Spectroscopic Study of Rotational Nonequilibrium in Supersonic Free Molecular Flows Hideo Mori, Ayumi Takasu, Kenji Niwa, Toshihiko Ishida and Tomohide Niimi Nagoya University, Furo-cho, Chikusa, Nagoya 464-8603, Japan Abstract. In highly rarefied gas flows, there appear nonequilibrium phenomena not only between translational and rotational energy modes but also in rotational mode. To analyze these highly rarefied gas flows, we established the experimental system for 2R+2 N2 -REMPI and applied it to the measurement of rotational temperature in a supersonic free molecular flow of nitrogen. The data in the Boltzmann plot, obtained from the measured REMPI spectra, cannot be fitted by one line but approximately by two lines, revealing the non-Boltzmann distribution of rotational energy in the ground state. In comparison with the Boltzmann distribution at the source condition, it is clarified that the rotational energy distribution at relatively high rotational level of J ≥ 13 is unchanged during the flow expansion, while there are rotational transitions from levels in J ≤ 12 to lower levels by the molecular collisions during the expansion. INTRODUCTION Equilibrium condition is established by a large number of intermolecular collisions. If there are enough intermolecular collisions in a system consisting of a large number of gas molecules, distributions of translational, vibrational, and rotational energy follow the Maxwell-Boltzmann distribution, and these distributions reveals the same temperature. However, if the number of intermolecular collisions is very low, there appear nonequilibrium phenomena among these energy modes, and these distributions reveals the different temperature. Moreover, deviation of rotational energy distribution from Boltzmann distribution occurs if the molecular collision number is too low to maintain the Boltzmann distribution. In this study, the resonantly enhanced Multiphoton ionization (REMPI) technique is applied to an analysis of a supersonic free molecular flow of nitrogen gas. In the 2R+2 N2 -REMPI technique[1, 2], nitrogen molecules are ionized by two steps, i.e., the first step from the ground state to the resonance state (two photons) and the second step from the resonance state to the ionization state (two photons). The nitrogen ions are detected as a signal and its spectrum depending on the wavelength of an irradiated laser beam is analyzed to measure the rotational energy distribution, or the rotational temperature. The experimental setup for 2R+2 N2 -REMPI is established, and is applied to a measurement of the rotational temperature in a supersonic free molecular nitrogen flow. From the Boltzmann plot of the REMPI spectra, the rotational energy distribution of nitrogen in the free molecular flow is analyzed, and the deviation of the rotational energy distribution from the Boltzmann distribution is discussed. ROTATIONAL ENERGY RELAXATION Now we consider the process of the rotational energy relaxation in supersonic free jets, by using the relaxation equation by Gallagher and Fenn[3]. The relaxation time τr is defined as DTrot Ttr − Trot = , Dt τr (1) where Trot and Ttr are the rotational and translational temperature, respectively. For a one-dimension steady flow, D/Dt is given as D/Dt = u · d/dx, using the flow velocity u and the distance x downstream from the nozzle exit. The CP663, Rarefied Gas Dynamics: 23rd International Symposium, edited by A. D. Ketsdever and E. P. Muntz © 2003 American Institute of Physics 0-7354-0124-1/03/$20.00 612 TABLE 1. Two-photon Hönl-London factors of nitrogen for the a1 Πg ← X 1 Σ+ g transition using linearly polarized light S(J ) Branch O P Q R S (∆J = −2) (∆J = −1) (∆J = 0) (∆J = 1) (∆J = 2) M(O)J (J − 2)/15(2J − 1) M(P)(J + 1)/30 M(Q)(2J + 1)/10(2J − 1)(2J + 3) M(R)J/30 M(S)(J + 1)(J + 3)/15(2J + 3) rotational collision number Zr is a number of collisions during the relaxation time τr and is defined as C̄ Zr ≡ τr . λ (2) C̄ and λ are the mean speed and the mean free path of molecules, respectively. If the isentropic flow is assumed, the relaxation equation can be written with the Mach number Ma as d(Trot /T0 ) −4Ddm 2 n0 π /γ (Trot /T0 ) 1 + (γ − 1)Ma 2 /2 − 1 = (3) γ /(γ −1) d(x/D) Zr Ma 1 + (γ − 1)Ma 2 /2 where D is the nozzle diameter. Eq. (3) can be solved numerically by the Runge-Kutta method with the empirical formula of the dependence of Ma on x at the center line of supersonic free jets proposed by Ashkenas and Sherman[4]. Figure 1 shows the relaxation in the free molecular flow calculated by solving Eq. (3) using the conditions of the study mentioned in the following chapters. The source pressure and the