Sunteți pe pagina 1din 11
International Journal of Engineering PERGAMON International Journal of Engineering Science 36 (1998) 1313-1323 =Science Bulk viscosity in the Navier-Stokes equations George Emanuel * School of Aerospace and Mechanical Engineering, The University of Oklahoma, Norman, OK 73019, USA Received 23 January 1996; accepted 24 November 1997 Abstract This article discusses the physics associated with the bulk viscosity coefficient jx, as it appears in the compressible Navier-Stokes equations. Thus, the active rotational and vibrational modes of a polyatomic molecule are discussed with emphasis on the importance of local thermodynamic equilibrium. This condition is necessary if j is to only depend on the thermodynamic state, as required for the Navier-Stokes equations. A new perturbation formulation is provided for the absorption and dispersion coefficients and the entropy production that holds even when py greatly exceeds the shear viscosity, as is the case for gases such as CO>. A comparison for CO; shows excellent agreement between exact results and the low-frequency formula used for the absorption coefficient. This agreement stems from a large value for the Peclet number. © 1998 Elsevier Science Ltd. All rights reserved. 1. Introduction Studies of sonic attenuation, from which the bulk viscosity can be determined, started in the later part of the nineteenth century. As demonstrated by review articles [1-9], a number of which list hundreds of references, it was a prolific research topic. Over time, the experimental technique improved with the use of crystals for generating and receiving the signal, better instrumentation, and with careful gas handling techniques [10]. Most of this work, however, both experimental and theoretical, is not pertinent to continuum fluid dynamics, which is the focus of this article. Our interest is in gas flows that are governed by the steady or unsteady, compressible, Navier-Stokes (N-S) equations, which here includes the mass and energy equations. In this context, thermodynamic and transport properties, including the bulk viscosity, pip, are functions only of the thermodynamic state. Moreover, these parameters are real, frequency independent, and a Kramers—Kronig relation [11] is not relevant. The condition for ji, to be real and frequency independent is discussed shortly. Tel: + 1-405-325-5011; fax: + 1-405-325-1088 0020-7225/98/S19.00 © 1998 Elsevier Science Ltd. All rights reserved. PII: $0020-7225(98)00020-2 1314 G. Emanuel | International Journal of Engineering Science 36 (1998) 1313-1323 For aerodynamics and the fluid dynamics of air, the terms containing jy in the N-S equations are often inconsequential. Although i, for air is not zero, it has long been conventional to take it as zero, i.e. Stokes hypothesis. For a laminar boundary layer, Stokes hypothesis is appropriate, since ty introduces a negligible third-order effect [12]. Stokes hypothesis, however, cannot be used in the analysis of shock wave structure, except for a monatomic gas. In addition, there are many gases for which the ratio & = 14/1, where is the shear viscosity, is large compared to unity. This circumstance is the primary subject of this article; it has been discussed for a laminar boundary layer [13], turbulence [14], and is of interest for planetary atmosphere aerodynamics [15, 16]. For instance, the Venus and Mars atmospheres are largely CO>, for which & is about 2100 [17]. Another application occurs when a vapor exhibits anomalous behavior [18,19], which happens near the coexistence curve and when the specific heats are large. Large specific heats, however, also result in a large value for 4 Thus, modeling of the complex viscous behavior of a retrograde [18] or Bethe-Zel'dovich-Thompson [19] fluid requires an accurate knowledge of He The bulk viscosity is determined by propagating a weak ultrasonic signal one-dimensionally through a gas. As is customary, wave attenuation is assumed to be caused by heat conduction and viscosity. For a polyatomic molecule, the bulk viscosity is finite because of the relaxation of internal rotational and vibrational modes. Hence, there is energy transfer between these modes and the translational mode. Rotational relaxation is generally quite rapid requiring only a few molecular collisions. In this situation, there are sufficient collisions to easily maintain local thermodynamic equilibrium (LTE) with the translational mode, On the other hand, vibrational relaxation may require several thousand collisions for equilibration. The relaxation