Sunteți pe pagina 1din 13

Ref. p.

285]

3.5 Excimer lasers

275

3.5 Excimer lasers


U. Sowada

3.5.1 Introduction
A molecule with a dissociative lower laser state is called excimer, if it consists of two atoms of the same kind. In the case of dissimilar atoms it is called exciplex. These words are derived from excited-state dimer or excited-state complex, i.e. they denote a molecule in a bound, excited state. At present all lasers using stimulated emission with this type of molecular species are called excimer lasers, regardless whether they are based on an excited-state complex or excited-state dimer. Excimers and exciplexes are particularly promising for laser action, because population inversion is readily achieved, once a few molecules are present. Fast dissociation of the lower laser state excludes the process of induced absorption, a decisive photon loss mechanism in normal two-state systems. In order to guarantee unrestricted molecular formation and dissociation reactions, the laser active medium of an excimer laser needs to be in the gaseous state. The output of excimer lasers is decisively inuenced by chemical reactions among the many short-lived intermediates following the pumping process. Diatomic excimers often contain at least one rare-gas atom, which due to its chemical inertness is a promising candidate to have a dissociative ground state. The lowest excited electronic state of a rare-gas atom produces a species with a single electron in the outermost shell, resembling an alkali metal. These elements are known to form stable compounds with halogen atoms, since halogens lack just one electron. Therefore, the most obvious candidates for excimer lasers are the rare-gas monohalides.

3.5.2 Wavelengths and stimulated emission cross sections


3.5.2.1 Rare-gas halogen excimers
3.5.2.1.1 Rare-gas monohalides Table 3.5.1 displays the wavelengths of rare-gas monohalide excimers. Several dierent wavelengths have been observed, depending upon upper and lower state in the transition. For the B-X-transitions we note the following tendency: The lighter the rare-gas atom, the shorter is the wavelength of the transition. The dependence with regard to the mass of the halogen is reversed: The lighter the halogen, the longer is the wavelength . The uorescence line widths are small ( is about 2 nm), i.e. the transitions are best represented by one of the graphs in Fig. 3.5.1b or c. The C-A-transitions are spectroscopically ten times wider (about 25 nm full width at half maximum), thus are represented by the scheme in Fig. 3.5.1a. As in the case of the B-X-transitions, increasing the mass of the rare gas will increase the wavelength, increasing the

Landolt-Brnstein o New Series VIII/1B1

276

3.5.2 Wavelengths and stimulated emission cross sections

[Ref. p. 285

Table 3.5.1. Wavelengths observed in rare-gas monohalide excimer lasers. B-X-transition [nm] NeF ArF ArCl KrF KrCl KrBr XeF XeCl XeBr XeI 108 193 175 249 222 206 351 308 282 253 Reference [77Ric] [76Hof] [77Way] [76Tel2] [76Mur2] [75Gol] [76Tel3] [76Tel1] [76Tel2] [76Tel2] C-A-transition [nm] Reference

275 240 222 460 345 300 265

[84Rho] [84Rho] [84Rho] [84Rho] [84Rho] [84Rho] [84Rho]

Energy E

Internuclear separation

Energy E

Internuclear separation

Energy E

Internuclear separation

Fig. 3.5.1. Transitions from an upper bound molecular state to a state with shallow minimum in the potential energy of the ground state. Depending upon the relative positions of the minima the bandwidth of the transition will be either spectrally broad (a) or spectrally narrow (b), (c), as indicated by the variation in length of the three arrows.

mass of the halogen will decrease it. The C-A-transition of XeF has been reported to show an electrical to optical eciency of roughly 0.1 %, output energy 100 J/(m3 atm), when pumped with an electron beam [83Nig]. The product of cross section for stimulated transition and lifetime for spontaneous decay is related to the transition wavelength and the bandwidth by (3.5.1) [66Len]: = 1 4 ln (2)
1/2

4 . c

(3.5.1)

Landolt-Brnstein o New Series VIII/1B1

Ref. p. 285]

3.5 Excimer lasers

277

The width of the band is meant to be observed in uorescence; c is the speed of light. The predictions and experimental results for the B-X-transitions in the two most important excimers KrF and XeCl are given in Table 3.5.2. The values of the cross section for a stimulated C-Atransition will be smaller by about a factor of 10, since the width of the band is about ten times greater.
Table 3.5.2. Cross sections for stimulated transition in KrF and XeCl (B-X-transition). [m2 ns] KrF XeCl 2.4 1028 5.6 1028 (expt.) [ns] 9 [76Nak, 77Bur] 7 [66Cal] 11 [82Mae] [1020 m2 ] 2.7 3.5 5.1 (expt.) [1020 m2 ] 4.0 0.4 [84Sza] 2.5 [81Lev]

Of the experimentally determined stimulated transition cross sections (last column of Table 3.5.2) only that one of [84Sza] appears to be accurate, since it has been measured with a short enough probe pulse to avoid repopulation of the upper laser state.

