2 (γ = 1.4); G = SF6 (γ = 1.33); G = SiBr4 (γ = 1.07)a

To improve laser beam overlap and flow control, it can be advantageous to pass the gas first through a contoured supersonic nozzle, before allowing it to expand as a free jet into the jet chamber, as illustrated in Figure 1. The gas can be pre-expanded smoothly in a nozzle (with less flaring) for example from ∼5 torr at T = 300 K in the feed chamber to about 0.1 torr and T ∼ 63 K at the nozzle exit, using an axi-symmetric Mach-3.4 nozzle with exit diameter of 0.6 cm (including boundary layer) and throat diameter of 0.26 cm. The gas mix subsequently enters the jet chamber where it expands further to p ≈ 0.005 torr and T ∼ 25 K with some flaring, forming a quasi-cylindrical jet. Actually any available Mach-3 or Mach-4 supersonic nozzle with throat radius 0.2–0.5 cm and exit radius 5–7 cm can be adapted for this function.

For increased flow rates slit nozzles with two-dimensional (instead of axisymmetric) jets can be used. In this case, the laser beam is passed cross-wise through the free jet (Eerkens, 1998). Pre-expansion of the gas mix through a two-dimensional supersonic slit nozzle is again desirable to tailor the free jet for better laser beam overlap.

7.2. Nucleation and particle growth

To determine whether significant concentrations of (QF6)N clusters could develop during the transit time ttr as the jet expands to the desired final temperature and partial pressure, we examine conditions that cause nucleation. Typically the final partial pressure of QF6 is pQ ∼ 2 × 10−4 torr if the final gas jet pressure is ppG ∼ 10−2 torr and yQ = 0.02. At adiabatic expansion temperatures of T ≤ 100 K, this pressure pQ is well above the equilibrium vapor pressure peQ of both SF6 and UF6 (and other QF6 species) listed in Table 5. Under normal equilibrium this would cause QF6 to condense, but in the gas phase the process is retarded because of the reduced surface tension (Kelvin effect) and reduced binding energy per monomer of small clusters. The actual pressure pdQ at which QF6 vapor would start to nucleate and grow (QF6)N clusters in the gas is much higher than peQ. As reviewed in Eerkens (2003), depending on pressure pQ and temperature T, a “critical” particle size r* or monomer loading N* can be calculated, requiring time tc to grow. N-mers with N < N* do not grow since they can not keep additional monomers on their surface during repeated collisions. That is, all sub-critical N-mers (dimers, trimers, …) exist only transiently in extremely small concentrations (Eerkens, 2001a).

Calculated nucleation and particle growth parameters for (QF6)N

At T > ∼150 K, the concentration of dimers at any instant is much lower than that of monomers, while trimer populations are orders-of magnitude smaller than those of dimers, etc. Thus nucleation is of no concern. However for a condensable vapor which is suddenly cooled in an expanding free jet to T < 100 K, significant particle (“snow”) formation might occur since N*(T) drops rapidly as T drops and consequently the concentration of critical N*-mers can increase steeply. Even then, this can only happen if the time constant tc for growth is less than ttr (∼10−3s). Although heterodimers QF6:Ar and hetero-N-mers (QF6)N:ArK also form, we briefly examine growth of homo-N-mers (QF6)N, and assume the results for hetero-NK-mers are similar.

To determine N* as a function of T and pQ or φQ = [ell ]n(pQ/peQ), a quartic equation must be solved, while an estimate of tc = tc(N*) requires analysis of a kinetics rate equation (Eerkens, 2003). Computer-calculated values of N*(T) for some selected temperatures considered earlier, are listed in Table 5. For a given N*, tc can be calculated via the relations (Eerkens, 2003):

Here Dα and εα are the dimer-bond well-depth and fundamental frequency discussed in Section 3.4, while tc is the time it takes for 20% of the QF6 to have condensed into growing particles. Units in Eq. (110) are: T(K), MQ(amu), nQ(molecules/cm3), and ΩQ is the volume of one QF6 molecule in the condensate in Å3. The latter is readily calculated from the relation:

where ρQ is the density of QF6 (liquid or solid) condensate, tabulated in Physics Handbooks. Values of Dα, εα, and ΩQ for SF6 and UF6 are listed in Table 5. Some tc times calculated by (105)–(111) are tabulated in Table 5, showing the shortest time tc ∼ 0.1 s. Thus for jet transit times of ttr < 10−3 s and pQ < 2 × 10−4 torr, there is not enough time for cluster development. However note from (110) that τ4 ∝ nQ−1 so clustering can occur if pQ > 10−2 torr.

