Skip to main content
Log in

Experimental assessment of the Tenti S6 model for combustion-relevant gases and filtered Rayleigh scattering applications

  • Published:
Applied Physics B Aims and scope Submit manuscript

Abstract

Accurate descriptions of Rayleigh–Brillouin scattering (RBS) spectra for gas-phase species are important in light scattering measurement techniques including filtered Rayleigh scattering (FRS). The current manuscript targets evaluation of the well-known Tenti S6 model for calculating the RBS spectra for combustion-relevant species over a broad range of temperatures with relevance towards FRS applications in reacting flows. In this work, testing of the Tenti S6 model was performed by comparing measured FRS signals to synthetic FRS signals generated through the combination of the Tenti S6 model and an experimentally verified I2 absorption model. First, temperature-dependent FRS signals were measured for a number of individual gases including Ar, N2, O2, CH4, H2, CO, and CO2 from 300 to 1400 K. Comparisons between the measurements and synthetic FRS signals show excellent agreement (< 4% average difference) over the full temperature range. For pure CO2, rotational Raman scattering effects must be taken into account when comparing measured and synthetic FRS signals. FRS measurements in binary mixtures were performed to assess the commonly used (but not verified) assumption that the total FRS signal from a mixture can be treated as the mole fraction-weighted average of the FRS signals from each component. Measured FRS signals in mixtures with large variations in both molecular weight and Rayleigh scattering cross section show a linear relationship with constituent mole fraction, indicating that this assumption is valid within the kinetic regime. Finally, FRS measurements were performed in near-adiabatic H2/air and CH4/air flames. Comparisons between measured and synthetic FRS signals show excellent agreement over a broad range of equivalence ratios (\(\phi\)), which includes a temperature range of 1100 < T(K) < 2400 and large relative changes in species mole fractions. Overall, the results indicate that the predicted RBS lineshapes calculated using the Tenti S6 model are sufficiently accurate in the context of FRS measurements for the species and temperatures evaluated.

This is a preview of subscription content, log in via an institution to check access.

Access this article

Price excludes VAT (USA)
Tax calculation will be finalised during checkout.

Instant access to the full article PDF.

Fig. 1
Fig. 2
Fig. 3
Fig. 4
Fig. 5
Fig. 6
Fig. 7
Fig. 8
Fig. 9
Fig. 10

Similar content being viewed by others

Notes

  1. In Ref. [41]. the temperature dependence of the Rayleigh scattering cross sections (σi) were reported at 355 and 266 nm. No temperature dependence was reported at 532 nm because the temperature variation of σI over the temperature range of 300–1500 K was less than the uncertainty (2%) of the measurements.

  2. The trace scattering cross section, \({\sigma }_{i}^{t}\) is calculated as \({{\upsigma }}_{\text{i}}-\left(4{\rho }_{i}/3-4{\rho }_{i}\right)\), where \({\rho }_{i}\) is the depolarization ratio of species i.

References

  1. J.W. Strutt, On the transmission of light through an atmosphere containing small particles in suspension, and on the origin of the blue of the sky. The London, Edinburgh, and Dublin Philosophical Magazine and. J. Sci. 47, 375–384 (1899)

    Google Scholar 

  2. R.W. Boyd, Nonlinear Optics, 3rd ed. (Academic Press, Cambridge, 2008)

    Google Scholar 

  3. A.T. Young, Rayleigh scattering. Appl. Opt. 20, 533–535 (1981)

    Article  ADS  Google Scholar 

  4. C.D. Boley, R.C. Desai, G. Tenti, Kinetic models and Brillouin scattering in a molecular gas. Can. J. Phys. 50, 2158–2173 (1972)

    Article  ADS  Google Scholar 

  5. G. Tenti, C.D. Boley, R.C. Desai, On the kinetic model description of Rayleigh–Brillouin scattering from molecular gases. Can. J. Phys. 52, 285–290 (1974)

    Article  ADS  Google Scholar 

  6. C.S. Wang Chang, G.E. Uhlenbeck, J. de Boer, The heat conductivity and viscosity of poly-atomic gases. in Studies in Statistical Mechanics, ed. by J. deBoer, G.E. Uhlenbeck (Wiley, New York, 1964)

    MATH  Google Scholar 

  7. A. Stoffelen, G.J. Marseille, F. Bouttier, D. Vasiljevic, S. De Haan, C. Cardinali, ADM-aeolus doppler wind lidar observing system simulation experiment. Q. J. R. Meteorol. Soc. 132, 1927–1947 (2006)

    Article  ADS  Google Scholar 

  8. B. Witschas, C. Lemmerz, O. Reitebuch, Daytime measurements of atmospheric temperature profiles (2–15 km) by lidar utilizing Rayleigh–Brillouin scatterin. Opt. Lett. 39, 1972–1975 (2014)

    Article  ADS  Google Scholar 

  9. M.O. Vieitez, E.J. van Duijn, W. Ubachs, B. Witschas, A. Meijer, A.S. de Wijn, N.J. Dam, W. van de Water, Coherent and spontaneous Rayleigh–Brillouin scattering in atomic and molecular gases and gas mixtures. Phys. Rev. A 82, 043836 (2010)

    Article  ADS  Google Scholar 

  10. Y. Ma, H. Li, Z. Gu, W. Ubachs, Y. Yu, J. Huang, B. Zhou, Y. Wang, K. Liang, Analysis of Rayleigh–Brillouin spectral profiles and Brillouin shifts in nitrogen gas and ai. Opt. Express 22, 2092–2104 (2014)

    Article  ADS  Google Scholar 

  11. B. Witschas, M.O. Vieitez, E.-J. van Duijn, O. Reitebuch, W. van de Water, W. Ubachs, Spontaneous Rayleigh–Brillouin scattering of ultraviolet light in nitrogen, dry air, and moist ai. Appl. Opt. 49, 4217–4227 (2010)

    Article  ADS  Google Scholar 

  12. Z.Y. Gu, W. Ubachs, W. van de Water, Rayleigh–Brillouin scattering of carbon dioxide. Opt. Lett. 39, 3301–3304 (2014)

    Article  ADS  Google Scholar 

  13. Z. Gu, W. Ubachs, A systematic study of Rayleigh–Brillouin scattering in air, N2, and O2 gases. J. Chem. Phys. 141, 104320 (2014)

    Article  ADS  Google Scholar 

  14. Z. Gu, B. Witschas, W. van de Water, W. Ubachs, Rayleigh–Brillouin scattering profiles of air at different temperatures and pressure. Appl. Opt. 52, 4640–4651 (2013)

    Article  ADS  Google Scholar 

  15. J.N. Forkey, W.R. Lempert, R.B. Miles, Accuracy limits for planar measurements of flow field velocity, temperature and pressure using Filtered Rayleigh Scatterin. Exp. Fluids 24, 151–162 (1998)