3.5.2.1.2 Polyatomic rare-gas halogen excimers Two dierent families of triatomic rare-gas halogen excimers have been studied, rst the type where two similar rare-gas atoms are bound to a halogen atom, and second the type where all three atoms are dierent. In Table 3.5.3 the wavelengths and the reported band widths as observed in uorescence are listed. Where no number is given, the width is not reported in the literature. The last column contains the reported (however, theoretical) cross sections for stimulated emission (3.5.1) [84Rho, p. 202].
Table 3.5.3. Triatomic rare-gas halides, wavelengths and cross sections for stimulated transition. Wavelength [nm] Ar2 F Ar2 Cl Kr2 F Kr2 Cl Kr2 Br Xe2 F Xe2 Cl Xe2 Br Xe2 I ArKrF ArKrCl KrXeCl KrXeBr KrXeI 285 245 420 325 318 610 490 440 375 305 270 370 330 260 Width of the band [nm] 50 30 70 30 130 80 60 70 50 80 50 80 Cross section [1022 m2 ] 0.95 3.28

14.05 4.56 2.65

Landolt-Brnstein o New Series VIII/1B1

278

3.5.3 Chemical reactions in the discharge

[Ref. p. 285

3.5.2.2 Rare-gas excimers


Since the positively charged dimer ions of the rare gases are stable, it is reasonable to investigate the pure rare-gas dimers. Table 3.5.4 gives the wavelengths as observed in uorescence or laser experiments.
Table 3.5.4. Excimers in the pure rare gases. Wavelength [nm] Ar 2 Xe 2 126 172 Reference [79Bra] [76Lor]

In the homonuclear rare-gas excimers two dierent electronic states with dierent radiative lifetimes can be distinguished: 1 u , 3 u [76Lor]. As a function of gas pressure laser power has been found to display a maximum at 20 atm [73Hof, 73Wal]. At high density of the excimer states a reaction may occur between excimers with the result of deexcitation. Counterproductive to laser action is the short wavelength which except for Xe appears to exclude amplication by stimulated 2 emission, since photoionization of the excimer seems to have a larger cross section.

3.5.2.3 Halogen excimers


Even though halogens have a stable molecular ground state, they are interesting candidates to be used in excimer lasers. The normal ground state has electronic symmetry. The lowest state is metastable, and transitions in the -system are useful for this purpose. Particularly fascinating is the F -excimer because of its short wavelength. The bandwidths of the halogen-excimers are small 2 (ca. 2 nm). In Table 3.5.5 the wavelengths of the homonuclear halogen excimers are listed.
Table 3.5.5. Wavelengths of the homonuclear halogen excimers. Wavelength [nm] F 2 Br 2 I 2 158 293 342 Reference [77Ric] [76Ewi, 76Mur1] [75Egg, 77Bur, 76Hay]

3.5.3 Chemical reactions in the discharge


Understanding the chemical reactions in the gas mixture and the role of the many intermediates is very important in computer modeling. Even the apparently simple gas mixtures with only three components in a normal discharge-pumped excimer laser exhibit a large number of possible reactions, because on the nanosecond time scale many short-lived, reactive species are formed. If numerical calculations of the behavior during discharge are attempted, all possible reaction paths

Landolt-Brnstein o New Series VIII/1B1

Ref. p. 285]

3.5 Excimer lasers

279

need to be included. As a result, optimal gas mixtures can be proposed, and temporal aspects of the beam can be related to the reactions. As examples for important excimer laser systems, two gas mixtures are considered, the rst one containing helium (He), xenon (Xe), and hydrochloric acid (HCl), the second one argon (Ar), krypton (Kr), and uorine (F2 ). HCl must be used as a halogen donor instead of Cl2 , because Cl2 is strongly absorbing light, in contrast to F2 . In chemical reactions the variation in time of the concentration of the reaction product P is proportional to the rate constant k and the product of the single concentrations of the reactants R. If we have n equal reactants R reacting to form the product P , we get (3.5.2): d n [P ] = k [R] . dt (3.5.2)

The square brackets here denote concentrations with unit cm3 . If we have dissimilar reactants R1 , R2 , . . . , Rn , (3.5.2) reads instead: d [P ] = k [R1 ] [R2 ] . . . [Rn ] . dt (3.5.3)