temperature are 1.2Torr (160Pa) and 293K, respectively, and the nozzle diameter D is 0.5mm. The rotational collision number Zr is assumed as 4.2, proposed by Miller and Andres[5]. In this case, at the ratio of the distance from the nozzle exit x to D is 10, translational temperature is derived as 17K, and the rotational temperature is 210K. From the result, we can expect that the nonequilibrium, or non-equipartition between the translational and the rotational energy, will occur in the condition, because of very low number density. THEORY OF REMPI SPECTRA The theory of 2R+2 N2 -REMPI spectra [1, 2, 6] is used to analyze the possibility of rotational temperature measurement of nitrogen. If the rotational energy distribution follows the Maxwell-Boltzmann distribution and the rotational temperature is Trot , the number of molecules NJ in the rotational level J is given by[7] NJ ∝ (2J + 1) exp(−Erot /kTrot ), (4) where k is Boltzmann’s constant and Erot is the rotational energy. Because each state with total angular momentum J has a (2J + 1)-fold degeneracy[7], the distribution is given by the product of the simple Boltzmann factor exp(−Erot /kTrot ) and the deneracy 2J + 1. When the rotational energy distribution follows Eq. (4), the rotational line intensity in 2R+2 N2 -REMPI spectra is given by IJ ,J = Ag(J )S(J , J ) exp(−Erot /kTrot ), (5) where A is the constant independent of the rotational quantum number J of the ground state and J of the resonance state, but including laser flux, number density, vibrational transition strength (Franck-Condon factor). g(J ) is the nuclear spin degeneracy depending on the parity of J and the spin Is of the nuclei[7]. For N2 (Is = 1), g(J ) takes 3 and 6 for odd and even J , respectively. S(J , J ) is the two-photon Hönl-London factor for the a1 Πg ← X 1 Σ+ g transition. In Table 1, the values of S(J , J ) for linearly polarized light[8, 9, 10] are listed (in this table, ∆J = J − J ). The M(O)–M(S) in Table 1 are the relative transition intensity factors of the branches given by the products of the electronic transition dipole moments. These factors depend not on J but on the kind of the branches. 613 Since the spectral line intensity I(J , J ) divided by g(J )S(J , J ) depends linearly on exp(Erot /kTrot ), as is easily derived from Eq. (5), the rotational temperature can be deduced from the REMPI spectra using the Boltzmann plot. Even the deviation of the rotational energy distribution from the Boltzmann distribution occurs, the rotational energy distribution itself can be derived from REMPI spectra, by multiplying I/gS by the degeneracy (2J + 1). Erot is a function of J and is given by[7] (6) Erot = hc Bv J(J + 1) − DvJ 2 (J + 1)2 + · · · , where h is Planck’s constant and c is the speed of light. Bv and Dv are the rotational constants in the vibrational level v, and the values are Bv = 1.9895781cm−1 and Dv = 5.74118 × 10−6cm−1 for nitrogen molecules at X 1 Σ+ g electronic state and v = 0[11]. EXPERIMENTAL APPARATUS Figure 2 shows the experimental apparatus that we constructed for measurements of REMPI spectra of a free molecular flow. Nitrogen gas is issued via a sonic nozzle, whose diameter D is 0.50mm, into a vacuum chamber. The vacuum chamber is evacuated by two turbo molecular pumps in parallel. For the stagnation pressure of 1.2Torr (160Pa) and the temperature of 293K, the pressure in the chamber is kept at 3.3 × 10−5Torr (4.4 × 10−3 Pa). The background temperature in the chamber is 293K, which is the room temperature of the laboratory. We use Nd-YAG pumped dye laser (Lambda Physik, SCANMATE OG 2E C-400) with Rhodamine 6G dye as the laser source, and the output is frequency-doubled by a BBO crystal with an auto-tracker. The energy, oscillation frequency, and duration time of the laser is 7mJ/pulse, 10Hz, and 7ns, respectively. The beam is focused through a quartz lens (f = 120mm) on the center line of the nitrogen free molecular flow. The ionized nitrogen molecules are detected by a secondary electron multiplier (Murata, CERATRON®). The ion signal is amplified by a current-input preamplifier and averaged by a boxcar integrator, and the intensity of the signal is stored to a personal computer. To collect the ions by the detector effectively, an anode plate (repeller) and two cathode plates with a hole (f = 10mm), which act as “lenses” for electric field, are placed in front of the detector. The signal intensity is integrated 100 times for each wavelength. The wavelength range is 283–284.1nm and the step of the scanning is 0.001nm. RESULTS AND DISCUSSIONS Rotational Energy Distribution Deduced from REMPI Figure 3 is a REMPI spectrum measured using