    Article  Google Scholar 

  16. R.B. Miles, W.R. Lempert, J.N. Forkey, Laser Rayleigh scattering. Meas. Sci. Technol. 12, R33 (2001)

    Article  ADS  Google Scholar 

  17. M. Boguszko, G.S. Elliott, On the use of filtered Rayleigh scattering for measurements in compressible flows and thermal field. Exp. Fluids 38, 33–49 (2005)

    Article  Google Scholar 

  18. U. Doll, G. Stockhausen, C. Willert, Pressure, temperature, and three-component velocity fields by filtered Rayleigh scattering velocimetr. Opt. Lett. 42, 3773–3776 (2017)

    Article  ADS  Google Scholar 

  19. U. Doll, G. Stockhausen, C. Willert, Endoscopic filtered Rayleigh scattering for the analysis of ducted gas flows. Exp. Fluids 55, 1690 (2014)

    Article  Google Scholar 

  20. F. Benhassen, M.D. Polanka, M.F. Reeder, Trajectory measurements of a horizontally oriented buoyant jet in a coflow using filtered rayleigh scattering. J. Aerosp. Eng. 30, 04016067 (2017)

    Article  Google Scholar 

  21. J. Brübach, J. Zetterberg, A. Omrane, Z.S. Li, M. Aldén, A. Dreizler, Determination of surface normal temperature gradients using thermographic phosphors and filtered Rayleigh scatterin. Appl. Phys. B 84, 537–541 (2006)

    Article  ADS  Google Scholar 

  22. P.M. Allison, T.A. McManus, J.A. Sutton, Quantitative fuel vapor/air mixing imaging in droplet/gas regions of an evaporating spray flow using filtered Rayleigh scattering. Opt. Lett. 41, 1074–1077 (2016)

    Article  ADS  Google Scholar 

  23. D. Hoffman, K.U. Münch, A. Leipertz, Two-dimensional temperature determination in sooting flames by filtered Rayleigh scattering. Opt. Lett. 21, 525–527 (1996)

    Article  ADS  Google Scholar 

  24. G. Elliott, N. Glumac, C. Carter, Molecular filtered Rayleigh scattering applied to combustion. Meas. Sci. Technol. 12, 452 (2001)

    Article  ADS  Google Scholar 

  25. D. Most, A. Leipertz, Simultaneous two-dimensional flow velocity and gas temperature measurements by use of a combined particle image velocimetry and filtered Rayleigh scattering technique. Appl. Opt. 40, 5379–5387 (2001)

    Article  ADS  Google Scholar 

  26. D. Most, F. Dinkelacker, A. Leipertz, Direct determination of the turbulent flux by simultaneous application of filtered rayleigh scattering thermometry and particle image velocimetry. Proc. Comb. Inst. 29, 2669–2677 (2002)

    Article  Google Scholar 

  27. A.P. Yalin, Y.Z. Ionikh, R.B. Miles, Gas temperature measurements in weakly ionized glow discharges with filtered Rayleigh scattering. Appl. Opt. 41, 3753–3762 (2002)

    Article  ADS  Google Scholar 

  28. S.P. Kearney, R.W. Schefer, S.J. Beresh, T.W. Grasser, Temperature imaging in nonpremixed flames by joint filtered Rayleigh and Raman scattering. Appl. Opt. 44, 1548–1558 (2005)

    Article  ADS  Google Scholar 

  29. J.R. Bonatto, W. Marques Jr., Kinetic model analysis of light scattering in binary mixtures of monatomic ideal gases. J. Stat. Mech. 2005, P09014–P09014 (2005)

    Article  Google Scholar 

  30. A.S. Fernandes, W. Marques, Sound propagation in binary gas mixtures from a kinetic model of the Boltzmann equation. Phys. A 332, 29–46 (2004)

    Article  Google Scholar 

  31. W. Marques, Coherent Rayleigh–Brillouin scattering in binary gas mixtures. J. Stat. Mech. 2007, 03013–03013 (2007)

    Article  Google Scholar 

  32. L. Letamendia, Light-scattering studies of moderately dense gas mixtures: Hydrodynamic regime. Phys. Rev. A Gen. Phys. 24, 1574–1590 (1981)

    Article  ADS  Google Scholar 

  33. L. Letamendia, P. Joubert, J.P. Chabrat, J. Rouch, C. Vaucamps, C.D. Boley, S. Yip, S.H. Chen, Light-scattering studies of moderately dense gases. II. Nonhydrodynamic regime. Phys. Rev. A 25, 481–488 (1982)

    Article  ADS  Google Scholar 

  34. J.N. Forkey, W.R. Lempert, R.B. Miles, Corrected and calibrated I 2 absorption model at frequency-doubled Nd:YAG laser wavelengths. Appl. Opt. 36, 6729–6738 (1997)

    Article  ADS  Google Scholar 

  35. T.L. Labus, E.P. Symons, Experimental investigation of an axisymmetric free jet with an initially uniform velocity profile, NASA Technical Report NASA-TN-D-6783, E-6801 (1972)

  36. J. Gauntner, P. Hrycak, D. Lee, J. Livingood, Experimental flow characteristics of a single turbulent jet impinging on a flat plate, NASA Technical Report NASA-TN D-5690 (1970)

  37. M.W. Thring, M.P. Newby, Combustion length of enclosed turbulent jet flames. Sym. (Int.) Combust. 4, 789–796 (1953)

    Article  Google Scholar 

  38. K.M. Tacina, W.J. Dahm, Effects of heat release on turbulent shear flows. Part 1. A general equivalence principle for non-buoyant flows and its application to turbulent jet flames. J. Fluid Mech. 415, 23–44 (2000)

    Article  ADS  MathSciNet  Google Scholar 

  39. R.D. Hancock, K.E. Bertagnolli, R.P. Lucht, Nitrogen and hydrogen CARS temperature measurements in a hydrogen/air flame using a near-adiabatic flat-flame burner. Combust. Flame 109, 323–331 (1997)

    Article  Google Scholar 

  40. M.J. Papageorge, C. Arndt, F. Fuest, W. Meier, J.A. Sutton, High-speed mixture fraction and temperature imaging of pulsed, turbulent fuel jets auto-igniting in high-temperature, vitiated co-flows. Exp. Fluids 55, 1763 (2014)

    Article  Google Scholar 

  41. J.A. Sutton, J.F. Driscoll, Rayleigh scattering cross sections of combustion species at 266, 355, and 532 nm for thermometry application. Opt. Lett. 29, 2620–2622 (2004)

    Article  ADS  Google Scholar 

  42. C. Carter, Laser-based Rayleigh and Mie scattering methods (Wiley, New York, 1996), pp. 1078–1093

    Google Scholar 

  43. R.L. McKenzie, Measurement capabilities of planar Doppler velocimetry using pulsed lasers. Appl. Opt. 35, 948–964 (1996)