Note from Table 5 that tc exceed 100 seconds both at high and low temperatures. At high T this is caused by the high value of N*(T). For example oligomers with N = 20, needed to start SF6 nucleation at T = 74 K, are extremely rare and have vanishingly small populations (< 10−80 cm−3). At temperatures of T = 10 K, the high tc values are due to low collision rates.

8. DISCUSSION AND CONCLUSIONS

To make an isotope separation scheme attractive, one needs a reasonable unit throughput Funit (moles of feed per hour) and energy consumption EC = Pe/Funit (kWhr/mole) per separator, besides a good enrichment factor β (> ∼1.8), high Θo (> ∼0.08), and low γ (< ∼0.93). There are many isotope separation processes (e.g., calutrons) with high β's but low Funit and high EC which are uneconomic compared to other separation methods.

Our analyses showed that to achieve high CRISLA enrichment factors β with good cuts Θo, operating pressures of less than 0.1 torr are necessary. If pGptot ∼ 10−2 torr, the QF6 partial pressure in the jet core is only pQ ∼ 2 × 10−4 torr if yQ = 0.02 (to minimize VV losses). This compares with pressures of 2 × 10−5 torr used in electromagnetic calutron isotope enrichments, whose economics is poor compared to ultracentrifuge isotope separations (presently the favored non-laser method) because of low throughputs. Fortunately a free jet moves at supersonic velocities (as do ions in a calutron) which partially off-sets low operating pressures.

With supersonic velocities of U ∼ 3 × 104 cm/s, QF6 feed through-puts will be on the order of 2.5 × 104 cm3/s per separator or approximately 0.01 moles/hr, if pQ = 2 × 10−4 torr, T = 25 K, assuming a single axisymmetric jet with (average) radius R = 0.5 cm. For small-quantity isotope productions needed in nuclear medicine, this rate is adequate. For larger scale operations the rate per separator can be substantially increased by utilizing two-dimensional slit nozzles and free-jets which are laser-irradiated cross-wise (Eerkens, 1998).

An advantage of laser isotope separation schemes over calutrons and other mass-action methods is that in the laser case, energy only needs to be spent on a specific isotopomer, whereas in calutrons the electron discharge must ionize every isotope of the atomic element. Also, excluding preparation and other losses, only ∼0.1 eV per desired isotope is spent in CRISLA, instead of ∼6 eV to ionize every isotope of the element in calutron separations. Thus a laser scheme can save energy if laser photon generation is efficient.

As mentioned, another advantage of the CRISLA process with free-jet harvesting is that feed and product streams are the same gas. Chemical reprocessing of products or tails as required in other MLIS schemes, is absent. Thus several enriching stages can be put directly in series in CRISLA to produce the desired final level of isotope enrichment.

In the analyses, it was tacitly assumed that hetero-dimers iQF6*:G provided the main source of isotope-enriching epithermal iQF6! molecules. Homo-dimers iQF6*:QF6 also form and after predissociation could give high kinetic energies after VT conversion (see Eq. 42). However for iQF6*:QF6 homo-dimers there is a high probability for VV transfer and isotope-scrambling prior to predissociation. This diminishes isotope enrichment and is undesirable. To minimize VV transfers, concentrations yQ of QF6 in the carrier gas are kept low (yQ < ∼0.05).

Because the mole fraction yQ of QF6 in carrier gas G is low, and a low abundance xi of laser-irradiated isotope iQ makes the concentration yQxi even lower, another tacit assumption made was that the absorbed and VT-released laser energy is insufficient to heat jet gases. If xi and/or yQ are high, laser heating of the jet gases might have to be considered. In most practical cases however, yQxi << 1 and this effect can be ignored.

It might appear desirable to have high iQF6! escape factors and thus to have the collision mean-free-path [ell ]c be on the order of the jet core radius Rj. However with [ell ]cRj, one would have nearly collisionless slip-flow and Knudsen-regime conditions. This would preclude enough collisions of iQF6* with G to produce an adequate population of iQF6*:G species for achieving isotope separation. The optimum gas pressures for isotope separation are thus in the low-pressure region of continuum fluid mechanics, where gaseous diffusion relations are still valid.