The proportionality constant k is called the reaction rate constant. It has the unit s1 (for n = 1), cm3 s1 (for n = 2), cm6 s1 (for n = 3), and so on. The reaction with n = 1 is called rst order, with n = 2 it is second order. Reactions described by (3.5.2) or (3.5.3) are called nth order. The value of a chemical rate constant k depends upon the energy of the molecules, here expressed by the gas temperature Tg , in case of electron reactions on electron energy, expressed by electron temperature Te . In this compilation these two temperatures are assumed to be given in electron volts. The positive ion of xenon is denoted Xe+ , the excited state Xe , the next higher excited state Xe (p). The symbol e is used for a free, kinetic electron, HCl(v) means vibrationally excited, XeCl(X) denotes the excited electronic molecular state. In Table 3.5.6 reactions in the gas mixture containing He, Xe, and HCl, following discharge pumping are listed. The process of stimulated emission with subsequent ground-state dissociation (No. 38) is decisive for laser action. Reactions with rst-order kinetic behavior occur only rarely in Table 3.5.6 (e.g. the spontaneous radiative decay of the excimer molecule, reaction No. 37). Most common is a second-order kinetic reaction (e.g. No. 1). Third-order reactions need to be included, too (e.g. No. 18). Photon absorption reactions (Nos. 38 through 42) are better described with cross sections than rate constants. Next to the ionic recombination reaction (No. 21), there are 4 more reactions leading to the excimer XeCl (Nos. 22, 24, 25, 27). The ingredients required in these 5 reactions have to be formed rst, e.g. by dissociative electron attachment to HCl (No. 16). If the electron attachment rate to HCl is too high, not enough electrons may be present to sustain the discharge through impact ionization (Nos. 1 and 11). This is the reason why the concentration of HCl must be kept very low. Reaction rates of the excimer XeCl , which may occur before stimulated emission, must be kept low enough, too (Nos. 28 through 37). The negative halogen ion is a very important ingredient, since it is one of the partners in creating the excimer molecule (No. 21). Reactions of this ion, which reduce its concentration in the discharge, will reduce the laser power. Due to the short wavelength of the photon eld in the resonator, electron photodetachment becomes a possible reaction (No. 42). It has been discovered [89Osb], however, that electron photodetachment has a positive eect on the discharge, since it restores free electrons to sustain the discharge and produce more electrons and positive rare-gas ions (reaction No. 1). It is illustrative to add another table (Table 3.5.7) with the reactions occurring in a gas mixture containing argon (Ar), krypton (Kr), and uorine (F2 ) [84Rho, pp. 131132]. The relevant excimer molecule to be formed is KrF .
Landolt-Brnstein o New Series VIII/1B1

280

3.5.3 Chemical reactions in the discharge

[Ref. p. 285

Table 3.5.6. Reactions in the gas mixture containing He, Xe, and HCl, following discharge pumping. No. Reaction Cross section [cm2 ] or reaction rate constant k ([s1 ] or [cm3 s1 ] or [cm6 s1 ])
0.72 k1 = 6.05 108 Tg exp(21.98/Te ) 8 0.72 k2 = 1.2 10 Tg exp(8.36/Te ) 0.72 k3 = 6.08 108 Tg exp(15.42/Te ) 8 0.71 k4 = 7.85 10 Tg exp(3.77/Te ) 7 0.72 k5 = 2.15 10 Tg exp(2.4/Te ) 7 0.5 k6 = 2 10 Te 0.5 k7 = 2 105 Te 10 k8 = 5 10 k9 = 1.5 107 0.79 k10 = 1.1 106 Tg exp(1.37/Te ) 9 0.72 k11 = 2.9 10 Tg exp(51.96/Te ) 9 0.72 k12 = 1.96 10 Tg exp(36.87/Te ) 8 0.72 k13 = 3.73 10 Tg exp(4.7/Te ) 8 0.5 k14 = 2.4 10 Te k15 = 2 109 1.5 k16 = 3 1010 Te k17 = 2 108 k18 = 2.5 1031 k19 = 1.1 1031 k20 = 6.5 1032 k21 = 2 106 (Tg /0.026)2.5 k22 = 2 106 (Tg /0.026)2.5 k23 = 5.6 1010 k24 = 5.6 1010 k25 = 6 1011 k26 = 5 1010 k27 = 5.6 1010 k28 = 1.4 109 k29 = 7.7 1010 k30 = 3.2 1011 k31 = 1012 k32 = 5 1032 k33 = 1.5 1031 k34 = 7.3 1031 k35 = 3 107 k36 = 1.5 1031 k37 = 9.3 107 38 = 2.5 1016 39 = 2.5 1017 40 = 6 1020 41 = 1018 42 = 2.1 1017 k43 = 3.5 1010 (0.026/Te ) k44 = 1 1010 (0.026/Te ) k45 = 3.5 1027 (0.026/Te ) k46 = 1 1027 (0.026/Te ) k47 = 2.8 1020 (0026/Te )4 k48 = 8 1021 (0.026/Te )4

Reference

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48

e + Xe Xe+ + 2e e + Xe Xe + e e + Xe Xe (p) + e e + Xe Xe+ + 2e e + Xe (p) Xe+ + 2e e + Xe+ Xe + Xe 2 e + Xe+ Xe (p) + Xe 2 Xe + Xe Xe+ + Xe + e Xe (p) Xe + h e + Xe Xe (p) + e e + He He+ + 2e e + He He + e e + He He+ + 2e e + HCl HCl(v) + e e + HCl H + Cl + e e + HCl H + Cl e + HCl(v) H + Cl Xe+ + Xe + Xe Xe+ + Xe 2 Xe+ + Xe + He Xe+ + He 2 He+ + He + He He+ + He 2 Xe+ + Cl XeCl Xe+ + Cl XeCl + Xe 2 Xe + HCl Xe + H + Cl Xe (p) + HCl XeCl + H Xe + HCl(v) XeCl + H Xe + HCl(v) Xe + H + Cl Xe (p) + HCl(v) XeCl + H XeCl + HCl Xe + HCl + Cl XeCl + HCl(v) Xe + HCl + Cl XeCl + Xe 2Xe + Cl XeCl + He He + Xe + Cl XeCl + 2He 2He + Xe + Cl XeCl + Xe + He 2Xe + He + Cl XeCl + 2Xe 3Xe + Cl XeCl + e Xe + Cl + e XeCl + Xe + He Xe + Cl + He 2 XeCl XeCl(X) + h XeCl + h Xe + Cl + 2h Xe+ + h Xe+ + Xe 2 Xe + h Xe+ + e Xe (p) + h Xe+ + e Cl + h Cl + e He+ + e He + He + h 2 He+ + e He + h 2 2 He+ + He + e He + 2He 2 He+ + He + e He + He + h 2 2 He+ + 2e He + He + e 2 He+ + 2e He + e 2 2