    Article  ADS  Google Scholar 

  44. R. Patton, J. Sutton, Seed laser power effects on the spectral purity of Q-switched Nd:YAG lasers and the implications for filtered rayleigh scattering measurements. Appl. Phys. B Lasers Opt. 111, 457–468 (2013)

    Article  ADS  Google Scholar 

  45. J.A. Sutton, R.A. Patton, Improvements in filtered Rayleigh scattering measurements using Fabry–Perot etalons for spectral filtering of pulsed, 532-nm Nd:YAG output. Appl. Phys. B 116, 681–698 (2014)

    Article  ADS  Google Scholar 

  46. P. Linstrom, W. Mallard, Nist standard reference database number 69 (2003), 2018

  47. B.J. McBride, S. Gordon, M.A. Reno, Coefficients for calculating thermodynamic and transport properties of individual species, NASA Technical Report NASA-TM-4513, E-7981 (1993)

  48. X. Pan, M.N. Shneider, R.B. Miles, Coherent Rayleigh–Brillouin scattering in molecular gases. Phys. Rev. A 69, 033814 (2004)

    Article  ADS  Google Scholar 

  49. J.D. Lambert, Vibrational and rotational relaxation in gases (Oxford University Press, Oxford, 1977)

    Google Scholar 

  50. H.E. Bass, J.R. Olson, R.C. Amme, Vibrational relaxation in H2O vapor in the temperature range 373–946 K. J. Acoust. Soc. Am. 56, 1455–1460 (1974)

    Article  ADS  Google Scholar 

  51. R. Kung, R. Center, High temperature vibrational relaxation of H2O by H2O, He, Ar, and N2. J. Chem. Phys. 62, 2187–2194 (1975)

    Article  ADS  Google Scholar 

  52. Q. Lao, P. Schoen, B. Chu, Rayleigh–Brillouin scattering of gases with internal relaxation. J. Chem. Phys. 64, 3547–3555 (1976)

    Article  ADS  Google Scholar 

  53. A. Meijer, A. de Wijn, M. Peters, N. Dam, W. van de Water, Coherent Rayleigh–Brillouin scattering measurements of bulk viscosity of polar and nonpolar gases, and kinetic theory. J. Chem. Phys. 133, 164315 (2010)

    Article  ADS  Google Scholar 

  54. M.S. Cramer, Numerical estimates for the bulk viscosity of ideal gases. Phys. Fluids 24, 066102 (2012)

    Article  ADS  Google Scholar 

  55. S.P. Kearney, S.J. Beresh, T.W. Grasser, R.W. Schefer, P.E. Schrader, R.L. Farrow, A filtered rayleigh scattering apparatus for gas-phase and combustion temperature imaging. Paper AIAA 2003 – 584, 41st Aerospace Sciences Meeting, Reno, NV, January 2003

  56. G.K. Wertheim, M.A. Butler, K. West, D.N.E. Buchanan, Determination of the Gaussian and Lorentzian content of experimental line shapes. Rev. Sci. Instrum. 45(11), 1369–1371 (1974)

    Article  ADS  Google Scholar 

  57. T. Ida, M. Ando, H. Toraya, Extended pseudo-Voigt function for approximating the Voigt profile. J. Appl. Crystallogr. 33(6), 1311–1316 (2000)

    Article  Google Scholar 

  58. C.M. Penney, R.L. St. Peters, M. Lapp, Absolute rotational Raman cross sections for N2, O2, and CO2. J. Opt. Soc. Am. 64(5), 712–716 (1974)

    Article  ADS  Google Scholar 

  59. A. Weber, in The Raman Effect, vol 2: Applications, ed. by A. Anderson (Dekker, New York, 1973) (Ch. 9)

    Google Scholar 

  60. G. Herzberg, Molecular Spectra and Molecular Structure I. Spectra of Diatomic Molecules (Van Nostrand, Princeton, 1950)

    Google Scholar 

  61. M.P. Bogaard, A.D. Buckingham, R.K. Pierens, A.H. White, Rayleigh scattering depolarization ratio and molecular polarizability anisotropy for gases. J. Chem. Soc. Faraday Trans. 1 Phys. Chem. Condensed Phases 74, 3008–3015 (1978)

    Google Scholar 

  62. K.S. Jammu, G.E. St. John, H.L. Welsh, Pressure broadening of the rotational raman lines of some simple gases. Can. J. Phys. 44(4), 797–814 (1966)

    Article  ADS  Google Scholar 

  63. J.W. Gallagher, R.D. Johnson, in NIST Chemistry WebBook, NIST Standard Reference Database Number 69, ed. by P.J. Linstrom, W.G. Mallard. Constants of Diatomics Molecules (National Institute of Standards and Technology, Gaithersburg). https://doi.org/10.18434/T4D303. Accessed 6 Nov 2018

    Chapter  Google Scholar 

  64. R. Kee, F. Rupley, J. Miller, M. Coltrin, J. Grcar, E. Meeks, H. Moffat, A. Lutz, G. Dixon-Lewis, M. Smooke, CHEMKIN Collection, Release 3.6 (Reaction Design, Inc, San Diego, 2000)

    Google Scholar 

  65. G. Prangsma, A. Alberga, J. Beenakker, Ultrasonic determination of the volume viscosity of N2, CO, CH4 and CD4 between 77 and 300 K. Physica 64, 278–288 (1973)

    Article  ADS  Google Scholar 

  66. B. Annis, A. Malinauskas, Temperature dependence of rotational collision numbers from thermal transpiration. J. Chem. Phys. 54, 4763–4768 (1971)

    Article  ADS  Google Scholar 

  67. T.G. Winter, G.L. Hill, High-temperature ultrasonic measurements of rotational relaxation in hydrogen, deuterium, nitrogen, and oxygen. J. Acoust. Soc. Am. 42, 848–858 (1967)

    Article  ADS  Google Scholar 

  68. R. Healy, T. Storvick, Rotational collision number and Eucken factors from thermal transpiration measurements. J. Chem. Phys. 50, 1419–1427 (1969)

    Article  ADS  Google Scholar 

  69. M. Camac, Avco Everett Research Laboratory. Research Report 172 (1963)

  70. R.J. Gallagher, J.B. Fenn, Rotational relaxation of molecular hydrogen. J. Chem. Phys. 60, 3492–3499 (1974)

    Article  ADS  Google Scholar 

  71. A.D. Gupta, T. Storvick, Analysis of the heat conductivity data for polar and nonpolar gases using thermal transpiration measurements. J. Chem. Phys. 52, 742–749 (1970)

    Article  ADS  Google Scholar 

  72. J. Tao, G. Ganzi, S. Sandler, Determination of thermal transport properties from thermal transpiration measurements. II. J. Chem. Phys. 56, 3789–3793 (1972)