time, z;, of internal mode / is given as pr, where p is the pressure and where this combination depends only on the temperature. For vibration, this dependence is provided by a Landau- Teller type formula [20]. As Tisza [17] has pointed out, a frequency-independent bulk viscosity suffices when cot,<1 for all active internal modes, where « is the angular frequency of the signal. In this situation, there are sufficient collisions to maintain LTE. At a sufficiently high frequency, of course, a rate equation approach with a set of characteristic relaxation times is required. In other words, if wr; is not sufficiently small a complex valued, frequency dependent bulk viscosity is obtained. In this circumstance, acoustic theory [21,22] is utilized that is based on the Euler equations. Consequently, a frequency dependent bulk viscosity depends on the relaxation processes but not on the Reynolds or Prandtl numbers. For a continuum flow, the N-S equations implicitly assume LTE conditions for internal modes, and any bulk viscosity measurement should satisfy Tisza’s condition if the parameter is to be used in these equations. In an ultrasonic measurement, « exceeds 20 kHz. This frequency is usually much larger than what is encountered in an unsteady flow, including a turbulent one. Hence, an ultrasonic py measurement, performed under LTE conditions, should hold for steady and most unsteady N-S flows. In passing, we note that at a high density the modified Enskog theory has jay proportional to the square of the density [23], even for a monatomic gas. This phenomenon is not associated with the relaxation of internal modes and j4, is generally much smaller than that of a polyatomic molecule with one or more active vibrational modes. This high density aspect is excluded from the discussion. G. Emanuel { International Journal of Engineering Science 36 (1998) 1313-1323 1315 Although the oz; requirement is important, it apparently is often overlooked. For example, Ref. [10] provides the absorption in COz over a temperature range and for a 53-147 kHz frequency range. The value of pt for CO,-CO; collisions [13] at 303 K is 6.8 x 10° atm s. Here, t is the relaxation time of the (0,1',0) vibrational state, which is the lowest lying v> excited state. Since relaxation times are given as pt, we have the LTE requirement (pt) > 1 and that Pr is of order unity. Since Re inversely depends on @, while Pr is independent of @, this corresponds to the low-frequency approximation. The expansions are introduced [30] % 1 x14 3,40(4), (8a) A 1 4,04, 18} 6 Ret 7 Re)” (18) and a and f; are evaluated from Eqs. (12a) and (12b), with the result _(@r+y+2) y—Prpt a2 = 5 — Bi + sp (19a) B= (9b) These formulas are in accord with those in Tsien and Schamberg [30] providing their u$=0 assumption is introduced. None of the foregoing approximations or assumptions, however, are invoked in Eqs. (13) and (15). It is common practice to use the large Reynolds number approximations to obtain a value for xf. To derive the form, e.g. shown in Prangsma et al. [25]. set * & (20a) (20b) G. Emanuel | International Journal of Engineering Science 36 (1998) 1313-1323 1321 Since is per unit wavelength, replace f with Re2*f*; hence, Eq. (9b) becomes 4g 7 6 = Reiter = oro] n(F+48 amy] en where Re is given by Eq. (17b). This yields ee [4 roa £ [Gert ( -z)st+aa} 22) P where quantities, such as a*, p*, ..., are evaluated for the undisturbed gas. The low-frequency formula used in the experimental determination of pi is thereby obtained. 2.5. Numerical results Calculations are performed for CO, with T)=300 K, po=10 atm, #=2100, and an angular frequency ranging from 20 to 60 kHz. Svehla [31] provides the other transport properties. The chosen po and c values satisfy condition (1). Equations (13) and (15) yield a and f values that agree with those from Eqs. (18a)-(18b) and (19a)-(19b) to, at least, six significant digits. Moreover, both « values are unity, also to six digits, while the 8 values are of the order of 10. As noted, a dispersion measurement cannot be used to evaluate 4 when 2 is so close to unity, while the attenuation per wavelength is small but measurable. Equation (22) is, therefore, accurate for an LTE measurement when 4->1, even though the original derivation utilizes assumptions that are no longer warranted. With 4>+1, Re is greatly reduced, while Pr is now quite large. What is important is that Eq. (22) holds when the Peclet number is large. 