the experimental apparatus shown in Fig. 2. The horizontal axis of Fig. 3 is wavelength of the laser and the vertical axis is the signal intensity normalized by the maximum. The scales drawn in the upper side of the figure indicate the spectral lines of nitrogen molecules, and the numbers attached to the scales indicate J of the spectral lines. Figure 4 is a Boltzmann plot using the spectra. The horizontal axis indicates the rotational energy of the ground level Erot divided by k. The vertical axis is logarithm of the signal intensity I divided by the statistical weight g and the transition probability S. The ratio of M factors in S shown in Table 1 are determined by the relative intensity of spectral lines with the same J but belonging to the different branch, and derived as M(P)/M(O) = 0.757 and M(R)/M(O) = 0.596. The solid line in this figure is given by a least-square fitting using the points, and the rotational temperature is deduced from the reciprocal of the slope as 306K. The translational temperature at the focal point is estimated as 17K assuming isentropic flow (see Fig. 1), and there is a difference among these temperatures. From the result, the nonequilibrium, or non-equipartition among the translational and the rotational energy exists. However, the deduced rotational temperature is slightly higher than the source temperature T0 = 293K. There seems to exist no (exothermic) chemical reaction in the plume of pure nitrogen gas, and no shockwave seems to be generated in such a highly rarefied flows; there are no causes to raise the temperature in the plume. In Fig. 4, the lower and the higher level look relatively highly populated, although the middle level looks less populated. There seems to be nonequilibrium in the rotational energy distribution, i.e., deviation from the Boltzmann distribution. If the Boltzmann plot is carried out with the rotational level of J ≤ 8, the slope is indicated by the solid 614 Temperature [K] 500 x/D = 10.0 Trot = 210K Ttr = 17K Ma = 9.0 13 3 n = 3.2×10 /cm 100 50 10 5 P0 = 1.2Torr T0 = 293K D = 0.5mm Translational Rotational (Zr=4.2) T=T0 1 0 10 FIGURE 1. 20 30 40 Axial Distance [ x/D ] 50 Results of Relaxation Equation Personal Computer Grating Controller YAG-pumped Dye Laser (Rhodamine 590) Averaged Sig. Trigger SHG Generator (BBO with Auto-tracker) Ref. Beam Sampler N2 Sig. Boxcar Integrator Photodiode Preamplifier Chamber Quartz Lens Secondary Electron Multiplier (CERATRON R ) Turbo molecular pump Rotary pump FIGURE 2. Experimental apparatus 615 S 1 R Q P 10 O 10 4 19 20 20 0 16 P0 = 1.2Torr T0 = 293K x/D = 10 0.5 0 283 283.5 Wavelength[nm] 0 3 4 5 6 7 -1 -2 8 J''=2 9 10 O-Branch P-Branch R-Branch 11 12 13 -3 14 15 16 200 400 17 18 19 Trot =306K -4 0 284 REMPI spectrum FIGURE 3. ln( I/gS) [a.u.] Intensity[a.u.] 1.5 600 800 Erot / k [K] FIGURE 4. Boltzmann plot 616 1000 1200 line of Fig. 5, and the deduced “temperature” is 165K. On the other hand, if the Boltzmann plot is carried out with the rotational level of J ≥ 8, the deduced “temperature” is 357K, as shown by the broken line. The temperature of 357K deduced from the Boltzmann plot at J ≥ 8 is higher than the background temperature of 293K. Therefore, the population distribution cannot be assumed as a mixture of the distribution of the cooled gas flow and that of the background gas. As a result, it should be thought that deviation from the Boltzmann distribution occurs in the plume itself produced with the experimental condition. Deviation of Rotational Energy Distribution from the Boltzmann Distribution In the experimental conditions in this study, the molecular collision number in the expanding flow seems to be very low because of low source pressure of P0 = 1.2Torr. As a result, the equilibrium cannot be maintained when the rotational energy is transfered to the translational energy along with the cooling of expanding flows, and deviation from the Boltzmann distribution occurs. Figure 6 is the deduced rotational energy distribution of molecules in the plume. The horizontal axis is J, and the vertical one indicates the population ratio, that is, the number of molecules NJ divided by ∑J NJ . The closed circles display the distribution that is deduced experimentally. The solid line indicates the Boltzmann distribution of Trot = 210K, which is the theoretical rotational temperature of nitrogen at x/D = 10 (see Fig. 1), and the broken line is that of the source temperature T0 = 293K. To calculate ∑J NJ and NJ / ∑J NJ of the experimetnal results, the NJ values at J < 2 and J > 19 are deduced by extrapolating the experimental results at 2 ≤ J ≤ 19, because they cannot be determined by the