rarefied gas dynamics

31
Ann. Rev. Fluid Mech. 1989.21: 38717 Copyright © /989 by Annual Reviews Inc. All rights reserved RAREFIED GAS DYNAMICS E. P. Muntz Department of Aerospace Engineering, University of Southern California, University Park, Los Angeles, Ca lifornia 90089- 1 19 1 INTRODUCTION Rarefied gas dynam ics is a diverse field, encompassing, for example, high- altitude hypersonic flow fields, the reflective and reactive characteristics of gases interacting with solid and l iquid surfaces, energy-transfer ph enomena in molecular col lisions, aerosol dynamics, cluster formation and topology, flows induced by evaporation and condensation, upper-atmospheric dy - namics, and the attainment of milli-Ke lvin temperatures by flow cooling techniques. Other subjects in the field include vacuum-pump perfor- mance, spacecraſt contamination, a variety of interactions due to thruster plumes, spacecraſt charging, and gas and isotope separations. Underlying all of these subjects is the central theme of the field of rarefied gas dynamics: the study of gas flow phenomena in which the discrete molecular nature of the gas cannot be safely ignored. The field has a rich heritage of analysis applied to the s tudy of flows where concep ts and techniques related to nonequilibrium statistical m e chanics are important. An equally respected tradition is the development and application of instrumentation techniques that can be used to study the details of molecular motion in gas flows, as well as to study flow-generated populations of internal energy states and the characteristics of gases after surface encounters. Rarefied gas dynamics is founded on the pioneering work in the kinetic theory of gases begun in the latter half of the nineteenth century, which has continued to the present. What today we call rarefied gas d ynamics was discussed as early as 1 934 by Zahm (1 934), but research began in earnest shortly after World War II (Tsien 1 946). Because of growing interest in space and flight at extreme altitude, attention was drawn to a flow regime that had not been seriously considered as a concern of either aerodynamics or gas dynami cs. In this flow regime, the molecular mean free path in the gas, , becomes significant compared with either a c harac- 387 0066-4189/89/01 15 -0387$02.00 Annu. Rev. Fluid Mech. 1989.21:387-422. Downloaded from www.annualreviews.org by West Virginia University on 01/22/14. For personal use only.

Upload: mohammad-kazemi

Post on 21-Oct-2015

135 views

Category:

Documents


3 download

DESCRIPTION

Dynamics of Rarefied Gas Dynamics

TRANSCRIPT

Page 1: Rarefied Gas Dynamics

Ann. Rev. Fluid Mech. 1989.21: 387-417 Copyright © /989 by Annual Reviews Inc. All rights reserved

RAREFIED GAS DYNAMICS

E. P. Muntz

Department of Aerospace Engineering, University of Southern California, University Park, Los Angeles, California 90089- 1 1 91

INTRODUCTION

Rarefied gas dynamics is a diverse field, encompassing, for example, high­altitude hypersonic flow fields, the reflective and reactive characteristics of gases interacting with solid and l iquid surfaces, energy-transfer phenomena in molecular collisions, aerosol dynamics, cluster formation and topology, flows induced by evaporation and condensation, upper-atmospheric dy­namics, and the attainment of milli-Kelvin temperatures by flow cooling techniques. Other subj ects in the field include vacuum-pump perfor­mance, spacecraft contamination, a variety of interactions due to thruster plumes, spacecraft charging, and gas and isotope separations. Underlying all of these subjects is the central theme of the field of rarefied gas dynamics: the study of gas flow phenomena in which the discrete molecular nature of the gas cannot be safely ignored. The field has a rich heritage of analysis applied to the s tudy of flows where concep ts and techniques related to nonequilibrium statistical mechanics are important. An equally respected tradition is the development and application of instrumentation techniques that can be used to study the details of molecular motion in gas flows, as well as to study flow-generated populations of internal energy states and the characteristics of gases after surface encounters.

Rarefied gas dynamics is founded on the pioneering work in the kinetic theory of gases begun in the latter half of the nineteenth century, which has continued to the present. What today we call rarefied gas dynamics was discussed as early as 1 934 by Zahm (1 934), but research began in earnest shortly after World War II (Tsien 1 946). Because of growing interest in space and flight at extreme altitude, attention was drawn to a flow regime that had not been seriously considered as a concern of either aerodynamics or gas dynamics. In this flow regime, the molecular mean free path in the gas, .A., becomes significant compared with either a charac-

387 0066-4189/89/0 1 1 5 -0387$02.00

Ann

u. R

ev. F

luid

Mec

h. 1

989.

21:3

87-4

22. D

ownl

oade

d fr

om w

ww

.ann

ualr

evie

ws.

org

by W

est V

irgi

nia

Uni

vers

ity o

n 01

/22/

14. F

or p

erso

nal u

se o

nly.

Page 2: Rarefied Gas Dynamics

388 MUNTZ

teristic distance for important flow-property changes to take place or the size of the flow field, or more simply the size of the object creating the flow field. From about 1 950 to the late 1 960s there was active research, supported by several governments; these efforts were principally in the United States and the USSR, but major contributions were also made in France, Japan, Canada, Australia, Italy, the Federal Republic of Germany, The Netherlands, India, Poland, Y ugoslavia, the United Kingdom, and elsewhere. During this period a biannual series of Inter­national Symposia on Rarefied Gas Dynamics was initiated under the auspices of an informal advisory committee. The proceedings of these meetings have been published, beginning in 1 960; they form a unique record of the field (see Table I).

As it turned out, rarefied gas dynamics initially was somewhat ahead of its time. When all was said and done there were no early, critical aerospace­engineering problems associated with the regime. In the 1 950s and 1 960s reentries were ballistic, and thus vehicles quickly transited the uncertainty of the rarefied transitional flow regime at altitudes around 80-1 00 km. Besides, important things like heat transfer and aerodynamic forces were small in the rarefied regime compared with their effects at lower altitudes. In this time period, continuous or quasi-continuous flight in the 80-1 50 km altitude range was not a major concern. At higher altitudes satel lites did not have sufficient mechanical or electronic life for the low-level effects of the ambient atmosphere to be of great significance. As a result , from the late 1 960s to the early 1 980s governmental interest in the field, par­ticularly in the United States, was at a low level . However, work continued in many countries, most notably in the USSR, Australia, Japan, and Europe. The International Symposia on Rarefied Gas Dynamics also continued because of intrinsic academic interest and because the field impacts many areas outside of space and high-altitude flight. Despite difficulties, significant scientific progress was made during this period.

We now come to what is the principal subject of this review: the likely consequences of a modern resurgence of governmental interest in rarefied gas dynamics and the direction work in this field can be expected to take in the coming decade. While there may be intense academic stimulation in a field, a fact of modern research life is that governments are usually required to energize major research endeavors. In the early and mid 1 980s several events have conspired to once again trigger governmental interest in work related to rarefied gas dynamics.

During the flights of Space Shuttle's STS-3 and STS-4 a low-light-level television camera on the orbiters recorded a gaseous glow above the windward tail surfaces of the vehicle (Banks et al. 1 983, Mende et al. 1 983). In the same period it was noticed that thermal insulation blankets returned

Ann

u. R

ev. F

luid

Mec

h. 1

989.

21:3

87-4

22. D

ownl

oade

d fr

om w

ww

.ann

ualr

evie

ws.

org

by W

est V

irgi

nia

Uni

vers

ity o

n 01

/22/

14. F

or p

erso

nal u

se o

nly.

Page 3: Rarefied Gas Dynamics

RAREFIED GAS DYNAMICS 389

Table 1 List o f the published proceedings o f the International Symposia o n Rarefied Gas Dynamics

RGD 1. Rarefied Gas Dynamics. 1960. Proceedings oJthe 1st International Symposium,

ed. F. M. Devienne. Paris: Pergamon RGD 2. Rarefied Gas Dynamics. 1 961. Proceedings oj the 2nd International Symposium,

ed. L. Talbot. New York: Academic RGD 3. Rarefied Gas Dynamics. 1963. Proceedings oJthe 3rd International Symposium,

Vols. 1 ,2, ed. J. A. Laurmann. New York: Academic RGD 4. Rarefied Gas Dynamics. 1965. Proceedings oj the 4th International Symposium,

Vols. 1,2, ed. 1. H. deLeeuw. New York: Academic RGD 5. Rarefied Gas Dynamics. 1967. Proceedin.qs oJ the 5th International Symposium,

Vols. 1, 2, ed. C. L. Bmndin. London: Academic RGD 6. Rarefied Gas Dynamics. 1969. Proceedings oj the 6th International Symposium,

Vols. 1 ,2, ed. L. Trilling, H. Y. Wachman. New York: Academic RGD 7. Rarefied Gas Dynamics. 1 971. Proceedings oJthe 7th International Symposium,

Vols. 1,2, cd. D. Dini, C. Ccrcignani, S. Nocilla. Pisa: Editrice Tecnico Scientifica RGD 8. Rarefied Gas Dynamics. 1974. Proceedings oJ the 8th International Symposium,

Vols. 1 ,2, ed. K. Karamcheti. New York: Academic RGD 9. Rarefied Gas Dynamics. 1974. Proceedings oj the 9th International Symposium,

Vols. 1,2, ed. M. Becker, M. Fiebig. Porz-Wahn, Germ: DFVLR-Press RGD 10. Rarefied Gas Dynamics. 1977. Proceedings oJthe 10th International Symposium,

Parts I and 2 of Vol. 51 of Progress in Astronautics and Aeronautics, ed. L. Potter. New York: AIAA

RGD 1 1. Rarefied Gas Dynamics. 1979. Proceedings oj the 11 th International Symposium,

Vols. 1 ,2, ed. R. Campargue. Paris: Commissariat a l'Energie Atomique RGD 12. Rarefied Gas Dynamics. 1 981. Proceedings oJ the 12th International Symposium,

Parts 1 and 2 of Vol. 74 of Progress in Astronautics and Aeronautics, ed. S. Fisher.

New York: AIAA

RGD 13. Rarefied Gas Dynamics. 1985. Proceedings oj the 13th International Symposium,

Vols. 1,2, ed. O. M. Belotserkovskii, M. N. Kogan, C. S. Kutateladze, A. K. Rebrov. New York: Plenum

RGD 14. Rarefied Gas Dynamics. 1 984. Proceedings oJ the 14th International Symposium,

Vols. 1,2, ed. H. Oguchi. Tokyo: University of Tokyo Press RGD 1 5. Rarefied Gas Dynamics. 1986. Proceedings oj the 15th International Symposium,

Vols. 1,2, ed. V. Bofti, C. Cercignani. Stuttgart: B. G. Teubner RGD 1 6. Rarefied Gas Dynamics. 1989. Proceedings oj the 16th International Symposium,

Progress in Astronautics and Aeronautics, ed. E. P. Muntz, D. Weaver, D. Campbell.