[82Shu] [81Lev] [82Shu] [81Lev, 82Shu] [81Lev] [79Bra] [79Bra] [81Lev] [81Lev] [81Lev] [82Shu] [82Shu] [82Shu] [82Shu] [81Dem] [80Nig] [80Nig] [81Lev] [79Bra] [81Miz] [80Bar] [80Bar] [81Lev] [81Lev] [81Lev] [81Lev] [81Lev] [82Mae] [82Mae] [82Mae] [82Mae] [81Miz] [81Miz] [79Bra] [81Lev] [80Bor] [82Mae] [81Lev] [81Lev] [82Mae] [81Lev] [81Lev] [81Miz] [81Miz] [81Miz] [81Miz] [81Miz] [81Miz]
Landolt-Brnstein o New Series VIII/1B1

Ref. p. 285]

3.5 Excimer lasers

281

Table 3.5.7. Reactions in the gas mixture containing Ar, Kr, and F2 , following discharge pumping. No. Reaction Cross section [cm2 ] or reaction rate constant k ([s1 ] or [cm3 s1 ] or [cm6 s1 ]) k1 = 8.1 1010 k2 = 3 1010 k3 = 1 106 k4 = 8.5 1010 k5 = 1 106 k6 = 8.1 1010 k7 = 8.5 1010 k8 = 1 106 k9 = 1 106 k10 = 1 106 k11 = 6 1010 k12 = 2 1011 k13 = 0 k14 = 8 1032 k15 = 5 1032 k16 = 2 1011 k17 = 2 1011 k18 = 6.5 1031 k19 = 1 109 k20 = 3 1010 k21 = 3 1010 k22 = 1 109 k23 = 2.5 1010 k24 = 3 1010 k25 = 1 109 k26 = 1 1010 k27 = 1 1010 k28 = 1 1032 k29 = 1 1032 k30 = 1 1032 k31 = 1.14 1032 k32 = 6 1012 k33 = 4 1010 k34 = 2.5 1031 k35 = 7.5 1010 k36 = 1 1031 k37 = 1 1031 k38 = 1 106 k39 = 3.2 1010 k40 = 1 107 k41 = 2.5 1031 k42 = 3 1011 Reference

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42

Kr + F2 KrF + F ArF + Kr KrF + Ar Kr+ + F KrF Ar + F2 ArF + F Ar+ + F ArF Kr (p) + F2 KrF + F Ar (p) + F2 ArF + F Kr+ + F KrF + Kr 2 Ar+ + F ArF + Ar 2 ArKr+ + F KrF + Ar ArKr + F2 KrF + Ar + F KrF + Kr 2Kr + F KrF + Ar Kr + Ar + F KrF + 2Ar ArKrF + Ar ArF + 2Ar Ar2 F + Ar Kr + ArKrF Kr2 F + Ar Ar + ArKrF Ar2 F + Kr KrF + Kr + Ar Kr2 F + Ar ArKrF + F2 Ar + Kr + F + F2 Kr + F2 Kr2 F + F 2 Kr + F KrF + Kr 2 Kr2 F + F2 2Kr + F + F2 Ar + F2 Ar2 F + F 2 Ar + F ArF + Ar 2 Ar2 F + F2 2Ar + F + F2 Ar2 F + Kr ArKrF + Ar ArKr + Kr Kr + Ar 2 Kr + 2Ar ArKr + Ar Ar + Kr + Ar ArKr + Ar Kr + Kr + Ar Kr + Ar 2 Ar + 2Ar Ar + Ar 2 Ar + Kr Kr + Ar Ar + Kr Kr + 2Ar 2 Ar+ + 2Ar Ar+ + Ar 2 Ar+ + Kr Kr+ + 2Ar 2 Ar+ + Kr + Ar ArKr+ + Ar Kr+ + 2Ar ArKr+ + Ar ArKr+ + F ArKrF ArKr+ + Kr Kr+ + Ar 2 ArKr+ + e Kr (p) + Ar Kr+ + 2Kr Kr+ + Kr 2 Ar+ + Kr Kr+ + Ar

[76Nak] [76Nak, 77Man] [77Man, 78Bar] [76Nak] [78Bar] [76Nak] [76Nak] [76Nak, 78Bar, 76Jac1] [76Nak, 78Bar, 76Jac1] [79Bra] (estimated) [76Nak] [76Nak] [79Bra] (estimated) [77Man, 77Shu] [76Nak] [77Man] [77Man] [78Rok] [76Nak] [76Nak] [76Nak] [76Nak] [76Nak] [76Nak] [76Nak] [76Nak] [76Nak] [76Nak] [76Nak] [76Nak] [76Nak] [73Pip, 76Nak] [77Man] [70McD] [70Boh] [78Lac] [79Bra] (estimated) [79Bra] (estimated) [79Bra] (estimated) [79Bra] (estimated) [70McD] [70Boh] (continued)