    Article  ADS  Google Scholar 

  73. J. Tao, W. Revelt, S. Sandler, Determination of thermal transport properties from thermal transpiration measurements. III. Polar gases. J. Chem. Phys. 60, 4475–4482 (1974)

    Article  ADS  Google Scholar 

  74. M.J. Assael, W.A. Wakeham, Thermal conductivity of four polyatomic gases. J. Chem. Soc. Faraday Trans. 1 Phys. Chem. Condensed Phases 77, 697–707 (1981)

    Google Scholar 

  75. G. Hill, T. Winter, Effect of temperature on the rotational and vibrational relaxation times of some hydrocarbons. J. Chem. Phys. 49, 440–444 (1968)

    Article  ADS  Google Scholar 

  76. P. Kistemaker, M. Hanna, A. Tom, A. De Vries, Rotational relaxation in mixtures of methane with helium, argon and xenon. Physica 60, 459–471 (1972)

    Article  ADS  Google Scholar 

  77. R. Holmes, G. Jones, N. Pusat, Combined viscothermal and thermal relaxation in polyatomic gases. Trans. Faraday Soc. 60, 1220–1229 (1964)

    Article  Google Scholar 

  78. A. Malinauskas, Thermal transpiration. rotational relaxation numbers for nitrogen and carbon dioxide. J. Chem. Phys. 44, 1196–1202 (1966)

    Article  ADS  Google Scholar 

  79. A. Tip, J. Los, A. De Vries, Rotational relaxation numbers from thermal transpiration measurements. Physica 35, 489–498 (1967)

    Article  ADS  Google Scholar 

  80. C. O’Neal Jr., R.S. Brokaw, Relation between thermal conductivity and viscosity for nonpolar gases. II. Rotational relaxation of polyatomic molecules. Phys. Fluids 6, 1675–1682 (1963)

    Article  ADS  Google Scholar 

  81. E. Mason, Molecular relaxation times from thermal transpiration measurements. J. Chem. Phys. 39, 522–526 (1963)

    Article  ADS  Google Scholar 

  82. R.G. Keeton, H. Bass, Vibrational and rotational relaxation of water vapor by water vapor, nitrogen, and argon at 500 K. J. Acoust. Soc. Am. 60, 78–82 (1976)

    Article  ADS  Google Scholar 

  83. H. Roesler, K.F. Sahm, Vibrational and rotational relaxation in water vapor. J. Acoust. Soc. Am. 37, 386–387 (1965)

    Article  ADS  Google Scholar 

Download references

Funding

This work was partially funded by the National Science Foundation (CBET-1055960) and Air Force Office of Scientific Research (FA9550-16-1-0366).

Author information

Authors and Affiliations

Authors

Corresponding author

Correspondence to Thomas A. Mcmanus.

Appendices

Appendix 1: calculation of rotational Raman scattering spectra and FRS signal contributions

If molecules are free to rotate, “Rayleigh scattering” consists of rotational Raman scattering as well as the Placzek trace scattering. Rotational Raman scattering comprised of an un-shifted Q-branch in addition to spectrally shifted Stokes and anti-Stokes branches. The Q-branch corresponds to no change of the rotational state (\(\Delta\)J = 0), while the Stokes and anti-Stokes branches follow selection rules of \(\Delta\)J = ± 2. The J → J + 2 transitions are the Stokes lines, while the J → J − 2 transitions are the anti-Stokes lines. The Raman shifts for the Stokes and anti-Stokes lines can be approximated as

$$\Delta {\nu _{J \to J^{\prime}}}= - \left( {4{B_o} - 6{D_o}} \right)(J+3/2)+8{D_o}{\left( {J+3/2} \right)^3},$$
(13)

and

$$\Delta {\nu _{J \to J^{\prime}}}=\left( {4{B_o} - 6{D_o}} \right)(J - 1/2) - 8{D_o}{\left( {J - 1/2} \right)^3},$$
(14)

respectively, where Bo is the rotational constant for the lowest vibrational level and Do is the centrifugal distortion coefficient.

The intensity of a single Stokes or anti-Stokes-shifted rotational Raman scattering line is given by

$${I_{J \to {J^\prime }}}=C{I_o}{n_i}{F_J}\sigma _{i}^{{J \to {J^\prime }}}.$$
(15)

If the gas is in thermal equilibrium at temperature T, the fraction of molecules in state J can be expressed as

$${F_J}={g_J}\left( {2J+1} \right)\exp \left( { - \frac{{{E_J}}}{{{k_B}T}}} \right)/Q,$$
(16)

where gJ is a nuclear degeneracy factor (dependent on nuclear spin), EJ is the rotational energy, and Q is the rotational partition function, determined by the normalization satisfying \(\mathop \sum \limits_{{J=0}}^{\infty } {F_J}=1\). The rotational energy can be accurately approximated by

$${E_J}=hc[{B_o}J\left( {J+1} \right) - {D_o}{J^2}\left( {J+1{)^2}} \right],$$
(17)

where h is Planck’s constant. Finally, the differential rotational Raman scattering cross section, \(\sigma _{i}^{{J \to J\prime }}\)is expressed as

$$\sigma _{i}^{{J \to {J^\prime }}}=\frac{{64{\pi ^4}}}{{45}}{b_{J \to J^{\prime}}}{\left( {{\nu _o} - \Delta {\nu _{J \to {J^\prime }}}} \right)^4}{\gamma ^2},$$
(18)

where \({b_{J \to J\prime }}\)is a Placzek–Teller coefficient and \(\gamma\) is the anisotropy of the molecular polarizability tensor. The Placzek–Teller coefficients for the Stokes and anti-Stokes lines are

$${b_{J \to J+2}}=\frac{{3\left( {J+1} \right)\left( {J+2} \right)}}{{2\left( {2J+1} \right)\left( {2J - 3} \right)}},$$
(19)

and

$${b_{J \to J - 2}}=\frac{{3J\left( {J - 1} \right)}}{{2\left( {2J+1} \right)\left( {2J - 1} \right)}},$$
(20)

respectively.