Washington: AIAA. In press

from orbit on the STS orbiters had been severely eroded (Whitaker 1 983, Peters et al. 1 983). In the early 1 980s i t also became evident that modern satellite computer systems frequently required re-booting due to upsets as a result of discharges traced to differential charging on satellite surfaces. The upsets were most frequent as the vehicles passed through the midnight to dawn quadrant of their orbits (Garrett 1 980, Fennell et al. 1 983) . All of these phenomena indicated that in addition to the effects of the

Ann

u. R

ev. F

luid

Mec

h. 1

989.

21:3

87-4

22. D

ownl

oade

d fr

om w

ww

.ann

ualr

evie

ws.

org

by W

est V

irgi

nia

Uni

vers

ity o

n 01

/22/

14. F

or p

erso

nal u

se o

nly.

Page 4: Rarefied Gas Dynamics

390 MUNTZ

magnetosphere's radiation belts, there are interactions between local space environments and satellite systems of practical, measurable sig­nificance. As extended satellite lifetimes are realized, the contamination of optical components and thermal control surfaces becomes increasingly important. Major sources of contamination are the flow fields of control­and station-keeping thrusters. Optical contamination as a result of the ambient high-speed atmosphere interacting with exhaust gases has also been recognized as a potentially serious space-system problem. These occurrences represent phenomena that have been studied within the field of rarefied gas dynamics. Generally, increasingly frequent visits to space, as well as the intention of several countries to man permanent low-Earth­orbit space stations and to establish permanent lunar or other bases, leads one to expect numerous rarefied-gas-dynamic problems associated with activities such as the mining and processing of native resources in space. Even the old idea of space scoops (Berner & Camac 1 961 ) has recently been rejuvenated in both its original form (Ramohalli et al. 1 987) and a variant (Muntz & Orme 1 987) .

Several nations have announced their intent to pursue the goal of trans­atmospheric flight with hypersonic winged vehicles. A representative flight corridor for such vehicles is shown in Figure 1 . While these airplanes tend to fly at rather low altitudes during acceleration in order to operate air­breathing engines, they also necessarily have quite sharp leading edges.

ItO 100

90

80

E = 70

� 60 � � 50 ..J <[

40

30

20

10

NASP RE-ENTRY

NASP CRUISE

AOTV-AFE '--I i I I I t /

NASP SPACE LAUNCH

8s � 10% STANDOFF RN- fern

o 2 3 4 5 6 7 8 9 10 11

SPEED (km/sec.) Figure 1 Flight corridor for NASP and aerobraking vehicles, with the ratio of the maximum slope shock thickness to shock standoff distance superimposed for leading-edge radii RN•

Ann

u. R

ev. F

luid

Mec

h. 1

989.

21:3

87-4

22. D

ownl

oade

d fr

om w

ww

.ann

ualr

evie

ws.

org

by W

est V

irgi

nia

Uni

vers

ity o

n 01

/22/

14. F

or p

erso

nal u

se o

nly.

Page 5: Rarefied Gas Dynamics

RAREFIED GAS DYNAMICS 39 1

Chapman et al. (1987) have pointed out that at the sharp leading edges of the vehicle, where heat-transfer problems are severe, a significant portion of the trans atmospheric flight corridor will produce shock-wave thick­nesses that are at least 1 0% of the shock standoff distance. To illustrate this point, lines indicating where the shbck-wave thickness has reached 1 0% of the standoff distance for 1 cm and 1 0 cm leading-edge radii are superimposed on the flight corridor in Figure 1 . For aero assisted orbital transfer trajectories [Aeroassist Flight Experiment (A FE)], the shock thick­ness is of the same order as the shock standoff distance even at a typical minimum altitude of 75 km (Bird 1 986). Thick shock waves are important because significant radiation and chemistry can take place within the shock wave (Bird 1985, 1986, 1987, Moss 1987) . Thus an old rarefied­gas-dynamics problem- shock-wave structure-appears with the addi­tional complications of chemistry and radiation within the shock struc­ture. Transatmospheric and aeroassisted vehicles also bring up, through attempts to extend continuum approaches to higher altitudes, velocity-slip and temperature-jump boundary conditions associated with the kinetic or Knudsen layer at the interface between the gas and the vehicle's surface. The hypersonic context of many of these effects has been surveyed by Cheng (1 988).

At any particular time, researchers in rarefied gas dynamics seem to be involved in one or more areas that are of great i nterest to other scientific fields. During the 1970s, gas-dynamic isotope separation techniques and the development of flow cooling to aid in the spectroscopic study of complicated molecules were two examples of such involvement. For the mid-1 980s and into the 1 990s, the unique properties of clusters created by gas-dynamic sources provide possibilities for stimulating interactions with investigators studying surface depositions of cluster complexes, cluster structures, and cluster chemistry. Related to this is the study of the dynamics of aerosol clouds, where the particles have very small diameters and may be continually condensing and evaporating, as well as chemically active. The presence of such clouds both in nature and in industrial applications has attracted the rarefied-gas-dynamics research community to address these problems.

In common with the rest of fluid mechanics, rarefied gas dynamics applies to a variety of technological devices in disparate scientific areas while at the same time having its own intrinsic attraction. In this article I review the status and some likely future directions for research in rarefied gas dynamics, keeping in mind the weighting factors introduced by govern­mental involvement. The particular subjects have been selected for two reasons: first, to represent the status or scientific position of the field in what are considered to be important areas, and second, to raise possibilities for interesting and useful future research.

Ann

u. R

ev. F

luid

Mec

h. 1

989.

21:3

87-4

22. D

ownl

oade

d fr

om w

ww

.ann

ualr

evie

ws.

org

by W

est V

irgi

nia

Uni

vers

ity o

n 01

/22/

14. F

or p

erso

nal u

se o

nly.

Page 6: Rarefied Gas Dynamics

392 MUNTZ

AVAILABLE LITERATURE

Rarefied gas dynamics' literature, as well as its basis in the kinetic theory of gases, is well organized. The chronological l isting of books, mono­graphs, and reviews presented here in Table 2 are for the convenience of readers who would like to obtain either background or in-depth knowl­edge of the field. A quick catch-up can be obtained from the reviews of

Table 2 Background reading list (in chronological order) for rarefied gas dynamics

Knudsen, M. 1934. The Kinetic Theory o/Gases. London: Methuen & Co. Loeb, L. B. 1 934. Kinetic Theory 0/ Gases. New York/Toronto/London: McGraw-Hill

Kennard, E. H. 1 938. Kinetic Theory o/Gases. New York: McGraw-Hill Jeans, J. 1 940. An Introduction to the Kinetic Theory o/Gases. London: Cambridge

Univ. Press Dushman, S. 1944. Scientific Foundations of Vacuum Technique. New York: Wiley Chapman, S., Cowling, T. G. 1 952. The Mathematical Theory 0/ Non-Uniform Gases.

Cambridge: Cambridge Univ. Press

Jeans, J. 1954. The Dynamical Theory o/Gases. New York: Dover. 2nd ed. Hirschfelder, J. 0., Curtis, C. F., Bird, R. B. 1954. Molecular Theory o/Gases and

Liquids. New York: Wiley Patterson, G. N. 1956. Molecular Flow o/Gases. New York: Wiley

Grad, H . 1958. Principles of the kinetic theory of gases. In Encyclopedia 0/ Physics, ed. G. Flugge, XII: 205-93. Bcrlin/G6ttingen/Heidclberg: Springer-Verlag

Present, R. D. 1958. Kinetic Theory o/Gases. New York: McGraw-Hill Schaaf, S. A., Chambre, P. L. 1 958. Flow of rarefied gases. In High Speed Aerodynamics

and Jet PropulSion, ed. H. W. Emmons, Vol. 3, Sect . H. Princeton, NJ: Princeton Univ. Press

Schaaf, S. A. 1958. Mechanics of rarefied gases. In Encyclopedia of Physics, ed. S. Flugge. VllI/2: 59 1-624. Berlin/G6ttingen/Heidelberg: Springer-Verlag

Vincenti, W. G., Kruger, C. H. 1 965 . Introduction to Physical Gas Dynamics. New York: Wiley

Shidlovskii, V. P. 1965. Introduction to the Dynamics 0/ Rarefied Gas. Moscow: Nauka (Trans\., 1967, ed. J. A. Laurmann. New York: Elsevier)

Cercignani, C. 1 969. Mathematical Methods.in Kinetic Theory. New York: Plenum Kogan, M. N. 1969. Rarefied Gas Dynamics. Moscow: Nauka (Transl., 1969,

ed. L. Trilling. New York: Plenum) Sherman, F. S. 1969. The transition from continuum to molecular flow. Ann. Rev. Fluid

M echo 1: 317-40 Patterson, G. N. 1 97 1 . Introduction to the Kinetic Theory o/Gas Flows. Toronto:

Univ. Toronto Press Kogan, M. N. 1 973. Molecular gas dynamics. Ann. Rev. Fluid Mech. 5: 383-404 Cercignani, C. 1 975. Theory 0/ Application o/the Boltzmann Equation. New York:

Elsevier Bird, G. A. 1 976. Molecular Gas Dynamics. Oxford: Clarendon Bird, G. A. 1978. Monte Carlo simulation of gas flows. Ann. Rev. Fluid Mech. 1 0:

ll-31

Ann

u. R

ev. F

luid

Mec

h. 1

989.

21:3

87-4

22. D

ownl

oade

d fr

om w

ww

.ann

ualr

evie

ws.

org

by W

est V

irgi

nia

Uni

vers

ity o

n 01

/22/

14. F

or p

erso

nal u

se o

nly.

Page 7: Rarefied Gas Dynamics

RAREFIED GAS DYNAMICS 393

Sherman (1969), Kogan (1973) , and Bird (1978). The books by Bird (1976), Cercignani (1975, 1988), and Vincenti & Kruger (1965) are also recom­mended. Since 1960, a detailed and comprehensive record of activity in the field is available from the published proceedings of the International Symposia on Rarefied Gas Dynamics (RGD 1-16 in Table 1). These proceedings are of high quality and represent refereed and invited articles that form a unique, concentrated reference source. As a reasonable rule of thumb, most work in the field since 1960 has appeared in some form in these proceedings. One unfortunate aspect is that there is no cumulative subject index for the RGD Proceedings. This problem has been recognized, and the International Advisory Committee will attempt to provide such a data base sometime after the upcoming RGD symposium (RGD 16, July 1988, Pasadena, California).