According to (42), a carrier gas G with high mass MG gives the highest kinetic energy to iQF6 and yields the highest β's, provided G does not possess any internal vibrations with a frequency close to ν3 of QF6. Thus use of heavy carrier gases such as Br2 (M = 160), SiCl4 (M = 172), SiBr4 (M = 348), whose vapor pressures still exceed 10 torr at room temperature, might appear advantageous at first glance. However Table 4 shows that a very low value of γ (e.g., γ = 1.07 for SiBr4) requires impossibly high reservoir pressures and small nozzle throat diameters. This suggests use of a three-component gas mixtures such as QF6 + Ar + SiBr4 studied in earlier CRISLA experiments (Eerkens, 1998; Eerkens et al., 1995, 1996). The effect of two carrier gases G1 and G2 on β and Θo of iQF6 will be left to a future study.

Noble gases would be most convenient as carrier gases but do not yield the very highest enrichments. Xe gives the highest β's of all noble gas carriers because of its high mass (M = 131) and favorable γ (= 1.67), but it is rather expensive. SF6 (M = 146) as carrier gas is less expensive and with γ = 1.33 is still usable for supersonic expansions according to Table 4. Thus for UF6 enrichments, carrier gas SF6 is preferred, provided VV transfers of UF6's ν3 quanta at 628 cm−1 to SF6's ν4 vibration at 616 cm−1 are minor. The advantage of SF6 is also that it can be cryo-pumped. At 77 K (liquid N2), the vapor pressure of SF6 is on the order of 10−5 torr, precisely the desired sink pressure for evacuating rim gases from the free-jet chamber.

Note that direct epithermal escapes are mostly operative in the outer region of the jet core at a distance sth ∼ 1.41 NT1/2[ell ]c from the imaginary “wall”(see Eq (53)). Thus at higher pressures where [ell ]c is small, only laser excitations in the annular region between (Rsth) and R of the jet core donate high-speed core-fleeing isotopomers directly. In the laser-illuminated interior of the jet core, most epithermals are slowed down before they can reach the jet “wall.” In the interior region, the main contribution to isotope separation comes from non-excited isotopomers jQF6 which form long-lived heavy jQF6:G dimers. These dimers migrate more slowly than laser-excitable iQF6 monomers which migrate faster at monomer (thermal or epi-thermal) speeds.

Of course the calculated β's in Figures 3, 4, 5, and 6 are only estimates because of mathematical approximations and uncertainties in the values of some physical constants. Nevertheless the theory should predict trends correctly and provide useful approximate values of enrichment factors needed for the design of CRISLA separators. Though we focussed on 33SF6 and 235UF6 isotope separations, the equations can be used equally well for 123TeF6 and 98MoF6 or any other iQF6 or iQXmYn isotopomer, assuming lasers with suitable frequencies are available.

ACKNOWLEDGMENTS

I am indebted to Dr. W.H. Miller at the University of Missouri (MU) for programming the computer-calculated curves of Figs 2 through 9. To former MU student Dr. Jaewoo Kim, I am grateful for performing some of the experiments that led me to investigate and re-examine condensation repression physics in free jets. I also wish to thank Dr. S. Loyalka, Dr. W.H. Miller, and Dr. M.A. Prelas at MU, as well as Dr. J.F. Kunze at Idaho State University for valuable discussions and support of my laser isotope separation research.