3. Concluding remarks The importance of LTE for the transport of energy between the internal and translational modes of a polyatomic molecule is stressed. This condition is required if the bulk viscosity is to be a function of the thermodynamic state in the compressible Navier-Stokes equations. This is generally the case for steady and unsteady (turbulent) flows of air. Although a viscosity coefficient, the bulk viscosity is associated with energy transfer rather than momentum transfer. A number of molecules, such as CO, have very large values for the ratio of the bulk viscosity to the shear viscosity, 2. Nevertheless, the low-frequency formula, currently in use for the bulk viscosity, still holds for these molecules. This formula depends on thermodynamic properties, such as the specific heats and the speed of sound, the thermal conductivity and shear viscosity, and the frequency and attenuation of the ultrasonic signal. Under LTE conditions, the dispersion of the signal is too small to be useful. There appears to be few, if any, ultrasonic measurements at temperatures well above room temperature. For instance, # should start to increase for air above 600 K when vibrational excitation of Oz occurs. At a somewhat higher temperature, the rate of increase should be more rapid when vibrational excitation of Nz occurs. Measurements of the bulk viscosity for 1322 G. Emanuel | International Journal of Engineering Science 36 (1998) 1313-1323 molecules of interest in dense gas flows [18, 19] are needed. These measurements should also be at elevated temperatures. Unlike other thermodynamic and transport properties, there currently is only one experimental method for measuring the bulk viscosity of a gas. A second method would certainly be desirable. Such a method has been proposed and theoretically assessed [32]. The approach should be applicable to polyatomics, such as CO, SFe, and Ip, as a dense gas. In this approach, a shock tube experiment yields a shock that is many thousands of mean free paths thick, and both the linear Newtonian assumption and the LTE condition hold. A measurement of the shock thickness along with a numerical solution for the structure of the shock wave then yields 1. Acknowledgements The author is indebted to Professor B.M. Argrow for his comments and to M. Ishmail for the numerical computations References U1] Herzfeld KF and Litovitz TA, Absorption and dispersion of ultrasonic waves. New York: Academic Press, 1959, {2} Truesdell C. J Rat Mech Anal 1953;2:643, [3] Hunt FV. J Acous Soc Am 1955;27:1019. [4] Lick W. Adv Appl Mech 1967;10:1 [5] Richards WT. Rev Mod Phys 1939:11:36, [6] Markham JJ, Beyer RT, Lindsay RB. Rev Mod Phys 1951;23:353 [7] Karim SM, Rosenhead L. Rev Mod Phys 1952;24:108, {8) Lighthill MJ, In: Batchelor GK, Davies RM, editors. Survey in mech. Cambridge: University Press, 1956:250. [9] Bhatia AB, Ultrasonic absorption. Oxford: Clarendon Press, 1967. [10] Overbeck CJ, Kendall HC. J Acous Soc Am 1941513:26. [11] O'Donnell M, Janes ET, Miller JG. J Acous Soc Am 1981;69:696. [12] Van Dyke M. In: Riddell FR, editor. Hypersonic flow research. New York: Academic Press, 1962:37. [13] Emanuel G. Phys Fluids 1992;44:491 [14] Orou JC, Johnson JA, IIL. Phys Fluids 1994:6:415, {15] Candler GV, Reprint from AIAA 1990;90;1695. [16] Park C, Howe JT, Jaffe RL, Candler GV, Reprint from AIAA 1991;91:0464, [17] Tisza L. Phys Rev 1942;61:531 [18] Thompson PA, Carofano GC, Kim Y-G, J Fluid Mech 1986;166:57. [19] Cramer MS. In: Kluwick A, editor, Nonlinear waves in real fluids, New York: Springer, 1991:91 [20] Vincenti WG, Kruger Jr. CH. Introduction to physical gas dynamics. New York: Wiley, 1965 [21] Kneser HO. In: Mason WP, editor. Physical acoustics, vol. Il, Part A. New York: Academic Press, 1965:133 [22] Bauer H-J. In: Mason WP, editor. Physical acoustics, vol. II, Part A. New York: Academic Press, 1965:47. [23] Hanley HJM, Cohen EGD. Physica 1976;83A:215. [24] Callear AB, Lambert JD. In: Bamford CH, Tipper CFH, editors. Chemical kinetics, vol. 3. The formation and decay of excited species. New York: Elsevier, 1969:182, [25] Prangsma GJ, Alberga AH, Beenakker JJM. Physica 1973;64:278, [26] Monchik L, Yun KS, Mason BA. J Chem Phys 1963;39:654, G. Emanuel} International Journal of Engineering Science 36 (1998) 1313-1323 [27] Meador WE, Miner GA, Townsend LW. Phys Fluids 1996;8:258. 28] Emanuel G. Phys Fluids 1996;8:1994. 29] Emanuel G, Analytical fluid dynamics. Boca Raton, Fl: CRC Press, 1994. [0] Tsien H-S, Schamberg R. J Acous Soc Am 1946;18:334. BI] Svehla RA, NASA Tech. Rep. TR-132, 1962, 32] Emanuel G, Argrow BM. Phys Fluids 1994;6:3203. 1323

S-ar putea să vă placă și