There exists no comprehensive recent review of the experimental tech­n iques of rarefied gas dynamics. Schaaf (1958) lis ts in his references (nos. 209-262) work on a number of experimental rarefied-gas-dynamics tech­niques. Sherman & Talbot (1960) present the status of comparisons between theory and measurement at that time. Sherman (1963) discusses early experimental techniques. The electron-beam fluorescence technique, which has provided many types of measurements in rarefied gas flows, is reviewed by Muntz (1969, 198Ia,b) and by Biitefisch & Vennemann (1974). To date, there has been no extensive application of laser techniques to the study of rarefied flows, although this subject is to be reviewed by Lewis in the proceedings of the upcoming RGD 16 (Lewis 1989). Hagena (1984) has reviewed cluster formation in expansions, and Fenn (1985) has dis­cussed the future (and the route to get there) for molecular beams. Thomas (1985) and Hurlbut (1985) have examined surface accommodation coefficient measurements. Measuremen ts of nonequilibrium processes in free-jet expansion flows have been reviewed by Rebrov (1977, 1985) .

THEORETICAL RAREFIED GAS DYNAMICS

There are many surveys of theoretical techniques in rarefied gas dynamics. A review of Cercignani (1975,1988), Kogan (1967,1973), Sherman (1969), Shidlovskii (1965), Grad (1958), and Schaaf (1958) provide a com­prehensive coverage. Works dealing specifically with discrete velocity models and Monte Carlo approaches arc brought up in more detail later. In this section I briefly look at theoretical techniques in order to provide a basis for later discussions.

The fundamental equation of the kinetic theory of gas flows, if only two-body collisions are important, is the Boltzmann equation:

Ann

u. R

ev. F

luid

Mec

h. 1

989.

21:3

87-4

22. D

ownl

oade

d fr

om w

ww

.ann

ualr

evie

ws.

org

by W

est V

irgi

nia

Uni

vers

ity o

n 01

/22/

14. F

or p

erso

nal u

se o

nly.

Page 8: Rarefied Gas Dynamics

394 MUNTZ

al +� a] = J(J,I) at } axj, Kn

(j= 1,2,3). (1)

Following Kogan's (1967, 1973) notation, Xj and 0 are nondimensional coordinates and speeds, respectively, of a molecule in thej-direction based on a characteristic length L and a characteristic speed [2(k/m)T]1/2. The local static temperature is T, k is Boltzmann's constant, and m is the molecular mass. The nonlinear collision integral J is complicated; it describes the net effect of populating and depopulating collisions on the distribution function. The distribution function 1 is the nondimensional molecular number density (based on a reference number density no) in a six-dimensional phase space formed from three velocity and three space coordinates. The Knudsen number Kn is the ratio of a molecular mean free path to the characteristic flow dimension. By multiplying the Boltzmann equation in turn by molecular mass, momentum, and energy and inte­

grating over all possible molecular velocities in each case, five conservation equations can be found for the transport of mass, momentum, and cncrgy.

As the average distance a molecule travels between collisions (its mean free path) becomes large compared with a characteristic flow dimension, the Knudsen number approaches infinity and the role of gas-phase col­lisions becomes unimportant. This limit is the free-molecule flow regime, where only gas-surface collisions between the gas and an obstacle or conduit determine the flow-field properties. In this regime the flow field can be determined even for complicated geometries by using computational techniques, which are limited only by the accuracy to which the surface­reflection characteristics are known.

For Kn --+ 0, gas-phase collisions become important and the flow approaches the continuum regime of conventional gas dynamics. In the limit of small Kn, the Boltzmann equation can be solved by expanding the distribution function as a series in Kn:

J= pO)+KnJ(I)+Kn2j<2)+ . . . . (2)

Substituting for J in Equation (I) and equating terms of equal order results in a set of integral equations (the Hilbert-Chapman-Enskog approach; Kogan 1973 ). Once the form of the distribution function is assumed, all of the hydrodynamic quantities including the stress terms and heat-flux vector can be written in terms of the distribution function, which can thus be specified in terms of flow quantities and their derivatives (cr. Kogan 1 973) .

The first term on the right-hand side of Equation (2) is the Maxwell or equilibrium distribution function, which is closely approximated if the flow is isentropic, and is given by

Ann

u. R

ev. F

luid

Mec

h. 1

989.

21:3

87-4

22. D

ownl

oade

d fr

om w

ww

.ann

ualr

evie

ws.

org

by W

est V

irgi

nia

Uni

vers

ity o

n 01

/22/

14. F

or p

erso

nal u

se o

nly.

Page 9: Rarefied Gas Dynamics

RAREFIED GAS DYNAMICS 395

� and u represent the nondimensional speeds of a molecule and of the flow, respectively. The first term in Equation (2) corresponds to the Euler equations. The second term corresponds to the Navier-Stokes equations, and the third to the Burnett equations. Later we wil l have occasion to use the form of Equation (2) that is consistent with the Navier-Stokes equations, namely

(4)

In Equation (4) the molecular speeds Cj and Ck are "peculiar" speeds, or speeds relative to the local flow speed, where Cj = 0-Sj' Sj = u) [2 (k/m)T] 1/2, and T is the local static temperature. The speed ratio Sj is the j-component of local flow velocity Uj' made nondimensional using the most probable molecular peculiar speed [2(k/m)Tf/2.

The distribution function of Equation (4) i s a local Maxwellian (i . e. isentropic flow) if the macroscopic gradient terms AO In T/oxj and AoS)oxk approach zero. Note that these terms are really the fractional change in temperature or change in speed ratio per molecular mean free path A. As is described by Holtz et al . (1971), limits on the size of these terms before the general form of Equation (4) becomes invalid can be estimated from comparison to experimental measurements of the distribution function.

For flows with large Kn, the term Kn-I can be used as a small parameter to expand the distribution function and substitute into Equation (l); a solution is subsequently obtained by equating terms of the same order from the left- and right-hand sides. This is the nearly free-molecule flow limit, where the distribution function is disturbed slightly from its collisionless value (see summary in Sherman 1969).

Much of the initial impetus for studies in rarefied gas dynamics was to be able to describe flows for arbitrary Kn, with particular emphasis on the transition regime between the continuum and free-molecule limits. There have been a variety of approaches developed to achieve this purpose. In addition to solutions based on expansions of the distribution function in terms of a small parameter (the Knudsen number for small Kn, or Kn-I

for large Kn near the free-molecule limit) or variants of this general approach (see Kogan 1967, Cercignani 1975), Equation (1) can be solved

Ann

u. R

ev. F

luid

Mec

h. 1

989.

21:3

87-4

22. D

ownl

oade

d fr

om w

ww

.ann

ualr

evie

ws.

org

by W

est V

irgi

nia

Uni

vers

ity o

n 01

/22/

14. F

or p

erso

nal u

se o

nly.

Page 10: Rarefied Gas Dynamics

396 MUNTZ

by linearizing the collision integral for flows where the average speed and temperature exhibit l ittle variation. This linearization has been extensively exploited by Cercignani (1975) to obtain solutions in the transitional-flow regime.

Another method for simplifying the collision integral can be found in models or intuitive forms. One popular model is the B GK equation (Bhatnagar et al. 1954), which is given by

aJ - aJ _ NN J(O) J ;1- + �i;;- - nv( - )/Kn. ut uXj

(5)

Here v is a dimensionless molecular-collision frequency that does not depend on molecular velocity but may depend.on static temperature. This model equation can be solved for many situations where a solution of the Boltzmann equation is not practical. An example is the shock-wave solu­tion of the B GK equation obtained by Liepmann et al. (1962) that was used to study thc characteristics of strong shock wavcs.

Another approach that has been used in specialized situations is the combination of two or more Maxwell distributions. Perhaps the most famous of these is the Mott-Smith assumption that the distribution func­tion in a shock wave is the weighted combination of the upstream and downstream Maxwell distributions (Mott-Smith 1951, Muckenfuss 1962). This rather simple assumption actually works very well for the prediction of shock-wave thickness over a wide range of shock Mach numbers (Alsmeyer 1 976).

COMPUTATIONAL RAREFIED GAS DYNAMICS

The dominant predictive tool in rarefied gas dynamics for the past decade at least has been the direct simulation Monte Carlo (DSMC) technique. This approach, which was introduced in 1963 and 1 964 by Graeme Bird (Bird 1963, 1965), has been developed, nurtured, and brought to an impressive level of ptoductive capability by Bird and others in the inter­vening years. The work reported by Moss ( 1 986) using the DSMC method to predict species concentrations and radiation in very energetic flow fields indicates thc power of the technique. The basic technique is described by Bird in h is book (Bird 1976); its use in chemicall y reacting flows has been presented by Koura ( 1 973) and Bird (1976, 1979). Recently, the technique has been extended to radiating flows (Bird 1 987). A short review of the DSM C technique, as well as the Hicks-Y en-N ordsieck method (N ordsieck & Hicks 1967, Yen 197 1 ) and the molecular-dynamics method, has been

Ann

u. R

ev. F

luid

Mec

h. 1

989.

21:3

87-4

22. D

ownl

oade

d fr

om w

ww

.ann

ualr

evie

ws.

org

by W

est V

irgi

nia

Uni

vers

ity o

n 01

/22/

14. F

or p

erso

nal u

se o

nly.

Page 11: Rarefied Gas Dynamics

RAREFIED GAS DYNAMICS 397

given by Bird (1978, 1 989; see also Yen 1984). The Hicks-Yen-Nordsieck (HYN) approach is a Monte Carlo method, but in this case the collision integral is solved by a Monte CarIo sampling technique; the remainder of the Boltzmann equation is solved using standard finite-difference methods. Appropriate implementations of the DSMC technique have been shown by Bird ( 1 976) and by Nanbu in a series of papers (Nanbu 1986) to be in principle an exact solution of the Boltzmann equation, although as Bird ( 1 989) argues, it is not entirely clear that such a connection is necessary.

In the DSMC method a large number of simulated molecules are fol­lowed simultaneously (Bird 1 976, 1989). Coll isions are handled on a prob­abilistic basis using the molecules found in a small geometric cell after each computational time step. The computation is started from some initial condition and followed step by step in time; steady flow is the condition that is reached at large times. The computational cell network is in physical space, and the time steps can be directly related to physical times. A most important feature of thc technique is that it can be applied so that the computation time is proportional to the first power of the number of simulated molecules.