References

REFERENCES

Adamson, T.C. & Nichols, J.A. (1959). On the structure of jets from highly underexpanded nozzles into still air. J. Aerospace Sciences 1, 23.Google Scholar
Benedict, M., Pigford, T.H. & Levi, H. (1981). Nuclear Chemical Engineering, 2nd Edition, New York, McGraw-Hill.
Bernstein, L.S. & Kolb, C.E. (1979). Understanding the IR continuum spectrum of the N2O dimer and other VanderWaals complexes at low temp's. J. Chem. Phys. 71.Google Scholar
Beswick, J.A. & Jortner, J. (1978). Perpendicular vibrational predissociation of t-shaped VanderWaals molecules. J. Chem. Phys. 69, 512.CrossRefGoogle Scholar
Beu, T.A. & Takeuchi, K. (1995). Structure and IR-spectrum calculations for small SF6 clusters. J. Chem. Phys. 103, 6394.CrossRefGoogle Scholar
Beu, T.A., Onoe, J. & Takeuchi, K. (1997). Calculations of structure and IR-spectrum for small UF6 clusters. J. Chem. Phys. 106, 5910.CrossRefGoogle Scholar
Burak, I., Houston, P., Sutton, D.G. & Steinfeld, J.I. (1970). Observation of laser-induced acoustic waves in SF6. J. Chem. Phys. 53, 9.CrossRefGoogle Scholar
Chow, R.R. (1959). On the separation of a binary gas mixture in an axisymmetric jet. Tech Report HE-150-175 (Series 110, Issue No. 5) University of California-Berkeley, Institute of Engineering Research.
Eerkens, J.W. (1957a). Design considerations and region of operation of the jet diffusion equipment. Tech Report HE-150-158 (Series 110, Issue No. 3) University of California-Berkeley, Institute of Engineering Research.
Eerkens, J.W. & Grossman, L.M. (1957b). Evaluation of the diffusion equation and tabulation of experimental gaseous diffusion constants. Tech Report HE-150-150 (Series 110, Issue No. 1). University of California-Berkeley, Institute of Engineering Research.
Eerkens, J.W., Sehgal, B. & Grossman, L.M. (1958). Some preliminary results on the separation of gaseous components in a jet. Tech Report HE-150-162 (Series 110, Issue No. 4). University of California-Berkeley, Institute of Engineering Research.
Eerkens, J.W. (1973). Model equations for photon emission rates and absorption cross-sections (Rocket Radiation Handbook Volume II), USAF-FTD-CW-01-01-74, AD-A007946.
Eerkens, J.W., Kunze, J.F. & Anderson, G. (1995). Mo-98 and Mo-99 laser isotope separation for nuclear medicine. Proc. Missouri Ac. of Science, Eng'ng Section.
Eerkens, J.W., Miller, W.H. & Puglisi, D. (1996). Laser separation of medical isotopes—QF5Cl Spectra. ANS Transactions of Nov. 11–14>, 1996, Washington DC Meeting.
Eerkens, J.W. (1998). Separation of isotopes by laser-assisted retardation of condensation (SILARC). Laser Part. Beams 16, 295.CrossRefGoogle Scholar
Eerkens, J.W. (2001a). Equilibrium dimer concentrations in gases and gas mixtures. Chem. Phys. 269, 189.Google Scholar
Eerkens, J.W. (2001b). Average diffusion-to-the-wall times for laser-tagged molecules in a long cylinder. Appl. Phys. B 72, 885.Google Scholar
Eerkens, J.W. (2003). Nucleation and particle growth in supercooled flows of SF6/N2 and UF6/N2 mixtures. Chem. Phys. 293, 111153.CrossRefGoogle Scholar
Eerkens, J.W. (2005). Condensation reduction and isotope separation of laser-excited molecules in wall-cooled subsonic gas streams, Nucl. Sci. Eng. 150, 122.Google Scholar
Geraedts, J., Setiadi, S., Stolte, S. & Reuss, J. (1981). Laser induced predissociation of SF6 clusters. Chem. Phys. Let. 78, 277.CrossRefGoogle Scholar
Geraedts, J., Stolte, S. & Reuss, J. (1982). Vibrational predissociation of SF6 dimers and trimers. Z. Physik A 304, 167175.CrossRefGoogle Scholar
Glasstone, S. & Edlund, M.C. (1952). The elements of nuclear reactor theory, Chapter 6; New York: Van Nostrand.
Herzberg, G. (1945). Infrared and raman spectra, New York: Van Nostrand.
Hirschfelder, J.O., Curtiss, C.F. & Bird, R.B. (1967). Molecular Theory of Gases and Liquids. Footnote on page 555 of Section 4a. John Wiley, New York, Fourth Printing 1967.
Kim, K.C. & Person, W.B. (1981). Study of absolute IR intensity of v3 absorption of UF6. J. Chem. Phys. 74, 171.CrossRefGoogle Scholar
Kraus, M. (1967). Compendium of ab initio Calculations of Molecular Energies and Properties, National Bureau of Standards, Technical Note NBS TN 438, Dec 1967, US Government Printing Office, Washington DC.
Lee, Y.T. (1977). Isotope separation by photodissociation of VanderWaals molecules. U.S. Patent 4,032,306, issued June 1977.
Liedenbaum, C., Heijmen, B., Stolte, S. & Reuss, J. (1989). IR predissociation of halogenated methane VanderWaals complexes. Z. Physik D 11, 175180.Google Scholar
Liepmann, H.W. & Puckett, A.E. (1953). Introduction to Aerodynamics of a Compressible Fluid, Ronald Press.
McCafferey, A.J. & Marsh, R.J. (2002). Vibrational predissociation of vanderWaals molecules; An internal collision, angular momentum model. J. Chem Phys. 117, 9275.CrossRefGoogle Scholar
Schwartz, R.N., Slawski, Z.I. & Herzfeld, K.F. (1952). Calculation of vibrational relaxation times in gases. J. Chem. Phys. 20, 159.CrossRefGoogle Scholar
Schwartz, R.N. & Herzfeld, K.F. (1954). Vibrational relaxation times in gases (three-dimensional treatment). J. Chem. Phys. 22, 5.CrossRefGoogle Scholar
Smalley, R.E., Levy, D.H. & Wharton, L. (1976). The fluorescence excitation spectrum of the HeI2 VanderWaals complex. J. Chem. Phys. 64, 3266.CrossRefGoogle Scholar
Snells, M. & Reuss, J. (1987). Induction effects on IR-predissociation spectra of (SF6)2, (SiF4)2, and (SiH4)2. Chem. Phys. Let. 140, 543.CrossRefGoogle Scholar
VanBladel, J.W.I. & VanderAvoird, A. (1990). The infrared photodissociation spectra and the internal mobility of SF6, SiF4, and SiH4 dimers. J. Chem. Phys. 92, 2837.CrossRefGoogle Scholar
VandenBergh, H. (1985). Laser assisted aerodynamic isotope separation,” Laser und Optoelektronik 3, 263.Google Scholar
Yardley, J.T. (1980). Introduction to Molecular Energy Transfer, Academic Press, New York
Figure 0