Landolt-Brnstein o New Series VIII/1B1

282
Table 3.5.7 continued. No. Reaction

3.5.3 Chemical reactions in the discharge

[Ref. p. 285

Cross section [cm2 ] or reaction rate constant k ([s1 ] or [cm3 s1 ] or [cm6 s1 ]) k43 = 2.5 1031 k44 = 1 1033 k45 = 1.1 109 k45 = 1 108 k46 = 4.1 1011 k47 = 6.4 107 k48 = 1.5 1015 k49 = 4.8 108 k50 = 8 107 k51 = 8 1011 k52 = 1.8 107 k53 = 2.8 1012 k54 = 6.6 107 k55 = 1.4 1019 k56 = 2.8 108 k57 = 9 107 k58 = 5.4 1012 k59 = 1.9 107 k60 = 1.1 107 k61 = 7.7 108 k62 = 3 1010 k63 = 1 1012 k64 = 1.5 108 k65 = 2.5 108 k66 = 3.3 106 k67 = 3.8 106 k68 = 6.7 107 k69 = 2 108 k70 = 3 106 k71 = 5 107 72 = 1.2 1020 73 = 5 1018 74 = 1.5 1017 75 = 1 1017 76 = 2.3 1018 77 = 4.5 1018 78 = 1 1019 79 = 3.2 1020 80 = 5 1018 81 = 1 1018 82 = 4.0 1016

Reference

43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60 61 62 63 64 65 66 67 68 69 70 71 72 73 74 75 76 77 78 79 80 81 82

Kr+ + Kr + Ar Kr+ + Ar 2 F + F + M F2 + M F2 + e F + F Kr + e Kr + e Kr + e Kr (p) + e Kr + e Kr+ + 2e Kr + e Kr+ + 2e Kr (p) + e Kr + e Kr + e Kr + e Kr (p) + e Kr+ + 2e Ar + e Ar + e Ar + e Ar (p) + e Ar + e Ar+ + 2e Ar + e Ar+ + 2e Ar (p) + e Ar + e Ar + e Ar + e Ar (p) + e Ar+ + 2e Kr+ + e Kr (p) + Kr 2 Ar + e Ar (p) + Ar 2 F2 + e 2F + e F + e F KrF Kr + F + h ArF Ar + F + h Kr 2Kr + h 2 Ar 2Ar + h 2 Kr2 F 2Kr + F + h Ar2 F 2Ar + F + h ArKr Ar + Kr + h ArKrF Ar + Kr + F + h F2 + h 2F F + h F + e Kr+ + h Kr+ + Kr 2 Ar+ + h Ar+ + Ar 2 Ar (p) + h Ar+ + e Kr (p) + h Kr+ + e Ar + h Ar+ + e Kr + h Kr+ + e Kr2 F + h products Ar2 F + h products KrF + h Kr + F + 2h

[77Man, 76Jac2] [79Bra] (estimated) [77Man, 77Che] [82Cha] [69Sha, 75Egg, 76Jac2] [76Jac1] [65Rap, 72Pet] [65Vri] [76Jac1] [69Sha, 75Egg, 76Jac2] [65Vri] [69Sha, 75Egg, 76Jac2] [76Jac1] [65Rap, 72Pet] [65Vri] [76Jac1] [69Sha, 75Egg, 76Jac2] [65Vri] [65Rap, 69Sha, 70Boh] [65Rap, 69Sha, 70Boh] [79Bra] (estimated) [79Bra] (estimated) [76Nak, 76Dun] [76Dun] [76Nak] [76Nak] [76Nak] [76Nak] [76Nak] [76Nak] [66Cal] [71Man] [78Mic] [78Mic] [79Bra] (estimated) [79Bra] (estimated) [79Bra] (estimated) [79Bra] (estimated) [79Bra] (estimated) [79Bra] (estimated) [84Sza]

Landolt-Brnstein o New Series VIII/1B1

Ref. p. 285]