The DSMC technique is a pure form of computational fluid dynamics. In principle it can contain all of the physics needed for any problem without the necessity of non equilibrium thermodynam ic assumptions that are required in nonequilibrium continuum-flow calculations. In practice the technique is computationally intensive compared with its continuum counterparts. However, calculations that overlap continuum calculations (with added slip effects) for STS orbiter flow fields have been accomplished (Moss 1986). The success of the technique depends on computational performance, so it is easy to anticipate further advances in its use. A number of research centers are now looking at efficiently matching DSMC and continuum techniques in a computational hybrid approach (Cheng & Wong 1 988). As indicated by Bird (1985), hypersonic flow fields can show widely different degrees of rarefaction at different locations, which makes a hybrid approach attractive for such flow fields. On the same subject, it was noted by Yen ( 1 984) that the HYN approach offers the possibility of not having to match d ifferent numerical techniques for solving mixed continuum and rarefied-flow problems.

Since in many situations where extreme nonequil ibrium effects exist, such as in shock transition zones around sharp hypersonic leading edges, the DSMC technique is the only realistic method for obtaining solutions, its validity is important. In some sense its validation may well serve as a prototype for many other computational fluid-dynamic techniques. The usual situation, and the one that applies here, is that for flow fields in which one really wants to use the computational technique, it i s very

Ann

u. R

ev. F

luid

Mec

h. 1

989.

21:3

87-4

22. D

ownl

oade

d fr

om w

ww

.ann

ualr

evie

ws.

org

by W

est V

irgi

nia

Uni

vers

ity o

n 01

/22/

14. F

or p

erso

nal u

se o

nly.

Page 12: Rarefied Gas Dynamics

398 MUNTZ

difficult to check it directly. The status of DSMC experimental validation is reviewed in a following section.

The success of the DSMC technique, as well as a continued desire to extend calculations to lower Knudsen numbers, has revived interest in discrete-velocity gas models. In these models, gases are allowed to have molecular velocities with only a l imited set of possible values. It is sur­prising how successful the very simplest velocity model of a gas, frequently presented in introductory chapters of kinetic-theory texts, can be in pre­dicting approximate transport properties and the equation of state. This simplest model permits molecular motion only parallel and antiparallel to the coordinate axes of a Cartesian system. Following the suggestions of Krook (1955) and Gross (1960), Broadwell (1964a,b) has extended versions of the simple model to solve Couette and Rayleigh flows and shock-wave structure. Broadwell's use of a very few discrete velocities permitted him to solve analytically both the infinite-Mach-number shock-wave transition problem and translational Couette flow for low Mach number and arbi­trary Knudsen number. Hamel & Wachman (1965) have used what they call the "discrete ordinate" technique for a study of Couette flow.

For discrete models, v Boltzmann equations can be written for v discrete velocities:

(7)

The rate of change of the number of molecules N(v) with velocity (v) on the left-hand side of Equation (7) is just equal to the difference between gain (G) and loss (L) terms on the right-hand side. There is clearly a set ofv differential equations that can be written where the "collision integral" (G - L) will be very simple if v is a small number (Broadwell 1 964b) . The surprising qualitative and even quantitative accuracy of the discrete­velocity models is related to the question expressed recently by Hasslacher (1987): "How detailed must micromechanics be to generate the qualitative behavior predicted by the Navier-Stokes equations?" A related question is, is there a use for discrete models to help accelerate DSMC-type cal­culations? If there is, a h ybrid DSMC-discrete-velocity approach would permit calculations to smaller Knudsen numbers, which is a worthwhile objective.

The discrete-velocity model of Broadwell retains the notion of con­tinuous space and time. Other models go beyond the discrete-velocity assumption to discrete physical space and time. As reviewed by Hasslacher (1987), the particles exist at nodes in a lattice; in each time step a particle will move in one of the possible discrete directions a d istance of exactly

Ann

u. R

ev. F

luid

Mec

h. 1

989.

21:3

87-4

22. D

ownl

oade

d fr

om w

ww

.ann

ualr

evie

ws.

org

by W

est V

irgi

nia

Uni

vers

ity o

n 01

/22/

14. F

or p

erso

nal u

se o

nly.

Page 13: Rarefied Gas Dynamics

RAREFIED GAS DYNAMICS 399

one node. Simple conservation collision rules are implemented at each node to determine the discrete particle velocities for the next step. In all cases the particle speeds are constant. Thus at the beginning of a time step only the particle velocities in the nearest neighbor nodes need be known in order to completely determine the results of the time step. Simple, massively parallel computations are possible. With appropriate choices of lattice configuration and collision rules, cellular spaces, or cellular au­tomata as these models are called, have been observed to evolve at least approximately the dynamics of nonlinear physical systems. The connection between the lattice gas and real gases (as represented by the Navier-Stokes equations) can be formed using analogs of the kinetic-theory techniques developed for obtaining transport coefficients from the Boltzmann equa­tion (Frisch et al. 1986). Results of lattice-gas calculations have provided remarkable, apparently qualitatively reasonable continuum flow field visualization (Hasslacher 1987). The quantitative accuracy of lattice-gas simulati on of continuum-gas-dynamics phenomena remains to be estab­lished, but the application or adaptation of these new directions to mas­sively parallel rarefied-flow computations are of great interest. The single­speed lattice gas has no way to have a gas temperature independent of gas velocity. To remedy this, a multispeed lattice gas is required and has been proposed by d'Humieres & Lallemand (1 986). Calculations based on this model will be reported at RGD 16 (Nadiga et al. 1989). It is clear that the computational efficiency of at least the simplest collision-rule automaton can be extremely attractive. The inclusion of more complicated rules to accommodate real gas effects (rotation, vibration, chemistry, etc.) is another matter; however, there is no question that the automata offer intriguing possibilities and certainly will be the subject of much future research.

EXPERIMENTAL RAREFIED GAS DYNAMICS

The history of experiments in the field of rarefied gas dynamics involves two rather different mainstreams that have, however, interacted in mutually beneficial ways. Special effort has been expended on developing techniques for producing and measuring rarefied flow fields. Determination of the details of the molecular motion and the populations of internal energy states has been a particular emphasis and has led to the development of several new techniques. In parallel, much effort has gone into the development of supersonic free-jet molecular-beam sources. This work was originally done for surface-interaction studies but has now been extensively used in chemical kinetics and a wide range of related investigations. The development of sources in the period up to about 1975 leaned heavily

Ann

u. R

ev. F

luid

Mec

h. 1

989.

21:3

87-4

22. D

ownl

oade

d fr

om w

ww

.ann

ualr

evie

ws.

org

by W

est V

irgi

nia

Uni

vers

ity o

n 01

/22/

14. F

or p

erso

nal u

se o

nly.

Page 14: Rarefied Gas Dynamics

400 MUNTZ

on flow-field instrumentation techniques that could observe energy-state populations and molecular motion in various gas species. In return, the simple free-jet flows were an ideal test-bed for fl ow-field instrumentation techniques. The two parallel experimental developments are a remarkable success story. Spectacular progress has been made in molecular beams and flow cooling for both chemical kinetics and surface studies. Detailed flow­field measurements were made a number of years ago and have been used very recently for comparison with DSMC calculations.

Instrumentation Techniques for Flow Fields The requirements for understanding rarefied flows stimulated the devel­opment of special instrumentation techniques, which were developed to sense the rotational, vibrational, and translational energy-state population distributions at points in rarefied flow fields.

The first flow-field instrument to indicate the directional details of the molecular motion in a gas flow was the free-molecule pressure probe developed by G. N. Patterson and his students (Patterson 1 956, Enkenhus et al. 1960). This device is a cylindrical probe with a small orifice in its side

wall. It is used in flows where the mean free path is much greater than the probe diameter. By placing the probe's axis in arbitrary directions and rotating the probe about its longitudinal axis, the plane of the probe's orifice can in principle be oriented in any direction. Pressure measurements as a function of orifice orientation provide information about the molec­ular motion as a function of direction. There have been many variants of this probe, including one in which the probe was used to feed a mass

spectrometer for the study of mixture flows (Koppenwallner 1986). The long time response of very small probes attached to finite gauge volumes and the difficulty of probe manufacture have l im ited the use of free­molecule pressure probes to flows that are at lower dens ity than is con­venient for many laboratory fl ow studies.

A second-generation flow-field probe is the electron-beam fluorescence technique. A collimated beam of, say, 20-kV electrons is projected through a rarefied flow. The electrons stimulate fluorescence from the flow gases, which can then be used to measure under various particular circumstances the translational, rotational, and vibrational population distributions of the gases in the flow. First introduced by Schumacher & Gadamer (1 958) for density measurements, the technique was later developed for trans­lational, rotational, and vibrational population distributions by Muntz and many other investigators (Muntz 1969, 1981a,b). Spatial resolution is obtained by observing the fluorescence at posit ions along the beam's trajectory and maintaining a beam diameter of around 1 mm or so. A wide

Ann

u. R

ev. F

luid

Mec

h. 1

989.

21:3

87-4

22. D

ownl

oade

d fr

om w

ww

.ann

ualr

evie

ws.

org

by W

est V

irgi

nia

Uni

vers

ity o

n 01

/22/

14. F

or p

erso

nal u

se o

nly.

Page 15: Rarefied Gas Dynamics

RAREFIED GAS DYNAMICS 401

variety of studies have been reported using electron-beam fluorescence measurements (Muntz 1 969, 1981a,b, Butefisch & Vennemann 1 974) .

A lmost all of the measurements that have recently been developed for optically accessible molecules using laser-induced fluorescence had been made 20 years ago using electron-beam fluorescence. Laser-induced fluo­rescence can operate at higher densities using saturation optical pulses and is more sensitive on a per-molecule basis. However, seeding the flow with optically accessible atoms or molecules i s frequently necessary. In rarefied gas flow it is not convenient to use seed molecules for use with laser­induced fluorescence, since they may not follow the flow that is being traced owing to mass differences; additionally, their internal energy-state population distributions will not reflect the flow molecules in a well­understood way. An alternative would be to study pure flows of, say, iodine or sodium vapor at relatively high pressures. This method presents its own set of problems.

Flow-field measurements over the next decade will likely continue in the tradition of electron-beam fluorescence, but with a difference. As discussed by Muntz et al. (1987) there are real advantages to be obtained by com­bining electron-beam- or X-ray-induced excitation with laser-induced fluorescence. The electron or X-ray beam acts to seed optically accessible species along its trajectory. A crossing laser beam can then probe these species. If great care is taken to follow the excitation kinetics, it is believed that a flow-measurement technique can be developed that will exhibit the best features o f both electron-beam and laser-induced fluorescence. I f optically accessible molecules are being studied, laser-induced fluorescence offers a number of advantages. It will certainly be applied on i ts own in future rarefied-gas-dynamics studies.