Schematic of CRISLA chamber with free jet and skimmer.

Figure 1

Molecular parameters for SF6(ν3) and UF6(ν3) gas-phase collisional VT relaxations

Figure 2

Molecular parameters for SF6(ν3) and UF6(ν3) gas-phase collisional VT relaxations

Figure 3

Dimerization parameters for SF6/G and UF6/G encounters

Figure 4

VT conversion probabilities for SF6*(ν3)/G gas mixtures.

Figure 5

VT conversion probabilities for UF6*(ν3)/G gas mixtures.

Figure 6

Thermalization parameters for epithermal SF6! and UF6! molecules

Figure 7

Enrichment factors βm(T) for SF6(ν3) monomer excitation at fixed pG.

Figure 8

Enrichment factors βm(pG) for SF6(ν3) monomer excitation at fixed T.

Figure 9

Enrichment factors βm(T) for UF6(ν3) monomer excitation at fixed pG.

Figure 10

Enrichment factors βm(pG) for UF6(ν3) monomer excitation at fixed T.

Figure 11

Enrichment factors βd(pG,T) for SF6:G dimer laser excitations.

Figure 12

Enrichment factors βd(pG,T) for UF6:G dimer laser excitations.

Figure 13

Isotope product cuts Θ(T) for 33SF6 at fixed pG.

Figure 14

Isotope product cuts Θ(pG) for 33SF6 at fixed T.

Figure 15

Isotope product cuts Θ(T) for 235UF6 at fixed pG.

Figure 16

Isotope product cuts Θ(pG) for 235UF6 at fixed T.

Figure 17

Measured free-jet contour and pressure distribution for G = Air; po = 3.86; p∞ = 0.21 torr (after Eerkens et al., 1958; Chow, 1959).

Figure 18

Measured free-jet contour and pressure distribution for G = Air; po = 0.80; p∞ = 0.15 torr. (after Eerkens et al., 1958; Chow, 1959).

Figure 19

Calculated one-step supersonic jet expansion parameters for QF6/G with G = He, Ar, Xe (γ = 1.67); G = H2, N2 (γ = 1.4); G = SF6 (γ = 1.33); G = SiBr4 (γ = 1.07)a

Figure 20

Calculated nucleation and particle growth parameters for (QF6)N