For each rotational line, Doppler and collisional broadening mechanisms lead to an intensity distribution described by a Voigt profile, which is the convolution of Gaussian (G) and Lorentzian (L) profiles. For the current work, the Voigt profile is approximated using a pseudo-Voigt profile [56]

$$\begin{aligned} & V_{{\text{p}}}^{{J \to {J^\prime }}}\left( {{\nu _{\text{r}}},\Delta {\nu _{J \to {J^\prime }}},\Delta {\nu _{\text{T}}}} \right) \\ & \quad =\eta L({\nu _{\text{r}}},\Delta {\nu _{J \to {J^\prime }}},\Delta {\nu _{\text{T}}})+(1 - \eta )G({\nu _{\text{r}}},\Delta \nu J \to {J^\prime }),\Delta {\nu _{\text{T}}}), \\ \end{aligned}$$
(21)

where the Gaussian and Lorentzian profiles are calculated as

$$\begin{aligned} & G\left( {{\nu _{\text{r}}},\Delta {\nu _{J \to {J^\prime }}},\Delta {\nu _{\text{T}}}} \right) \\ & \quad =\frac{{2\sqrt {\ln (2)} }}{{\Delta {\nu _{\text{T}}}\sqrt \pi }}\exp \left[ { - 4\sqrt {\ln (2)} \frac{{{{({\nu _r} - \Delta {\nu _{J \to {J^\prime }}})}^2}}}{{\Delta {\nu _{\text{T}}}^{2}}}} \right] \\ \end{aligned},$$
(22)

and

$$L\left( {{\nu _{\text{r}}},\Delta {\nu _{J \to {J^\prime }}},\Delta {\nu _{\text{T}}}} \right)=\frac{{\Delta {\nu _{\text{T}}}}}{{2\pi }}\frac{1}{{{{\left( {{\nu _{\text{r}}} - \Delta {\nu _{J \to {J^\prime }}}} \right)}^2}+{{\left( {\Delta {\nu _{\text{T}}}/2} \right)}^2}}},$$
(23)

and η is a function of the total full width at half maximum (FWHM) parameter, \(\Delta {\nu _{\text{T}}}\). An accurate formula for η is [57]

$$\eta =1.36603\left( {\Delta {\nu _{\text{L}}}/\Delta {\nu _{\text{T}}}} \right) - 0.47719{\left( {\Delta {\nu _{\text{L}}}/\Delta {\nu _{\text{T}}}} \right)^2}+0.11116{\left( {\Delta {\nu _{\text{L}}}/\Delta {\nu _{\text{T}}}} \right)^3},$$
(24)

where \(\Delta v_{L}\) is the Lorentzian FWHM parameter (due to collisional broadening) and the total FWHM parameter is given by

$$\Delta {\nu _{\text{T}}}=\left[ {\Delta \nu _{{\text{D}}}^{5}+2.69269\Delta \nu _{{\text{D}}}^{4}\Delta {\nu _{\text{L}}}+2.42843\Delta \nu _{{\text{D}}}^{3}\Delta \nu _{{\text{L}}}^{2}+4.47163\Delta \nu _{{\text{D}}}^{2}\Delta \nu _{{\text{L}}}^{3}+0.07842\Delta {\nu _{\text{D}}}\Delta \nu _{{\text{L}}}^{4}+\Delta \nu _{{\text{L}}}^{5}} \right],$$
(25)

where \(\Delta {\nu _{\text{D}}}\) is the Gaussian FWHM parameter (due to thermal broadening). The appropriate expressions for the FWHM of the Doppler (Gaussian) and collisional (Lorentzian) broadening profiles are

$$\Delta {\nu _{\text{D}}}=7.16 \times {10^{ - 6}}\sqrt {\frac{T}{{{M_{\text{w}}}}}\frac{{\left( {2\nu _{o}^{2}+2{\nu _o}\Delta {\nu _{J \to {J^\prime }}}} \right)}}{2}+{{\left( {\Delta \nu _{{J \to {J^\prime }}}^{2}} \right)}^2}},$$
(26)

and

$$\Delta {\nu _{\text{L}}}=P\sqrt {\frac{{{T_{{\text{ref}}}}}}{T}} \left( {AJ+B} \right),$$
(27)

where P is the pressure in atmospheres, Tref is a reference temperature which is 300 K in the current work, and A and B are the empirical constants describing the rotational level-dependent collisional broadening. Both the Gaussian and Lorentzian profiles (Eqs. 22 and 23) are normalized such that

$$\mathop \int \limits_{\nu }^{{}} G\left( {\nu ,\Delta {\nu _{\text{T}}}} \right)=\mathop \int \limits_{\nu }^{{}} L\left( {\nu ,\Delta {\nu _{\text{T}}}} \right)=1.$$
(28)

Using the above expressions, the spectral lineshape for each individual Stokes and anti-Stokes rotational line is defined as

$$\mathcal{R}_{i}^{{J \to {J^\prime }}}\left( {{\nu _{\text{r}}},\Delta {\nu _{J \to {J^\prime }}}} \right)= V_{{\text{p}}}^{{J \to {J^\prime }}},$$
(29)

and a complete Stokes/anti-Stokes spectrum for species i is generated by summing each individual rotational line as

$${S_{{\text{S}}/{\text{AS}}}}\left( \nu \right)=\mathop \sum \limits_{{J=0}}^{\infty } {F_J}\sigma _{i}^{{J \to {J^\prime }}}\mathcal{R}_{i}^{{J \to {J^\prime }}}\left( {{\nu _r},\Delta {\nu _{J \to {J^\prime }}}} \right).$$
(30)

Using the results of Kattawar et al. [58] and Miles et al. [16], the relative magnitude of the intensity (differential scattering cross section) of the Q-branch to the S/AS components of the rotational Raman scattering is 1/3 or

$${I_{\text{Q}}}={\text{C}}{{\text{I}}_{\text{o}}}{n_i}\sigma _{i}^{Q}=\mathop \sum \limits_{{J=0}}^{\infty } {I_{J \to {J^\prime }}}/3=\frac{C}{3}{I_{\text{o}}}{n_i}\mathop \sum \limits_{{J=0}}^{\infty } {F_J}\sigma _{i}^{{J \to {J^\prime }}}.$$
(31)

The lineshape of the Q-branch is described by a Voigt profile just as the Stokes and anti-Stokes lines and the same procedure outlined above using Eqs. (21)–(29) is followed with the exception that \(\Delta {\nu _{J \to {J^\prime }}}=0\) for the Q-branch:

$$V_{{\text{p}}}^{{\text{Q}}}\left( {{\nu _{\text{r}}},\Delta {\nu _{\text{T}}}} \right)=\eta L\left( {{\nu _{\text{r}}},{\nu _{\text{T}}}} \right)+\left( {1 - \eta } \right)G\left( {{\nu _{\text{r}}},\Delta {\nu _{\text{T}}}} \right),$$
(32)
$$G\left( {{\nu _{\text{r}}},\Delta {\nu _{\text{T}}}} \right)=\frac{{2\sqrt {\ln \left( 2 \right)} }}{{\Delta {\nu _{\text{T}}}\sqrt \pi }}\exp \left[ { - 4\sqrt {\ln \left( 2 \right)} \frac{{{{\left( {{\nu _{\text{r}}}} \right)}^2}}}{{\Delta {\nu _{\text{T}}}^{2}}}} \right],$$
(33)
$$L\left( {{\nu _{\text{r}}},\Delta {\nu _{\text{T}}}} \right)=\frac{{\Delta {\nu _{\text{T}}}}}{{2\pi }}\frac{1}{{{{\left( {{\nu _{\text{r}}}} \right)}^2}+{{\left( {\Delta {\nu _{\text{T}}}/2} \right)}^2}}},$$
(34)
$$\Delta {\nu _{\text{D}}}=7.16 \times {10^{ - 6}}{\nu _{\text{o}}}\sqrt {\frac{T}{{{M_{\text{w}}}}}},$$
(35)
$$\Delta {\nu _{\text{L}}}=p\sqrt {\frac{{{T_{{\text{ref}}}}}}{T}} B,$$
(36)
$$\mathcal{R}_{i}^{{\text{Q}}}\left( {{\nu _r}} \right)=V_{{\text{p}}}^{{\text{Q}}}.$$
(37)

The values used for Bo, Do, gJ, γ2, A, and B are taken from the literature [58,59,60,61,62,63] and extrapolated to 532 nm where necessary.