Gas-Dynamic Molecular Beams and Free Jets

The suggestion by Kantrowitz & Grey (1 95 1) that intense molecular beams be formed using supersonic or hypersonic flows, from which a high-fluence beam is skimmed, has resulted in a large body of research in the area of free-jet expansions and molecular beams. A convenient source for hypersonic gas was found to be the free jet (Ashkenas & Sherman 1965, French 1 966). Research on both the characteristics of free-jet flows and on the formation of intense molecular beams has been prolific. Freezing of the translational random motion in the direction of the expanding flow (the parallel "temperature") was identified by Anderson & Fenn (1965) using molecular-beam sampling techniques and time-of-flight analysis of the parallel molecular velocity d istribution in the molecular beams pro­duced by the sampling process. The freezing was subsequently predicted

Ann

u. R

ev. F

luid

Mec

h. 1

989.

21:3

87-4

22. D

ownl

oade

d fr

om w

ww

.ann

ualr

evie

ws.

org

by W

est V

irgi

nia

Uni

vers

ity o

n 01

/22/

14. F

or p

erso

nal u

se o

nly.

Page 16: Rarefied Gas Dynamics

402 MUNTZ

by Hamel & Willis (1966) and Edwards & Cheng (1966). The transition region from equil ibrium to frozen translational temperature in jet expan­sions was observed initially by Muntz (1967) and later more com­prehensively by Cattolica et al. (1974). The internal-state relaxation in jets was initially studied by Marrone (1967) and M iller & Andres (1967). There have been a large number of subsequent publications on this subject, and these are perhaps best summarized by Rebrov (1977, 1985) and by Fenn ( 1 985).

Because of their convenience, free jets are the method of choice for generating gas-dynamic molecular beams. Flow-field measurement tech­niques, such as the electron-beam fluorescence probe and the molecular beams obtained by skimming free-jet flows, have been used to study jet flow properties (Anderson & Fenn 1965). After considerable research by many investigators, the art of flow sampling has been refined to such an extent that very l ittle interference is introduced by the sampling process (cf. Bossel 1 971) . Because of this capability, extremely low temperatures can be measured in the molecular beams sampled from the jet flows.

Free-jet expansions have been used extensively as refrigerators. Fenn (1985), in a wonderful review of his long affair with jets, discusses several uses of jet cooling. By taking advantage of large increases in coll ision cross sections at very low temperatures, groups led by Campargue (Campargue & Lebehot 1974, Campargue et al. 1981) and by Toennies (Miller et al. 1974) have managed to avoid translational freezing and to generate extremely low temperature (milli-Kelvin) molecular beams. The low tem­peratures permit surface-reflection studies that can sense single-phonon interactions of, say, helium w ith a surface (Brusdeylins et al. 1981). These extremely monoenergetic beams have also been used in inelastic gas­scattering experiments (Faubel et al. 1980).

Cooling with the jet gas used as a bath makes it possible to depopulate the majority of the rotational and vibrational levels of complicated mol­ecules. Indeed, owing to their low density in the bath, the molecules can be cooled well below their normal condensation temperatures. This effect has spawned a remarkable quantity of recent papers on laser-induced­fluorescence molecular spectroscopy, as described by Fenn (1985).

The ubiquitous free-jet expansion into a h igh vacuum has been used extensively for generating controlled agglomerations of molecules in the condensed phase. The ability to generate molecular groups ranging from dimers to aerosol-size particles allows technologically important phenomena to be studied (Fenn 1985). Cluster formation in jets has been reviewed by Hagena (1984) and by Rebrov (1985) as part of a review of nonequilibrium processes. The subject is wide ranging and interesting but cannot be dis­cussed here in detail owing to space l im itations. The many scientific studies

Ann

u. R

ev. F

luid

Mec

h. 1

989.

21:3

87-4

22. D

ownl

oade

d fr

om w

ww

.ann

ualr

evie

ws.

org

by W

est V

irgi

nia

Uni

vers

ity o

n 01

/22/

14. F

or p

erso

nal u

se o

nly.

Page 17: Rarefied Gas Dynamics

RAREFIED GAS DYNAMICS 403

made possible by a convenient source of clusters indicate that this subject will continue to attract attention in the future.

ASSESSING COMPUTATIONAL RAREFIED GAS DYNAMICS

Because of its central importance in predicting rarefied flows, the validation of the DSMC technique by experiments is discussed in this section. The status of the validation is important for simple gases, for chemicall y react­ing gases, and for gases with internal degrees of freedom. The DSMC technique, or some variant, can in principle account for all of the correct physics during a collision. However, to do so implies an extreme com­putational overhead. The question raised earlier about how much micro­knowledge has to be included to arrive at a macroscopically satisfactory result needs to be answered; however, note that there is no one answer, since the level at which macroscopic satisfaction is achieved will depend on the application. Comparing DSMC results with experiments can achieve this goal. As an illustration of the possibilities, two flow fields are examined here in order to present the status of comparisons between DSMC results, other predictions, and measurements. Additionally, it is im­portant to validate the DSMC method because of the typical computa­tional-fluid-dynamics role it has assumed as a surrogate for experiment.

Before looking at specific flow fields it is useful to remember that molec­ular interactions are generally modeled by point centers of repulsion in computational rarefied gas dynamics. In this model the intermolecular force during a collision is assumed to be inversely proportional to a power 1] of the distance between the molecules. Based on the Chapman-Enskog theory described earlier, the coefficient of viscosity can be related to the intermolecular force law by {t = K\Tw, where w = (1]+3)/2(1]-1 ) and K\ is a constant given by kinetic theory and the characteristics of the molecule. The mean free path of a molecule can be calculated from its coefficient of viscosity and is thus directly related to w or 1]. In the DSMC technique a collision cross section is required. As discussed by Bird ( 1 976), an effective hard-sphere momentum-transfer cross section that varies with the relative kinetic energy of the collision partners can be written in terms of w or 1]. Calculations made for different theories or models can be compared by using the same intermolecular force law to provide either viscosity coefficients or collision cross sections as necessary.

Normal Shock Waves There is now a considerable body of data on the density profiles of normal shock waves in monatomic gases for a wide range of shock Mach numbers.

Ann

u. R

ev. F

luid

Mec

h. 1

989.

21:3

87-4

22. D

ownl

oade

d fr

om w

ww

.ann

ualr

evie

ws.

org

by W

est V

irgi

nia

Uni

vers

ity o

n 01

/22/

14. F

or p

erso

nal u

se o

nly.

Page 18: Rarefied Gas Dynamics

404 MUNTZ

Most of the available data are reviewed by Alsmeyer (1976), who also presents important new data with very little relative scatter. Other data for very low shock Mach numbers have been presented by Garen et al. ( 1 977). These two studies were done in shock tubes; one used observations of the attenuation of an electron beam to measure the density profiles, the other a laser interferometric technique. Argon and xenon were the gases used in the studies. We first consider Figure 2, where various quantities are defined that relate to a shock-wave density profile. It is generally convenient to use the nondimensional density fi as a position variable in the shock and to nondimensionalize distances by the free-stream mean free path. Using only Alsmeyer's ( 1 976) data for maximum-slope shock thickness, which really lies in the middle of a rather wide scatter band from many other investigations based on a variety of techniques, the comparison shown in Figure 3 (due to Fiscko & Chapman 1 988) between experiment and prediction with 11 = 10 (w = 0.72) is consistent with the temperature sensitivity of measured argon viscosities at temperatures above 3 00 K (appropriate for shock-tube experiments). The trend for the DSMC tech­nique to predict a thinner shock at low Mach numbers was also noted by Alsmeyer ( 1 976). He ascribed this tendency to the rather significant attractive well of argon (which of course is not modeled by a simple inverse-power repulsion) playing a more important role in the low-Mach­number shocks. Sturtevant & Steinhilper ( 1 974) performed DSMC cal­culations incorporating various attractive-well depths but at higher Mach numbers, finding smaller thicknesses in terms of the upstream mean free path for decreased attractive-well depths.

It is interesting to note that Alsmeyer's data are fit quite well by the M ott-Smith model as developed by Muckenfuss ( 1 962). The fit is best at the lower Mach numbers. A similar satisfactory fit was found recently by Velikodnyi et al. ( 1 985), who applied Struminskii's ( 1 980) method of

>- n2 I-en z w o a: w CD � Z

FLOW -

Q=� A2

SHOCK WAVE NUMBER DENSITY PROFILE

o L-______ -L __ L- ______________ � o

Figure 2 Characteristic parameters that can be derived from a shock-wave density profile. The maximum-slope shock thickness is.5" and AI is the upstream mean free path.

Ann

u. R

ev. F

luid

Mec

h. 1

989.

21:3

87-4

22. D

ownl

oade

d fr

om w

ww

.ann

ualr

evie

ws.

org

by W

est V

irgi

nia

Uni

vers

ity o

n 01

/22/

14. F

or p

erso

nal u

se o

nly.

Page 19: Rarefied Gas Dynamics

0.4

0.3 A- DSMC

b b

/---. ......... � / CURVE THOU:::':' " ' :..:.. ;- - -

ALSMEYER'S DATA. ARGON , T1 " 300 K

MOTT - SMITH, w = 0.72

o 1!-L--±-3.....l.----;!5,------lL...,7!;--L--±-9.....l.�11 SHOCK MACH NUMBER

RAREFIED GAS DYNAMICS 405

Figure 3 Comparison of experiment and predictions for argon normal shock waves (after Chapman et al. 1987).

systems of identical particles to the shock-structure problem. The Burnett equations, which were once thought to have no solution above M � 1.9, were integrated in 1 976 to M = 4 by Simon & Foch ( 1 977) using modern numerical techniques for the case of a normal shock with w = 1 . Very recently, Fiscko & Chapman ( 1988) have integrated the Burnett equations (less one small but numerically troublesome term) to M ::::: 35. The shock­thickness predictions by F iscko & Chapman are in quite good agreement with DSMC calculations for the case of OJ = 1 .