3.5 Excimer lasers

283

All important information concerning discharge behavior or gain as a function of time can be extracted from this data, if the coupled dierential equations are solved and the rate constants properly applied. This second set of data in Table 3.5.7 appears to be more accurate than the one in the previous table, Table 3.5.6, because there is an even larger number of reaction pathways considered. Again we have electron reactions, including impact ionization (Nos. 48, 49, 50, 55, 56, 59) or attachment (Nos. 45 and 63). Radiation emission (Nos. 64 through 71) and absorption reactions (Nos. 72 through 82) are also included. Predictions of this scheme may in spite of this higher level of completeness be less reliable, however, since the reaction rate constants in Table 3.5.7 regarding KrF /Ar do not contain an energy dependence like the data for the XeCl /He in Table 3.5.6. Electron reactions are very strongly dependent on electron energy, and neglecting this will not allow to understand the consequences of increasing or decreasing the electric-eld strength which produces the discharge. Heating eects of the plasma by the electric discharge will increase the kinetic energy of the atoms and molecules, which cannot be without inuence on molecular reactions. But it is not possible for each of the many reactions to extract reliable data from the literature on the energy dependence of the rate constants. It is very instructive to compare the rate constants for dissociative electron attachment to the halogen molecule in Table 3.5.6 (No. 16) with that in Table 3.5.7 (No. 45). Even though literature data vary (probably due to the strong eect of electron energy), attachment to F2 seems to be faster than to HCl by a factor of 10. This is the principal reason for the dierent behavior of the discharge in excimer lasers. As a consequence, gas mixtures containing F2 will tend to arc rather than sustain a homogeneous glow discharge, while those containing HCl are less critical in this aspect. The data in Table 3.5.6 and Table 3.5.7 reveal another important tendency, which is of relevance to discharge stability. Argon atoms have a larger rate to be ionized by electron impact than helium (No. 55 in Table 3.5.7 compared to No. 13 in Table 3.5.6). This is quite natural, since the ionization energy is lower for the heavier noble gas. Glow discharges are most stable when the gas is at low pressure, i.e. localized discharge instabilities are less likely to form. At the high pressures present in excimer lasers, the lighter rare gas is better suited to be the buer gas, because it will not favor rapid electron density changes. Due to the relatively high photon energies of the laser transition in these excimer lasers we need to consider photodetachment, i.e. liberation of electrons from negative ions by photon absorption (No. 73 in Table 3.5.7). It is an important reaction because of two reasons: rst, it is a photon loss mechanism, and second, it is a mechanism to deliver electrons for further sustaining the discharge. For photodetachment from F theoretical predictions [67Rob, 74Ish, 78Res, 83Clo, 87Rad] and experimental results [71Man, 88Wan] show good agreement. A value of = 6 1022 m2 applies for photon energies just at threshold of 3.45 eV, which corresponds to a photon wavelength of 360 nm. Increasing the photon energy will reduce the cross section proportionally. Photodetachment will transform a lot of the still existing negative halogen ions into halogen atoms plus free electrons. Electron attachment will follow and transform the free electrons again into negative ions, however with a delay of a few nanoseconds. It may be argued that the observed slow recovery (2.5 ns for XeCl [82Cor] and 4 ns for KrF [84Sza]) of the upper laser state after depletion by an intense pulse of light is due to this eect.

Landolt-Brnstein o New Series VIII/1B1

284

3.5.4 Beam properties

[Ref. p. 285

3.5.4 Beam properties


3.5.4.1 Pulse energy and pulse duration
The rst excimer lasers have been pumped by a high-power electron beam [75Sea, 84Rho]. This is technically dicult. Pumping by a transverse high-voltage discharge like in normal TEA-lasers (TEA: Transverse Electrical discharge in gas at Atmospheric pressure) is a cheaper and more reliable method, and it is used successfully in all commercial lasers. Technological reasons restrict the useful pumped volume to several 100 cm3 , and the output energies of rare-gas monohalide excimers are usually in the range 0.1 . . . 1 J [86Pum]. Higher values for the pumped volume are possible by increasing the cross section with the aid of X-ray preionization [87Cir]. This way multi-Joule pulse energy values have been created. The high gas pressures produce signicant technical problems for longitudinally pumped excimer lasers [86Eic]. High-power microwaves have been shown to be applicable, too [91Kli]. Pulse duration in most cases is approximately 10 . . . 100 ns. Realization of longer pulses is restricted to XeCl [87Klo]. Short pulses can be made by injecting a short pulse of the correct wavelength, as obtained from a dye laser, and using the tube of the excimer laser as a single-pass amplier [84Sza].

3.5.4.2 Output power


High average output powers with excimer lasers scale with total gas pressure [85Ern], with repetition frequency and with pulse energy. Local safety regulations pose limits on the gas pressure, especially in large-volume lasers. The product of pressure (above atmospheric) and volume should not exceed 200 bar-liters to keep potential energy low enough; otherwise regular pressure-vessel testings become necessary for commercial systems. At high pulse repetition frequencies, which are capable of exciting standing acoustic waves in the laser gas (above about 1 kHz), the output power shows a decrease due to a disturbance of discharge homogeneity [80Fah, 85Mat, 86Fon]. High pulse energies are achieved by increasing the size of the pumped volume. The output energy for the excimer system XeCl is about 1000 J/(m3 bar). It appears as an optimal system, because the seed electrons from the pre-ionization do not disappear too quickly by attachment (cf. reactions No. 45 in Table 3.5.7 and No. 16 in Table 3.5.6). The eect of electron attachment must not be too strong, otherwise it will be almost impossible to achieve a homogeneous gas discharge, which is a necessity for homogeneous production of excited states. In order to have favorable competition with spontaneous decay, resonator roundtrip time must not be made much longer than a few ns. Therefore, attempts to increase the pulse energy aim at a larger cross section of the discharge [82Wat, 85Cha]. Preionization quality becomes an important prerequisite [81Her, 86Tay]. If pulsed high-power X-rays are used [86Osb], the transition from glow discharge to arc discharge sets the limits to output eciency (dened as output pulse energy divided through the electrical energy deposited in the discharge), which usually stays below 2 %. With the prepulse-technique a value for the eciency of 4.2 % has been reported [83Lon].