Figure 11a shows an example of a calculated light scattering spectrum for N2 at T = 296 K and P = 1 atm that compares the relative intensity of each of the components: (1) Placzek track, (2) Q-branch rotational Raman, and (3) stokes and anti-Stokes rotational Raman scattering. Integrating the various components over frequency, rotational Raman scattering comprises approximately 1.4% (0.35% Q-branch; 1.05% S/AS) of the total signal. For the majority of the species considered, the contribution of the rotational Raman scattering signal to the total scattering signal is very small as in the case of N2 and can be neglected. However, CO2 is an exception with the rotational Raman scattering comprising more than 6% of the total scattering signal. Figure 11b shows a comparison of the Stokes and anti-Stokes spectra for CO2 and N2. The total differential Raman scattering cross section of CO2 is approximately ten times larger than that of N2, while the total differential Rayleigh scattering cross section only is 2.39 times larger.

Fig. 11
figure 11

a Calculated light scattering spectrum for N2 at T = 296 K and P = 1 atm. b Comparison of Stokes and anti-Stokes rotational Raman scattering for N2 ad CO2 at T = 296 K and P = 1 atm. c “Zoomed in” region of the spectra from b showing the interaction between the rotational Raman lines and the I2 filter spectrum. For S/AS rotational Raman scattering, the differential scattering cross section is plotted as \({F_J}\sigma _{i}^{{J \to {J^\prime }}}\)

For the current FRS experiment, there are two more considerations to take into account: (1) the overlap of the rotational Raman scattering spectra with the 532-nm bandpass filter (BPF) and (2) the overlap of the rotational Raman scattering spectra with the I2 absorption spectra. Both of these factors will alter the fraction of the rotational Raman scattering component that is collected within the total FRS signal. Figure 11b shows the measured spectral distribution of the BPF transmission around the center laser frequency (\(\Delta\)ν = 0). It is clear that the transmission of the BPF changes significantly over the span of the rotational Raman spectra, where species such as CO2 that have small rotational constants (and smaller spacing between adjacent rotational lines) have a higher fraction of their total signal transmitted through the BPF.

For FRS experiments, the interaction between the Rayleigh–Brillouin and I2 spectra is well established, but because of the large spectral bandwidth of I2, there will be overlap between the rotational Raman lines and the I2 spectra. Figure 11c shows a “zoomed in” portion the rotational Raman spectra shown in Fig. 11b along with an overlay of a portion of the \(B\left( {{}_{0}^{3}{\Pi _{{0^+}u}}} \right) \leftarrow X\left( {{}_{0}^{1}{\Sigma _g}^{+}} \right)\) electronic transition of iodine calculated with the code of Forkey et al. [34]. Figure 11c shows that the I2 spectrum is quite dense and that many of the I2 lines overlap with individual rotational Raman transitions. Considering the effects of the bandpass and I2 filters, the total rotational Raman scattering signal transmitted to the detector can be written as

$${S_{{\text{RR}},i}}={{\text{CI}}_0}{n_i}\left[ {\sigma _{i}^{{\text{Q}}}\mathop \int \limits_{{{\varvec{\upnu}}}} \mathcal{R}_{i}^{{\text{Q}}}\left( {{\nu _r}} \right){\tau _{I2}}\left( \nu \right){\tau _{{\text{BPF}}}}\left( \nu \right){\text{d}}\nu +\mathop \sum \limits_{{J=0}}^{\infty } \sigma _{i}^{{J \to {J^\prime }}}{F_J}\mathop \int \limits_{\nu } \mathcal{R}_{i}^{{J \to {J^\prime }}}\left( {{\nu _r},\Delta {\nu _{J \to {J^\prime }}}} \right){\tau _{I2}}\left( \nu \right){\tau _{{\text{BPF}}}}\left( \nu \right){\text{d}}\nu } \right],$$
(38)

where \({\tau _{I2}}\left( \nu \right)\;{\text{and}}\;{\tau _{{\text{BPF}}}}\left( \nu \right)~\)is the transmission of the I2 cell and bandpass filter, respectively. Since the location and width of the individual rotational Raman lines are species-specific, the exact level of the rotational Raman scattering signal contribution to the total FRS signal collected by the detector depends on the individual species. For example, calculations for the current experimental conditions show that for N2 at T = 296 K and P = 1 atm, approximately 36% and 30% of the Q-branch and Stokes/anti-Stokes rotational Raman scattering signal, respectively, transmit through the I2 and bandpass filters, while only 22% of the Cabannes line transmits through the filters. This implies that the fractional percent of the rotational Raman scattering component (FRR) increases for FRS as compared to traditional Rayleigh scattering (2% vs 1.4%). For CO2 at T = 296 K and P = 1 atm, approximately 44% and 49% of the Q-branch and Stokes/anti-Stokes rotational Raman scattering signal, respectively, transmit through the I2 and bandpass filters, while only 16% of the Cabannes line signal transmits through the filters and is collected. Thus, for CO2, there is a significant increase in FRR, increasing from 6% for traditional Rayleigh scattering to 16% for FRS.

The value of FRR can vary as a function of temperature since the Cabannes linewidth, \({\text{~}}{\mathcal{R}_i}\left( {{\nu _r}} \right)\) and the Raman linewidths, \(\mathcal{R}_{i}^{{\text{Q}}}\left( {{\nu _{\text{r}}}} \right)\), and \(\mathcal{R}_{i}^{{J \to {J^\prime }}}\left( {{\nu _{\text{r}}},\Delta {\nu _{J \to {J^\prime }}}} \right)\) can change width (and shape) as a function of gas temperature, while the I2 lines remain constant for a given I2 filter cell setting. Figure 12 shows the variation of FRR as a function of temperature for five gases. For all species considered, FRR decreases with increasing temperature. This is due to the fact that the Cabannes linewidth, \({\text{~}}{\mathcal{R}_i}\left( {{\nu _r}} \right)\) increases with increasing temperature due to Doppler broadening, while the rotational Raman linewidths remain largely constant due to competing effects of Doppler broadening and collisional broadening. Thus, with increasing temperature, an increasingly larger fraction of the RBS signal (as compared to the rotational Raman scattering signal) transmits through the I2 filter.