A convenient single-parameter measure of the shape of a shock profile is the quotient Q defined in Figure 2 (Schmidt 1 969). The plot in Figure 4 shows Q as a function of shock Mach number as calculated by S imon & Foch ( 1977) with OJ = 1 for the Navier-Stokes and Burnett equations. Superimposed are the results of Oaren et al. (1977). The data seem to support the Burnett results above M � 2.0 (however, remember that OJ = 1

f.20

1.15 GAREN .f 01 DATA. xENON T,= 300K

BURNETT--7" ....... -w . f ,, _ or

o // / ' , _ . ' . ' 4.05 // ,

If' ,/ , I ,/ 1.00 / '

___ ALSMEYER

I " ,.,.-DATA. ARGON I Y T ' 300K J '/ 4

I :

2 3 4 5 SHOCK MACH NUMBER

Figure 4 Comparisons of the quotient Q (see Figure 2) for Navier-Stokes and Bur­nett solutions to normal shock waves with two sets of experiments.

Ann

u. R

ev. F

luid

Mec

h. 1

989.

21:3

87-4

22. D

ownl

oade

d fr

om w

ww

.ann

ualr

evie

ws.

org

by W

est V

irgi

nia

Uni

vers

ity o

n 01

/22/

14. F

or p

erso

nal u

se o

nly.

Page 20: Rarefied Gas Dynamics

406 MUNTZ

in the predictions). Between M = 1.0 and 2 .0 the data do not really favor either version. I am not aware of Q-values having been presented for low­Mach-number DSMC calculations, although accurate values could surely be obtained with present techniques and computers (cf. Bird 1989).

So far we have looked only at density, which i s a very simple moment of the molecular-velocity distribution function in a shock wave. It is of interest to see what level of complexity in the description of collisions is required for the DSMC or other techniques to be able to predict the distribution function in a shock wave. M easurements are available from electron-beam fluorescence studies done a number of years ago in helium at M = 1.59 and M = 20 (Muntz & Harnett 1969; E. P. Muntz, unpub­lished work, 1968) and in argon at M = 7 (Holtz & Muntz 1984). For low­Mach-number shock waves, a convenient single-parameter assessment of the shape of the molecular-velocity distribution function is its full width at half-maximum, i .e. its half-width (Holtz et a1. 1971). Data for the half­widths of the distribution for molecular motion in the direction of the flow (parallel) and perpendicular to it (perpendicular) are shown in Figure 5 as a function of non dimensional density rise fi in the M = 1.59 helium shock. For comparison, plots are also shown of the Navier-Stokes half-widths [using Equation (4) for the distribution function], M ott-Smith half-widths, and DSMC half-widths, all with w = 0.647. Note that in this case, the DSMC and HYN techniques do a good j ob of matching the data, whereas the Chapman-Enskog distribution is quite poor. The M ott-Smith dis-

1 .2 r-r-'-.,...,�'-,..,.-'-"""--,,,,,,,,,,.-r-,-,......,--',-,,,,,-,

1.1

<;; 1.0 :i 0.9 .... e 0.8 ;;

..:. 0.7 ...J � 0.6 (f) 0.5 (f)

':J 0.4 z � 0.3 z � 0.2 a 0.1

o

W = W - W1

� .or:..

�_

w2 - w, y .....

.

W : half width

{

.Y,.:.<

--

-

-

-

4-

;r;

1 - Upstream t/.:,.-'- ;?p 2 - Downstreom k/ ./ . . ;.:'

I ....... ;tjr / ..... /.:

/ ..... f ,/.�� Experiment

/ /. ,::/' Muntz B Harnett

/ .... / /;/' <} .L d istribution ;.r . . ;§ { /1 distribution I .. ):/ .. ' f .

.. )Y" . . . . ......... ... Mott - Smith II

fi .. . ' ... - . , . Moll -Smith .L I ,:/' - - - Chopmo n - Enskog

I ; :t. - -- - .. HYN, w : 0.5 .. 'I' - - '-- HYN , w : 1.0

-- DSMC II , w :O.647 - - - - - - - - DSMC .L w :0.647

-0.1 L..J......L.J.......l.-'--...L....L..l.-'-l.....t.....l....-'-'---'-....L-<-...l.-,-..L....Jc....J o 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.6 0.9 1.0

R EDUCED NUMBER DENSITY, � Figure 5 Half-widths of distribution functions for an M = 1 .59 normal shock wave.

Ann

u. R

ev. F

luid

Mec

h. 1

989.

21:3

87-4

22. D

ownl

oade

d fr

om w

ww

.ann

ualr

evie

ws.

org

by W

est V

irgi

nia

Uni

vers

ity o

n 01

/22/

14. F

or p

erso

nal u

se o

nly.

Page 21: Rarefied Gas Dynamics

RAREFIED GAS DYNAMICS 407

tribution is nearly as successful as the DSMC results. Further results of a similar nature, including comparisons to M = 7 argon data, will be pre­sented soon (Erwin & Pham Van Diep 1 989). As an example, the Mott­Smith parallel distribution function for an M = 20 helium shock is com­pared with experimental results (E. P. Muntz, unpublished work, 1 968) at fi = 0.62 in Figure 6. Note that while the Mott-Smith shock thickness is in reasonable agreement with experiment (Alsmeyer 1 976), the distribution function is not, exhibiting a large population of scattered molecules un­accounted for in the Mott-Smith model. The distributions in Figure 6 have each been normalized by their maximum values in order to compare the shapes of the two functions.

The preceding detailed comments are presented here as an example of one of the opportunities that exist for establishing true-bench-mark validations of computational rarefied gas dynamics. Similar studies of mixture shocks and shock waves in gases with internal degrees of freedom in order to study the details of inelastic collisions would be extremely useful. In both of these cases, measurements beyond those already avail­able may be necessary. Space limitations do not permit further discussion on these subjects here, although both arc extremely important for the complete validation of computational rarefied gas dynamics.

Flow-Field Studies and Knudsen Layers

The prediction of rarefied flow fields about simple shapes with detailed comparisons to experiments remains to be completed. The issue here is essentially how the kinetic layer near the interface between gas and solid (the Knudsen layer) is treated. Kogan ( 1 973, 1 985) and Cercignani ( 1 978,

1 .0

0.9 z � 0.8 3 0.7 ::::> g, 0.6 a.

0.5 w � 0.4 :5 0.3 w 0:: 0.2

Mp = 20

;; = 0.62

x EXPERIMENT. Fp

-- PREDICTION. MOTT-SMITH

0.1

O �����WL�����LL����LLLL�� 2.0 1.5 1.0 0.5 0 -0.5

(Veil X W- 5 ( cm/s )

Figure 6 Comparison of the shapes of parallel distribution functions at a point in an

M = 20 helium shock wave.

Ann

u. R

ev. F

luid

Mec

h. 1

989.

21:3

87-4

22. D

ownl

oade

d fr

om w

ww

.ann

ualr

evie

ws.

org

by W

est V

irgi

nia

Uni

vers

ity o

n 01

/22/

14. F

or p

erso

nal u

se o

nly.

Page 22: Rarefied Gas Dynamics

408 MUNTZ

1 979) have addressed the problem from a theoretical point of view. Cheng ( 1 988) describes the near-continuum rarefaction problem as shock slip and surface slip. Shock slip is the interaction between a thick shock wave and a fully viscous layer between the shock and the body. Surface slip refers to the jump conditions that occur in the kinetic layer (a thickness of order one mean free path) at the interface between gas and surface.

Crawford & Vogenitz ( 1 974) presented detailed DSMC flow-field cal­culations for a cylinder. These results included perpendicular and parallel temperature predictions. Davis & Harvey ( 1 979) have studied a high­Mach-number nitrogen flow over a flat plate (actually a large-diameter hollow cylinder aligned with the flow) and a plate with a forward facing step. The latter authors compared experiments with DSMC calculations by using several energy-transfer models to account for rotational energy transfer in collisions. Crusciel & Pool ( 1 986) have studied the Knudsen layer and other flow-field details for flows over a sphere and a sharp cone (Crusciel & Pool 1 987). In the latter case the flow-field density measure­ments of Hickman ( 1 967) were compared with the predictions. Measure­ments by Krylov et al. ( 1 985) of surface stress using a floating element balance are available for comparison to predictions. Molodtsov & Ryabov ( 1 985) compare measurements of flow around a sphere with DSMC and Navier-Stokes calculations. The Knudsen layer has been probed by Gottes­diener ( 1 984) with a free-molecule wire probe; this experiment is one of the few to resolve profiles in the Knudsen layer. Earlier velocity-profile measurements in a Knudsen layer have been presented by Reynolds et al. ( 1 974). Electron-beam fluorescence density measurements have been compared by Allegre et al . ( 1 986) with Navier-Stokes solutions with slip boundary conditions for a NACA 001 2 airfoil in an M = 2 flow. Velocity­distribution function measurements in the flow field over a flat plate made by Becker et al. ( 1 974) have been compared with DSMC predictions by Hermina ( 1 987) and Hurlbut (1 987).

It is clearly time for a detailed review of flow-field comparisons between experiments and predictions, as well as of the related issue of the Knudsen­number limits of continuum-flow approaches that include slip. Space limitations do not permit such a review here.

RADIATING, REACTING, NONEQUILIBRIUM COMPUTATIONAL RAREFIED GAS DYNAMICS

As noted in the introduction there is a growing need to be able to predict highly nonequilibrium, radiating and reacting flow fields. A major triumph of the past 5 to 10 years has been the development of the DSMC technique to provide this capability. The DSMC, or any similar particle treatment

Ann

u. R

ev. F

luid

Mec

h. 1

989.

21:3

87-4

22. D

ownl

oade

d fr

om w

ww

.ann

ualr

evie

ws.

org

by W

est V

irgi

nia

Uni

vers

ity o

n 01

/22/

14. F

or p

erso

nal u

se o

nly.

Page 23: Rarefied Gas Dynamics

RAREFIED GAS DYNAMICS 409

of a gas where classes of particles are followed (in this case only in a statistical sense), intrinsically is able to provide a representative reflection of nature and can be correct in highly nonequil ibrium situations. This of course is only true if all important elastic and inelastic collision cross sections are known as a function of energy. However, the possibil ity is there, whereas in continuum treatments reliance must be made on the assumptions of nonequilibrium thermodynamics. In the past decade, Graeme B ird and his associates at the Langley Research Center have been able to develop in "an engineering context" (Bird 1 985) the DSMC techn ique so as to provide predictions for h ighly nonequilibrium real air flow fields including radiation (Moss et al. 1 988). The "engineering context" refers to the numerous necessary approximations that must be made in describing the collision processes and energy transfers during collis ions. The com­putational resources and physical information needed to avoid these approximations are lacking at this time. The exciting point is that the engineering calculations are possible and provide reasonable agreement with the very few experimental results that are available (Bird 1 987). It is also clear that the agreement may be fortuitous, since there are a very large number of interacting approximations and not very well known collision cross sections involved.