Landolt-Brnstein o New Series VIII/1B1

References for 3.5

285

References for 3.5


65Rap 65Vri 66Cal 66Len 67Rob 69Sha 70Boh 70McD Rapp, D., Englander-Golden, P.: J. Chem. Phys. 43 (1965) 1464. Vriens, L.: Physica 31 (1965) 395. Calvert, J.G., Pitts, J.N.: Photochemistry, New York: Wiley and Sons, 1966. Lengyel, B.A.: Introduction to Laser Physics, New York: Wiley and Sons, 1966. Robinson, E.J., Geltmann, S.: Phys. Rev. 153 (1967) 4. Shaper, M., Scheibner, H.: Beitr. Plasmaphys. 9 (1969) 45. Bohme, D.K., Adams, N.G., Muselman, M., Donkin, D.B., Ferguson, E.E.: J. Chem. Phys. 52 (1970) 5094. McDaniel, E.W., Ermak, V., Dalgarno, A., Ferguson, E.E., Friedman, L.: Ion-Molecule Reactions, New York: Wiley Interscience, 1970. Mandl, A.: Phys. Rev. A 3 (1971) 251. Peterson, L.R., Allen jr., J.E.: J. Chem. Phys. 56 (1972) 6068. Ho, P.W., Swingle, V.C., Rhodes, C.K.: Opt. Commun. 8 (1973) 128. Piper, L.G., Velazco, J.E., Setser, D.W.: J. Chem. Phys. 59 (1973) 3323. Wallace, S.C., Hodgson, R.T., Dreyfus, R.W.: Appl. Phys. Lett. 23 (1973) 672. Ishihara, T., Foster, T.C.: Phys. Rev. A 9 (1974) 2350. Eggarter, E.: J. Chem. Phys. 62 (1975) 833. Golde, M.F.: J. Mol. Spectrosc. 58 (1975) 261. Searles, S.K., Hart, G.A.: Appl. Phys. Lett. 27 (1975) 243. Dunning, T.H., Hay, P.J.: Appl. Phys. Lett. 28 (1976) 649. Ewing, J.J., Jacob, J.H., Magnano, J.A., Brown, H.A.: Appl. Phys. Lett. 28 (1976) 656. Hays, A.K., Homann, J.M., Tysone, G.C.: Chem. Phys. Lett. 39 (1976) 353. Homan, J.M., Hays, A.K., Tisone, G.C.: Appl. Phys. Lett. 28 (1976) 538. Jacob, J.H., Mangano, J.A.: Appl. Phys. Lett. 28 (1976) 724. Jacob, J.H., Mangano, J.A.: Appl. Phys. Lett. 29 (1976) 467. Lorents, D.C.: Physica C 82 (1976) 16. Murray, J.R., Swingle, J.C., Turner jr., C.E.: Appl. Phys. Lett. 28 (1976) 538. Murray, J.R., Powell, H.T.: Appl. Phys. Lett. 29 (1976) 252. Nakano, H.H., Hill, R.M., Lorents, D.C., Huestis, D.L., McCusker, M.V.: SRI (Stanford Research Institute) Report MP-76-99, 1976. Tellinghuisen, J., Homan, J.M., Tisone, G.C., Hays, A.K.: J. Chem. Phys. 64 (1976) 2484. Tellinghuisen, J., Hays, A.K., Homan, J.M., Tisone, G.C.: J. Chem. Phys. 65 (1976) 4473. Tellinghuisen, J., Tisone, G.C., Homan, J.M., Hays, A.K.: J. Chem. Phys. 64 (1976) 4796. Burnham, R., Searles, S.K.: J. Chem. Phys. 67 (1977) 5967. Chen, H., Center, R.E., Trainor, D.W., Fyfe, W.I.: Appl. Phys. Lett. 30 (1977) 99.

71Man 72Pet 73Hof 73Pip 73Wal 74Ish 75Egg 75Gol 75Sea 76Dun 76Ewi 76Hay 76Hof 76Jac1 76Jac2 76Lor 76Mur1 76Mur2 76Nak 76Tel1 76Tel2 76Tel3