Fig. 12
figure 12

Variations in calculated fraction percent (FRR) of total FRS signal due to rotational Raman scattering (Q-branch and Stokes/anti-Stokes) as a function of temperature

Appendix 2: temperature-dependent values of \({c_{\text{int} }}~\)and \({\mu _{\text{B}}}\) for combustion-relevant species

Kinetic models of scattered light spectra require collision properties of gases that are used within the collision integral of the Boltzmann equation. For the Tenti S6 model [5], the collision integral is estimated using gas transport properties including the internal specific heat capacity (\({c_{\operatorname{int} }}\)) and the bulk viscosity (\({\mu _{\text{B}}}\)). Both \({c_{\operatorname{int} }}\) and \({\mu _{\text{B}}}\) depend on the relaxation rates of the internal degrees of freedom of the molecule.

2.1 Internal specific heat capacity (\({c_{\text{tint} }}\))

In general, the internal specific heat capacity can be calculated as

$${c_{\operatorname{int} }}=\frac{{\left( {5 - 3\gamma } \right)}}{{2\left( {\gamma - 1} \right)}},$$
(39)

where γ is the ratio of specific heats, which can be expressed in terms of the total number of internal degrees of freedom \(\left( f \right),\) as

$$\gamma =\frac{2}{f}+1,$$
(40)

and thus

$${c_{\operatorname{int} }}=\frac{{\left( {f - 3} \right)}}{2}.$$
(41)

The total number of degrees of freedom is written as f = ft + fr + fv, where ft = 3 is the number of translational degrees of freedom, fr is the number of rotational degrees of freedom given by

$${f_{\text{r}}}=\left\{ {\begin{array}{*{20}{l}} 2&{\quad {\text{linear}}\;{\text{molecule}}} \\ 3&{\quad {\text{non-linear}}\;{\text{molecule}}} \end{array}} \right.,$$
(42)

and \({f_v}\) is the number of vibrational degrees of freedom, given by

$${f_{\text{v}}}=\left\{ {\begin{array}{*{20}{l}} {2\left( {3N - 5} \right)}&{\quad {\text{linear}}\;{\text{molecule}}} \\ {2\left( {3N - 6} \right)}&{\quad {\text{non-linear}}\;{\text{molecule}}} \end{array}} \right.,$$
(43)

where N is the total number of atoms in the molecule. Note, the factor of 2 preceding the parenthetical arguments in Eq. (42) corresponds to the fact that each active vibrational mode has 2° of freedom. Since the number of accessible vibrational modes increases with increasing temperature, \(\gamma =\gamma \left( T \right)\) and thus the temperature dependence of \({c_{\operatorname{int} }}\) is embedded within \(\gamma\).

If the relaxation time scale of any internal degree of freedom is much longer than the characteristic time scales of sound propagating through the media, then the internal motion associated with that particular degree of freedom remains frozen on the timescale of the density fluctuations. For light scattering experiments, the characteristic sound frequencies are on the order of 1 GHz and thus internal degrees of freedom with relaxation times scales significantly longer than 1 ns are not active in the light scattering. For N2, O2, H2, CH4, CO, and CO2, the vibrational relaxation time scales are longer than 10− 7 seconds, even at the highest temperatures, and therefore only rotational modes contribute to the scattering. Under these conditions, an effective value of γ (due to rotational motion only) is written as

$${\gamma _{\text{r}}}=\frac{{{f_{\text{r}}}+5}}{{{f_{\text{r}}}+3}},$$
(44)

and the internal specific heat capacity is determined as

$${c_{\operatorname{int} }}=\left\{ {\begin{array}{*{20}{l}} 1&{\quad {\text{linear}}\;{\text{molecule}}} \\ {3/2}&{\quad {\text{non-linear}}\;{\text{molecule}}} \end{array}} \right..$$
(45)

For H2O the vibrational relaxation times range from ~ 10 ns at 300 K to < 1 ns seconds at flame temperatures and cannot be neglected based on timescale arguments. In this manner, \({c_{\operatorname{int} }}\) is calculated in the current work using Eq. (39), where \(\gamma ={C_{\text{p}}}/{C_{\text{v}}}\) and \({C_{\text{p}}}\) and \({C_{\text{v}}}\) are the heat capacity at constant pressure and volume, respectively. The values of \({C_{\text{p}}}\) are calculated using temperature-dependent polynomial curve fits from Kee et al. [64], and \({C_{\text{v}}}\) is calculated as \({C_{\text{v}}}={C_{\text{p}}}-R\), where \(R\) is the universal gas constant.

When all vibrational modes are active, H2O has 12 total internal degrees of freedom (3 translational, 3 rotational, and 6 vibrational), which leads to 3/2 ≤ \({c_{\operatorname{int} }}\) ≤ 9/2 over a full range of temperatures. The calculation of \({c_{\operatorname{int} }}\) for H2O using Eq. (39) at any given temperature is an estimation, thus the sensitivity of the synthetic FRS signals to the specified value of \({c_{\operatorname{int} }}\) is assessed. Figure 13 shows calculated synthetic FRS signals from pure H2O vapor at P = 1 atm over the full range of possible values of \({c_{\operatorname{int} }}\) (solid lines) at four different temperatures. The synthetic FRS signals for each value of \({c_{\operatorname{int} }}\) are normalized by the average synthetic FRS signal over the full range of \({c_{\operatorname{int} }}\) values. The symbols show the value of \({c_{\operatorname{int} }}\) calculated using Eq. (39) at each temperature. The results show minimal sensitivity of the calculated synthetic FRS signals to the specified value of \({c_{\operatorname{int} }}\) at 1 atm. For example for T > 700 K, varying \({c_{\operatorname{int} }}\) over the entire possible range of values results in variations of \({S_{{\text{FRS}},{{\text{H}}_2}{\text{O}}}}\) of less than 0.5%. At higher pressures (or increased y parameter values), the sensitivity of \({S_{{\text{FRS}},{{\text{H}}_2}{\text{O}}}}\) to variations in \({c_{\operatorname{int} }}~\)is expected to increase.