The possib il ity of making flow-field predictions for multidimensional, reacting and radiating flows (also of course viscous) with the DSMC technique impl ies a clear direction for future work. Significant effort in rarefied gas dynamics will be aimed at developing inelastic coll ision models and establishing their validity in situations that are less complicated than a nonequilibrium air chemistry flow field. As an example, the standard way of treating inelastic collisions with transfer of rotational energy in the DSMC technique is based on the model of Borgnakke & Larsen ( 1 975). In the model a certain fraction of collisions are considered inelastic; for these collisions, new translational and rotational energies are sampled from the distribution of these quantities that would occur in an equilibrium gas with a spcci fic cnergy that is the same as the specific energy available in the collisions. This of course is not very realistic, but if the fraction of inelastic collisions is chosen so that the rotational relaxation rate in the gas is matched, it seems to work quite satisfactorily in an "engineering context." Vibration and electron excitation have been handled by analo­gous techniques. Experiments can be done quite easily to investigate models of, say, rotational energy transfer [for instance, rotational-level population distributions in a shock wave are already available (Robben & Talbot 1 966)]. It is to be expected that different inelastic collision models (e.g. Anderson et al. 1 986) will appear, and that experimental studies in simple situations will be undertaken to validate their predictions.

Ann

u. R

ev. F

luid

Mec

h. 1

989.

21:3

87-4

22. D

ownl

oade

d fr

om w

ww

.ann

ualr

evie

ws.

org

by W

est V

irgi

nia

Uni

vers

ity o

n 01

/22/

14. F

or p

erso

nal u

se o

nly.

Page 24: Rarefied Gas Dynamics

410 MUNTZ

In summary, because of the DSMC technique's potential importance in many scientific and technological fields (e.g. astrophysics, radiation trans­fer, exhaust gas radiation) the details of i ts application to reacting, radi­ating flow fields should be studied and improved where appropriate in order to construct a sound foundation for the technique's predictive capa­bilities. I t is likely that more detailed knowledge can be included in the future owing to increasing computational performance. The necessary research will be an active area in rarefied gas dynamics in the next decade (for an example, see Kunc 1 989).

BRIEF DISCUSSIONS OF SOME INTERESTING SUBJECTS

Evaporation- and Condensation-Induced Flows

In 1970 Bochkaryov et al. ( 1 970) reported on the study of radially expand­ing flows from cylindrical and spherical sources constructed from per­forated sheets. A similar, although plane, source was used by Y trehus ( 1977) to simulate an evaporating surface. Considerable attention has been paid to the evaporation-condensation problem because of its technological importance and i ts intrinsic interest as a problem in rarefied gas dynamics. A Knudsen layer is at the interface between the surface and the vapor for the case of evaporation or condensation. Condensation or evaporation can be described as strong or weak (Kogan 1985). Strong conditions are identified with sufficient vapor mass flow such that, to leading order in the approximation at the outer edge of the Knudsen layer, the Navier-Stokes terms can be neglected; a drifting Maxwellill;n and a finite flow velocity outside of the Knudsen layer are appropriate. If, at the edge of the Knudsen layer, Navier-Stokes terms are present in the approximation, then the flow is considered weak. While the evaporation and condensation Knudsen layers are symmetric and seem theoretically well understood for weak evaporation and condensation, this is not the case for strong condensation layers. Hatakeyama and Oguchi have identified a subsonic-supersonic bifurcation (Hatakeyama & Oguchi 1979, Oguchi 1 9 8 1 ) that involves conditions in the flow outside of the Knudsen layer. Kogan ( 1 985), in his RGD 1 3 review of transport processes in Knudsen layers, outlines the theoretical status as of 1 982. Also at the same symposium (RGD 1 3), Schilder et al. ( 1985) reported on the first known measurements in an evaporative Knudsen layer, which in this case was produced by intensive evaporation of iodine. They used laser-induced fluorescence to measure fluorescence emission-line profiles from which temperatures parallel and perpendicular to the evaporating flow were determined. Further theoretical studies of the subsonic-supersonic transition in strong condensation have

Ann

u. R

ev. F

luid

Mec

h. 1

989.

21:3

87-4

22. D

ownl

oade

d fr

om w

ww

.ann

ualr

evie

ws.

org

by W

est V

irgi

nia

Uni

vers

ity o

n 01

/22/

14. F

or p

erso

nal u

se o

nly.

Page 25: Rarefied Gas Dynamics

RAREFIED GAS DYNAMICS 41 1

been presented by Hatakeyama (1 984), who expands on the properties of the subsonic and supersonic branches. Ytrehus & Aukrust (1 984) also discuss flow breakdown of condensation Knudsen layers. Unsteady strong condensation has been analyzed by Sone et al. ( 1986) with emphasis on the establishment of a steady-state flow.

Two situations of practical interest for space applications-the evapora­tion of spherical and cylindrical sources into a high vacuum-have been studied by Edwards & Collins (1 969) and by Knight (1 976). Based on the work of Fuchs & Legge (1 979), Muntz et al. (1 984), Muntz & Orme (1 987), Anders & Frohn (1 984), and Faubel (1 989), it has become clear how to investigate evaporating liquid surfaces in a vacuum. It is expected that a number of studies concerning the characteristics of liquid surface evap­oration in vacuum will appear over the next few years. Detailed studies of the associated vapor flow fields are possible, with some initial work already reported (Takens et ai. 1984, Faubel 1989). Surface Interactions at Medium Energies

The study of surface interactions at low energies (up to say I eV) has long been of interest to the rarefied-gas-dynamics community. Unfortunately, there is insufficient space in the present review to address this area. New emphasis, however, has been given to surface interactions at intermediate energies (say 1 to 20 eV) as a result of the observations of glow and material erosion in low Earth orbit, as mentioned in the Introduction. Even prior to this most recent interest, Knuth and his associates had studied momentum transfer for satellite surfaces (Knuth 1 980, Liu et al. 1 979). An interesting aside is the work of Steinheil et al. (1 977), which demonstrated an impres­sive reduction in tangential accommodation coefficient for specially pre­pared surfaces, with at least one surface being stable to exposure to atmo­spheric-pressure gases. The advent of new surface preparation techniques and designer-type materials, as well as of the new surface measurement techniques such as scanning tunneling microscopy, indicates that the time may be about right for significant advances in the control of surface accommodation coefficients. Uses for such control are not entirely clear, since the drag of practically shaped space vehicles is not very sensitive to the tangential accommodation coefficient. Nevertheless, a very low coefficient could significantly affect the drag of stabilized slender shapes. It is also worth noting that an overall reduction in momentum accom­modation will reduce the density of the cloud of reflected gases that concentrate ahead of the windward surfaces of space vehicles.

The status of intermediate-energy surface interaction has been reviewed by Hurlbut (1 985, 1 989) . Studies of surface-interaction measurements from space flights are reported by Karr et al. (1 986) and Gregory & Peters (1 986) .

Ann

u. R

ev. F

luid

Mec

h. 1

989.

21:3

87-4

22. D

ownl

oade

d fr

om w

ww

.ann

ualr

evie

ws.

org

by W

est V

irgi

nia

Uni

vers

ity o

n 01

/22/

14. F

or p

erso

nal u

se o

nly.

Page 26: Rarefied Gas Dynamics

4 1 2 MUNTZ

The production of medium-energy oxygen atoms in ground facilities has been reported by Cross et al. (1 986) and by Caledonia (1 989). A useful review of low-Earth-orbit atmospheric effects is given by Green et al. ( 1 985).

Non-Navier-Stokes Gas Dynamics

Kogan and his colleagues (cf. Kogan 1 986) have investigated continuum­limit flows that have Burnett terms contributing effects of the same order as the Navier-Stokes terms. When a uniformly heated body is immersed in a gas at a temperature different from the gas temperature, such as when a hot body is immersed in a significantly cooler gas, stress terms appear in the Burnett approximation for so-called slow, nonisothermal flows. Kogan indicates that these terms can lead to surprising results. If a hot sphere is in the Stokes-flow regime, then for large enough heating (say 1 0 times the gas temperature) the drag o n the sphere can become negative. The energy for the work done comes from maintaining the sphere's temperature (Kogan 1 986). Burnett term effects related to the ones just described have been observed experimentally for one specific case in elegant work by Hermans et al. ( 1 977, 1 979). In these experiments, Burnett viscomagnetic heat-flux predictions have been found to agree well with experiments.

Based on the Burnett terms' importance even in low-Mach-number shock waves, as well as the recent successful integration of the Burnett equations at higher Mach numbers and their significant contributions to non-Navier-Stokes continuum flows, it seems that the earlier perception that the Burnett approximation was not very useful may be mistaken. Investigations involving the Burnett equations appear likely to continue.

CONCLUSIONS

Several research subjects within the broad field of rarefied gas dynamics have been discussed here, chosen on the basis of what I consider to be their importance as research directions in the next decade. In this effort I have undoubtedly slighted some, but I hope only a limited number, of my professional colleagues; I trust they will forgive the sins of my omissions. Briefly, I expect the most fundamental issues in rarefied gas dynamics during the next decade to be the following:

I . Assessing the accuracy of computational rarefied-gas-dynamics tech­niques so that they can be used as surrogates for experiments, including investigations of the level of knowledge of the microkinetics necessary for appropriate descriptions of gas flows with active internal degrees of freedom and chemistry.

Ann

u. R

ev. F

luid

Mec

h. 1

989.

21:3

87-4

22. D

ownl

oade

d fr

om w

ww

.ann

ualr

evie

ws.

org

by W

est V

irgi

nia

Uni

vers

ity o

n 01

/22/

14. F

or p

erso

nal u

se o

nly.

Page 27: Rarefied Gas Dynamics

RAREFIED GAS DYNAMICS 413

2. Development of hybrid flow-field computational techniques that match continuum (Navier-Stokes and Burnett computations) with particle computations (DSMC, cellular automata) and that can apply to highly nonequilibrium, radiating, chemically active flows.

3. Validation of computational rarefied gas dynamics by means of experi­mental studies of Knudsen layers in simple flows with a detailed fol­lowing of internal energy-state populations, and gas-mixture flows.

4. A continuing interest in the theoretical analysis of rarefied-gas-dynamics problems, which is essential for understanding and developing com­putational methods.

5. Investigation of designer surfaces for achieving stable, low momentum and thermal accommodation coefficients.

6. Studies of the interactions at 5-20 eV between atmospheric species and both surfaces and gases.

7. Evaporation and condensation experiments and theory relating to events close to surfaces.

8 . Studies relating to the formation, properties, and use of clusters. 9. The analysis of evaporating and condensing particle and aerosol clouds

in both Earth and space environments.