77Bur 77Che

Landolt-Brnstein o New Series VIII/1B1

286 77Man 77Ric 77Shu 77Way 78Bar 78Lac 78Mic 78Res 78Rok

References for 3.5 Mangano, J.A., Jacob, J.H., Rokni, M., Hawryluk, A.M.: Appl. Phys. Lett. 31 (1977) 26. Rice, J.: 7th Winter Colloquium on High Power Visible Lasers, Park City, Utah, 1977. Shui, V.H.: Appl. Phys. Lett. 31 (1977) 50. Waynant, R.W.: Appl. Phys. Lett. 30 (1977) 234. Bardsley, J.N.: Appl. Phys. Lett. 32 (1978) 76. Lacina, W.B., Cohn, D.B.: Appl. Phys. Lett. 32 (1978) 106. Michels, H.H., Hobbs, R.H., Wright, L.A.: Int. J. Quantum Chem. 12 (1978) 257. Rescigno, T.N., Bender, C.F., McKoy, B.V.: Phys. Rev. A 17 (1978) 645. Rokni, M., Mangano, J.A., Jacob, J.H., Hsia, J.C.: IEEE J. Quantum Electron. 14 (1978) 464. Brau, Ch.A.: in: Excimer Lasers, Topics in Applied Physics, Vol. 30, Rhodes, Ch.K. (ed.), Berlin, Heidelberg, New York: Springer-Verlag, 1979, p. 87. Bardsley, J.N., Wadehra, M.: Chem. Phys. Lett. 72 (1980) 477. Borisov, V.M., Vysika F.I., Mamonov, S.G., Napartovich, A.P., Stepanov, Yu.Yu.: lo, Sov. J. Quantum Electron. (English Transl.) 10 (1980) 333. Fahlen, T.S.: IEEE J. Quantum Electron. 16 (1980) 1260. Nighan, W.L., Brown, R.T.: Appl. Phys. Lett. 36 (1980) 498. Demyanov, A.V., Dyatko, N.A., Kochetov, I.V., Napartovich, A.P., Pal, A.F., Pevgov, V.G., Perevoznov, A.F., Persiantsev, I.G., Starostin, A.N.: Sov. J. Plasma Phys. (English Transl.) 7 (1981) 768. Herziger, G., Wollermann-Windgasse, R., Banse, K.H.: Appl. Phys. 24 (1981) 267. Levin, L.A., Moody, S.E., Klostermann, E.L., Center, R.E., Ewing, J.J.: IEEE J. Quantum Electron. 17 (1981) 2282. Mizunami, T., Maeda, M., Uchino, O., Shimomura, O., Miyazoe, Y.: Rev. Laser Eng. (Reza Kenkyu) 9 (1981) 512. Chantry, P.J.: Applied Atomic Collision Physics Vol. 3, New York: Academic Press, 1982, p. 53. Corkum, P.B., Taylor, R.S.: IEEE J. Quantum Electron. 18 (1982) 1962. Maeda, M., Takahashi, T., Mizunami, T., Miyazoe, Y.: Jpn. J. Appl. Phys. Part 1 21 (1982) 1161. Shuaibov, A.K., Shevera, V.S., Gerts, S.Yu., Malinin, A.I.: Izv. Vyssh. Uchebn. Zaved. Fiz. 8 (1982) 121. Watanabe, S., Endoh, A.: Appl. Phys. Lett. 41 (1982) 799. Clodius, W.B., Stehman, R.M., Woo, S.B.: Phys. Rev. A 27 (1983) 333. Long, W.H., Plummer, M.J., Stappaerts, E.A.: Appl. Phys. Lett. 43 (1983) 735. Nighan, W.L., Nachshon, Y., Tittel, F.K., Wilson jr., W.L.: Appl. Phys. Lett. 42 (1983) 1006. Rhodes, Ch.K. (ed.): Excimer Lasers, Topics in Applied Physics, Vol. 30, 2nd ed., New York: Springer-Verlag, 1984. Szatmari, S., Schfer, F.P.: Appl. Phys. B 33 (1984) 219. a Champagne, L.F., Dudas, A.J., Feldman, B.J.: CLEO 85, Book of Abstracts, THQ3, 1985, p. 230.

79Bra

80Bar 80Bor 80Fah 80Nig 81Dem

81Her 81Lev 81Miz

82Cha 82Cor 82Mae 82Shu 82Wat 83Clo 83Lon 83Nig

84Rho 84Sza 85Cha

Landolt-Brnstein o New Series VIII/1B1

References for 3.5 85Ern 85Mat

287

Ernst, G.J, Nieuwenhuis, A.B.M., Abramski, K.M.: IEEE J. Quantum Electron. 21 (1985) 1127. Matera, M., Bua, R., Burlamacchi, P., Fini, L., Salimbeni, R.: Rev. Sci. Instrum. 56 (1985) 205. Eichler, H.J., Hamisch, H., Nagel, B., Schmid, W.: Laser 85 Optoelectronics in Engineering, Waidelich, W. (ed.), Berlin: Springer-Verlag, 1986, p. 15. Fontaine, B.L., Forestier, B.M., Sentis, M.L., Gevaudan, A.: CLEO 86, Book of Abstracts, WK29, 1986, p. 192. Osborne, M.R., Hutchinson, M H.R.: Appl. Phys. Lett. 49 (1986) 7. Pummer, H.: Laser 85 Optoelectronics in Engineering, Waidelich, W. (ed.), Berlin: Springer-Verlag, 1986, p. 3. Taylor, R.S.: Appl. Phys. B 41 (1986) 1. Cirkel, H.-J., Baumgartl, R., Bette, W., Friede, D., Mller, R.: Laser 87 Optoelecu tronics in Engineering, Waidelich, W. (ed.), Berlin: Springer-Verlag, 1987, p. 16. Klopotek, P., Brinkmann, U., Oesterlin, P., Mckenheim, W.: Laser 87 Optoelectronics u in Engineering, Waidelich, W. (ed.), Berlin: Springer-Verlag, 1987, p. 12. Radojevic, V., Kelly, H.P., Johnson, W.R.: Phys. Rev. A 38 (1987) 2117. Wang, W.C., Lee, L.C.: J. Phys. D 21 (1988) 675. Osborne, M.R., Wineld, R.J., Green, J.M.: J. Appl. Phys. 65 (1989) 5242. Klingenberg, H.H., Gekat, F.: Appl. Phys. B 54 (1991) 205.

86Eic 86Fon 86Osb 86Pum 86Tay 87Cir 87Klo 87Rad 88Wan 89Osb 91Kli

Landolt-Brnstein o New Series VIII/1B1

S-ar putea să vă placă și