Fig. 13
figure 13

Variations in calculated water vapor FRS signals at \(P=1~\) atm as a function of specified \({c_{{\text{int}}}}\) values

Table 2 Temperature-dependent species properties

2.2 Bulk viscosity (µ B)

Temperature-dependent values of the bulk viscosity can be estimated from relaxation time measurements found within the literature. Assuming that rotational and vibrational modes relax independently, each with a single time scale, the bulk viscosity can be written as

$${\mu _{\text{B}}}={\mu _{{\text{B}},{\text{r}}}}+{\mu _{{\text{B}},{\text{v}}}},$$
(46)

where \({\mu _{{\text{B,r}}}}\) and \({\mu _{{\text{B,v}}}}\) are the rotational and vibrational contributions to the bulk viscosity and given by the following expressions:

$${\mu _{{\text{B,r}}}}={\left( {\gamma - 1} \right)^2}\frac{{{f_{\text{r}}}}}{2}p{\tau _{\text{r}}},$$
(47)
$${\mu _{{\text{B,v}}}}=~{\left( {\gamma - 1} \right)^2}\left( {\frac{{Cv}}{R} - \frac{{{f_{\text{r}}}+3}}{2}} \right)p{\tau _{\text{v}}},$$
(48)

where p is the pressure, τr is the rotational relaxation time, and \({\tau _v}\) is the vibrational relaxation time. As discussed above, when the vibrational relaxation times are long compared to the characteristic time scale of the density fluctuations, there is no vibrational–translational energy exchange and the vibrational modes are assumed to be “frozen”. For N2, O2, H2, CH4, CO, and CO2, it is assumed that only rotation contributes to µB and the bulk viscosity is calculated using Eq. (47) with \({\gamma _{\text{r}}}\) replacing \(\gamma\). For water vapor, both rotational and vibrational contributions are included and Eqs. (46)–(48) are used to calculate the bulk viscosity. The rotational and vibrational relaxation times for the considered species at various temperatures are taken from sources within the literature as reported in Table 2. The values of \(p{\tau _{\text{r}}}\) and \(p{\tau _{\text{v}}}\) are fit to either power law or logarithmic expressions that serve as convenient fits to the data. There is sufficient relaxation time data for all species at room temperature, but for many species the relaxation time data are sparse at higher temperatures. For these cases, existing bulk viscosity models were used to help guide the choice of fit (power law or logarithmic), but were not used as data to determine the fit. When compiling all the entirety of the relaxation rates, \(p{\tau _{\text{r}}}\) and \(p{\tau _{\text{v}}}\) (for water vapor) were fit to

$$p\tau =A{T^{\text{B}}}+C\ln T+D~.$$
(49)

Table 2 lists the constants for each species, the range of temperatures in which relaxation times were obtained, and the suggested range of temperatures over which the fit of Eq. (49) is valid. For H2O, the first row of constants correspond to \(p{\tau _{\text{r}}}\) and the second row of constants correspond to \(p{\tau _{\text{v}}}\). Table 2 also lists the calculated values of \({\mu _{\text{B}}}/\mu\) at 300 K. For N2, O2, H2, CH4, CO, and CO2, the values of µB/µ derived from Eqs. (47) and (49) compare very favorably with those reported within the literature, including the values inferred from direct measurements of the RBS spectra. For example, the current fit yields a value of \({\mu _{\text{B}}}/\mu\) = 0.74 for N2, while Gu et al. [14] and Gu and Ubachs [13] determined \({\mu _{\text{B}}}/\mu\) = 0.79 at room temperature. For CO2, the current fit yields a value of \({\mu _{\text{B}}}/\mu\) = 0.30, while Pan et al. [48], Lao et al. [52], Meijer et al. [53], and Gu et al. [12] determined \({\mu _{\text{B}}}/\mu\) = 0.25, 0.31, 0.39, and 0.38, respectively.

Because of the limited amount of bulk viscosity data, it is not possible to evaluate the accuracy of the temperature-dependent expressions for µB at elevated temperatures. Figure 14 shows example synthetic FRS signals as a function of temperature for four different species, N2, CH4, CO2, and H2O using (1) a constant \({\mu _{\text{B}}}/\mu\) ratio determined at room temperature (solid black line), implying that the temperature dependence of µB is exactly that of µ and (2) a temperature-dependent µB/µ ratio (solid red line), where µB is determined using Eqs. (46)–(49). In Fig. 14, the values of \({\mu _{\text{B}}}/\mu\) are shown with dashed lines. The results shown in Fig. 14 indicate little observable difference between the synthetic FRS signals calculated with the two different sets of \({\mu _{\text{B}}}/\mu\) values, even when there are significant differences in \({\mu _{\text{B}}}\) as is the case for CH4.

Fig. 14
figure 14

Normalized synthetic FRS signal curves as a function of temperature for constant µB/µ (solid black line) and a temperature-dependent value of µB/µ (solid red line) according to Eqs. (46)–(49). The ratio of µB/µ for each calculation is given as the dashed lines and shown on the secondary y-axis. The reference condition is N2 at T = 296 K

To further investigate the sensitivity of the synthetic FRS signals to the chosen value of \({\mu _{\text{B}}}\), the synthetic FRS signals are calculated for all species as a function of temperature for varying values of \({\mu _{\text{B}}}\) The percent difference between the synthetic FRS signal calculated with an arbitrary value of the bulk viscosity, \(\mu _{{\text{B}}}^{*}\), and the synthetic FRS signal calculated with the bulk viscosity determined using Eqs. (46)–(49) is defined by

$${\epsilon _{{\mu _{\text{B}}}}}=\frac{{\left| {{S_{{\text{FRS}}}}\left( {\mu _{B}^{*}} \right) - {S_{{\text{FRS}}}}\left( {{\mu _B}} \right)} \right|}}{{{S_{{\text{FRS}}}}\left( {{\mu _B}} \right)}}.$$
(50)

Figure 15 shows the results of varying the value of the bulk viscosity by a factor of 100 (0.1 ≤ \(\mu _{{\text{B}}}^{*}/\mu\) ≤ 10) for temperatures of 300 K, 700 K, 1200 K, and 2000 K. For higher temperatures (T ≥ 700 K), the variation of the bulk viscosity leads to negligible variations in the calculated FRS signal over the full range of bulk viscosities tested. Only at room temperature (T = 300 K), do very small values of the bulk viscosity lead to notable changes in the calculated FRS signal. However, over the range of bulk viscosities reported within the literature, variations in µB lead to changes in the synthetic FRS signal by less than 1%. Overall, the results shown in Figs. 14 and 15 demonstrate that the calculated FRS signal is insensitive to the value of bulk viscosity, as long as µB does not approach zero. However, it should be noted that the FRS signal will likely become increasingly sensitive to the values of µB as pressure (y parameter) increases.

Fig. 15
figure 15

Sensitivity of the synthetic FRS signals to changes in values of bulk viscosity for a T = 300 K, b T = 700 K, c T = 1200 K, and d T = 2000 K

Rights and permissions

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Mcmanus, T.A., Monje, I.T. & Sutton, J.A. Experimental assessment of the Tenti S6 model for combustion-relevant gases and filtered Rayleigh scattering applications. Appl. Phys. B 125, 13 (2019). https://doi.org/10.1007/s00340-018-7121-8

Download citation

  • Received:

  • Accepted:

  • Published:

  • DOI: https://doi.org/10.1007/s00340-018-7121-8

Navigation