REMPI spectra, The NJ=0 and NJ=1 can be extrapolated by the Boltzmann plot by J ≤ 8 as shown in Fig. 5 and multiplying the deduced I/gS by (2J + 1). Using the Boltzmann plot for J ≥ 8, NJ at J > 19 can be deduced in the same manner. If the experimental distribution is compared with the Boltzmann distribution of T0 = 210K, the experimental population at x < 11 is lower than the theoretical one, while the experimental one at x > 11 is higher than the theoretical one. It is also worth noting that the population ratio at J ≥ 13 is very close to that of the source condition at Trot = 293K (the broken line of Fig. 6). This result reveals that there are rotational transitions from levels in J ≤ 12 to lower levels by the molecular collisions during the expansion, but no transitions from levels in J ≥ 13. It can be expected that the probability of rotational energy exchange at high J is lower than that at low J. CONCLUSION In this study, the rotational energy distribution in supersonic free molecular flow with very low number density is measured using 2R+2 N2 -REMPI method, and the nonequilibrium phenomena of a highly rarefied gas flows are analyzed. Following concluding remarks are obtained. 1. The nonequilibrium of rotational energy distribution, i.e., deviation from the Boltzmann distribution, is detected experimentally in supersonic free molecular flow. If the Boltzmann plot is carried out with the rotational level of J ≤ 8, the “temperature” is deduced as 165K, while the temperature deduced with the rotational level of J ≥ 8 is 357K, which is higher than the source temperature T0 = 293K. The rotational energy distribution cannot be assumed as a mixture of the energy distribution of the cooled gas flow and that of the background gas. 2. In our experimental condition, the population ratio at J < 11 is lower than that of the Boltzmann distribution at Trot = 210K, which is deduced using a rotational relaxation equation, while higher rotational energy level of J > 11 are overpopulated. Moreover, the population ratio at J ≥ 13 is very close to that of the source condition. From the result, it can be expected that the probability of rotational energy exchange at high J is lower than that at low J. ACKNOWLEDGMENTS The present work was supported by "Molecular Sensors for Aero-Thermodynamic Research (MOSAIC)", the Special Coordination Funds, and a grant-in-aid for Scientific Research (B) of Ministry of Education, Culture, Sports, Science and Technology. Hideo Mori also wishes to thank Japan Society for the Promotion of Science for a research fellowship. 617 3 4 ln( I/gS) [a.u.] 0 -1 5 6 7 8 J''=2 9 10 -2 -3 O-Branch P-Branch R-Branch 11 12 13 14 15 16 J ≤ 8 Trot =165K J ≥ 8 Trot =357K -4 0 200 400 600 800 Erot / k [K] 17 18 19 1000 1200 FIGURE 5. Boltzmann plot using two lines Experimental Theoretical (Trot=210K) Trot=293K NJ /ΣNJ 0.1 0.05 0 0 5 FIGURE 6. 10 J 15 Rotational energy distribution of molecules 618 20 REFERENCES 1. Mori, H., Ishida, T., Aoki, Y., and Niimi, T., “Spectroscopic Study of REMPI for Rotational Temperature Measurement in Highly Rarefied Gas Flows,” in Rarefied Gas Dynamics: 22nd International Symposium, edited by T. J. Bartel and M. A. Gallis, 2001, vol. 585 of AIP Conference Proceedings, pp. 956–963. 2. Dankert, C., and Nazari, B. K., “Spectroscopy Measurements in Rarefied Plumes in the Cryopumped Chamber STG,” in Rarefied Gas Dynamics: 22nd International Symposium, edited by T. J. Bartel and M. A. Gallis, 2001, vol. 585 of AIP Conference Proceedings, pp. 730–736. 3. Gallagher, R. J., and Fenn, J. B., J. Chem. Phys., 60, 3487–3491 (1974). 4. Ashkenas, H., and Sherman, F. S., “The Structure and Utilization of Supersonic Free Jets in Low Density Wind Tunnels,” in RAREFIED GAS DYNAMICS, 4th International Symposium on Rarefied Gas Dynamics, Academic Press, 1966, vol. 2, pp. 84–105. 5. Miller, D. R., and Andres, R. P., J. Chem. Phys., 46, 3418–3423 (1967). 6. Nazari, B. K., Beylich, A. E., and Dankert, C., “Rotational Temperature Measurements in the New DLR–High Vacuum Test Facility STG by Means of REMPI,” in RAREFIED GAS DYNAMICS, 21st International Symposium on Rarefied Gas Dynamics, 1999, vol. 1, pp. 583–590. 7. Herzberg, G., SPECTRA of DIATOMIC MOLECULES, Van Nostrand Reinhold, New York, 1950. 8. Bray, R. G., and Hochstrasser, R. M., Molecular Physics, 31, 1199–1211 (1976). 9. Halpern, J. B., Zacharias, H., and Wallenstein, R., J. Mol. Spectrosc., 79, 1–30 (1980). 10. Bruno, A. E., Schubert, U., Neusser, H. J., and Schlag, E. W., Chem. Phys. Lett., 131, 31–36 (1986). 11. Trickl, T., Proch, D., and Kompa, K. L., J. Mol. Spectrosc., 162, 184–229 (1993). 619
© Copyright 2025 Paperzz