ACKNOWLEDGMENTS

I thank Gail Dwinell and Carolyn Gautier for much assistance with the preparation of this article. In part, the time spent on this review was supported by a Hypersonic Training and Research grant from NASA, AFOSR, and ONR (NASA NAGW 1 06 1 ). Finally, my greatest obligation is to my colleagues from around the world, with whom I have had many exciting discussions mixed with an eclectic assortment of beverages over the past 25 (maybe a few more) years.

Literature Cited I

Allegre, J., Raffin, M., Gottesdiener, L. 1986. Slip effects on supersonic fiowfields around NACA 001 2 airfoils. In RGD 15, pp. 548-57

Alsmeyer, R. 1976. Density profiles in argon and nitrogen shock waves measured by the absorption of an electron beam. J. Fluid Afech. 74: 497-5 1 3

Anders, K., Frohn, A. 1 984. Experimental investigation of droplet evaporation in a

I Complete publication information on the Proceedings of the International Symposia on Rarefied Gas Dynamics (RGD) can be found in Table I .

wide Knudsen number range. I n RGD 14, pp. 975-82

Anderson, J., Fenn, J. B. 1 965. Velocity dis­tributions in molecular beams from nozzle sources. Phys. Fluids 8: 780-87

Anderson, J. B., Foch, J. D., Shaw, M. J., Stern, R. C., Wu, B. J. 1986. Statistical theory of electronic energy relaxation. In RGD 15, pp. 4 1 3-21

Ashkenas, R., Sherman, F. S. 1965. Free jet testing in low density wind tunnels. In RGD 4, 2: 84-105

Banks, P. M., Williamson, P. R., Raitt, W. J. 1 983 . Space shuttle glow observations. Geophys. Res. Lett. 10: 1 18--2 1

Becker, M., Robben, F. , Cattolica, R. 1 974.

Ann

u. R

ev. F

luid

Mec

h. 1

989.

21:3

87-4

22. D

ownl

oade

d fr

om w

ww

.ann

ualr

evie

ws.

org

by W

est V

irgi

nia

Uni

vers

ity o

n 01

/22/

14. F

or p

erso

nal u

se o

nly.

Page 28: Rarefied Gas Dynamics

414 MUNTZ Velocity distribution function near lead­ing edge of a fiat plate. AIAA J. 1 2: 1 247-53

Berner, F., Camac, M. 1 96 1 . Air scooping vehicle. Planet. Space Sci. 4: 1 59-83

Bhatnagar, P. L., Gross, E. P., Krook, M . 1 954. A model for collision processes in gases. Phys. Rev. 94: 5 1 1-24

Bird, G. A. 1963 . Approach to translational equilibrium in a rigid sphere gas. Pilys. Fluids 6: 1 5 1 8- 1 9

Bird, G. A. 1 965. Shock wave structure i n a rigid sphere gas. In RGD 4 , pp. 2 1 6-2 1

Bird, G . A . 1 976. See Table 2 Bird, G. A. 1 978. See Table 2 Bird, G. A. 1979. Simulation of multi-dimen­

sional and chemically reacting flows. In RGD 11, pp. 365-88

Bird, G. A. 1985. Low density aerothermo­dynamics. AlA A Pap. No. 85-0994. 9 pp.

Bird, G. A. 1 986. Direct simulation of typical AOTV entry flows. AIAA Pap. No. 86-1310. 9 pp.

Bird, G. A. 1 987. Nonequilibrium radiation during re-entry at 10 km/s. AIAA Pap. No. 87-1543. 10 pp.

Bird, G. A. 1989. The perception of numeri­cal methods in rarefied gas dynamics. In RGD 16

Bochkaryov, A. A., Rebrov, A. K., Chek­marev, S. P., Sharafutdinov, R. G. 1970. Experimental investigation of spherical and cylindrical rarefied gas expansion with a stationary shock wave. In RGD 7 (Book of abstracts), pp. 401-6

Borgnakke, c., Larsen, P. S. 1 975 . Statistical collision model for Monte Carlo simu­lations of polyatomic gas mixtures. J. Comput. Phys. 1 8 : 405-20

Bossel, U. 197 1 . Nozzle beam properties in presence of skimmer interference. In RG D 7, pp. 663-75

Broadwell, J. I 964a. Shock structure in a simple discrete gas. Phys. Fluids 7(8): 1 243-47

Broadwell, J. 1 964b. Study of rarefied shear flow by the discrete velocity method. J. Fluid Mech. 19: 401-14

Brusdeylins, R., Doak, R. B. , Toennies, J. P. 1 98 1 . Observation of selective desorption of one-photon inelastically scattered He atoms from an LiF crystal surface. J. Chem. Phys. 75: 1 784 92

Biitefisch, K. A., Vennemann, D. 1974. The electron-beam technique in hypersonic rarefied gas dynamics. Prog. Aerosp. Sei. 15: 217-53

Caledonia, G. 1989. Laboratory simulations of energetic atom interactions occurring in low Earth orbit. In RGD 16

Camparguc, R., Lcbehot, A. 1974. High intensity molecular beams with extremely

narrow energy spreads in the 0-37 eV range. In RGD 9, pp. C. 1 1 -1-12

Carnpargue, R., Lebehot, A., Lernonnier, J . c., Marette, D. 1 98 1 . Measured very nar­row velocity distribution for heated Xe and argon seeded nozzle-type molecular beams of He and H2 skimmed from free jet zones of silence, Xe energies up to 30 eV. In RGD 12, pp. 823-43

Cattolica, R., Robben, F., Talbot, L., Willis, D. R. 1 974. Translational non-equilib­rium in free jet expansions. Phys. Fluids 17 : 1793-1807

Cercignani, C. 1 975. See Table 2 Ccrcignani, C. 1 978. Knudsen layer: theory

and experiments. In Recent Developments in Theoretical and Experimental Fluid Mechanics, ed. U. Miller, R. G. Roesner, B. Schmids. Berlin: Springer-Verlag

Cercignani, C. 1979. Transition regime and Boltzmann equation. In RGD 11, pp. 1 4 1 -64

Cercignani, C. 1 988. The Boltzmann Equa­tion and Its Applications. Berlin: Springer­

Verlag Chapman, D. R., Fiscko, K. A., Lumpkin,

F. E. 1987. Fundamental problem in com­puting radiatillg flow fields with thick shock waves. Presented at SPIE Symp. Innova­tive Sci. and Techno!., Los Angeles. 7 pp.

Cheng, H. K. 1988. Problem of in viscid-v is­co us interaction in hypersonic flows. In Progress in Astronautics alld Aeronautics, ed. M . M . Hafez. New York: AIAA. In press

Cheng, H. K., Wong, E. 1988. Fluid dy­namics modelling and numerical simu­lation of low-density hypersonic flows. A IA A Pap. No. 88-2731. 15 pp.

Crawford, D. P., Vogenitz, F. W. 1 974. Monte Carlo calculations of the shock layer structure on adiabatic cylinders in rarefied supersonic flow. In RGD 9, pp. B.24- 1-9

Cross, J. N., Kremers, D. A., Spangler, L. H., Hofibauer, M. A., Archuleta, F. A. 1 986. CO2 laser sustained CW discharge atomic beam source. In RGD 15, pp. 657-66

Crusciel, G. T., Pool, L. A. 1986. Details of the Knudsen layer for a cold wall sphere. In RGD 15, pp. 226-35

Crusciel, G. T., Pool, L. A. 1 987. Monte Carlo flow field sensitivity study. AIAA Pap. No. 87-1544. 10 pp.

Davis, J., Harvey, J. K. 1 979. Experimental studies using a hollow cylinder to evaluate the collision kernels for Monte Carlo cal­culations. In RGD 11, pp. 277-86

d'Humieres, D., Lallernand, P. 1 986. Lattice gas models for 3D hydrodynamics. Euro­phys. Lett. 2(4): 291-97

Edwards, R. H., Cheng, H. K. 1 966 . Steady

Ann

u. R

ev. F

luid

Mec

h. 1

989.

21:3

87-4

22. D

ownl

oade

d fr

om w

ww

.ann

ualr

evie

ws.

org

by W

est V

irgi

nia

Uni

vers

ity o

n 01

/22/

14. F

or p

erso

nal u

se o

nly.

Page 29: Rarefied Gas Dynamics

expansion of a gas into a vacuum. AlA A J. 4: 558-60

Edwards, R. H. , Collins, R. L. 1969. Evap­oration from a spherical source into a vacuum. In RGD 6, pp. 1489-96

Enkenhus, K. R., Harris, E. L., Patterson, G. N. 1960. Pressure probes in free molecule flow. UTIA Rep . No. 62, Inst. Aerophys., Univ. Toronto, Can. 16 pp.

Erwin, D., Pham Van Diep, G. 1989. Com­parison of shock wavc velocity dis­tribution functions with predictions for low, medium and high Mach numbers. In RGD 16

Faubel, M. 1989. Free molecular evapora­tion studies on very thin jets of volatile liquids. In RGD 16

Faubel, M., Kohl, K. H . , Toennies, 1. P. 1 980. Molecular beam time of flight measurements of resolved rotational tran­sitions He + N" CO and CH4 collisions. J. Chem. Phys. 73: 2506-7

Fenn, 1. 1985. Where are we going with molecular beams? In RGD 13, pp. 761 76

Fennell, 1. F., Koons, H. C, Leung, M. S. , Mizera, P. F. 1983. A review of SCATHA satellite results: charging and discharging. Proc. ESLAB Symp. Spacecr. Plasma Interactions and Their Influence on Field and Part. Meas., 1 7th. ESA SP-198, pp. 3 -11

Fiscko, K. A., Chapman, D. R. 1988. Com­parison of shock structure solution using independent continuum and kinetic theory approaches. Presented at SPIE Symp. In­novative Sci. and TechnoL, Los Angeles. 10 pp.

French, 1. B. 1966. Molecular beams for rarefied gasdynamic research. A GARDo­graph No. 112. 85 pp.

Frisch, U . , Hasslacher, 8., Pomeau, Y. 1986. Lattice gas automata for the Navier­Stokes equations. Phys. Rev. Lett. 56: 1505-8

Fuchs, H., Legge, H . 1979. Flow of a water jet into vacuum. Acta Astronaut. 6: 12 13-26

Garen, W., Synofzik, R., Wortberg, G., Frohn, A. 1 977. Experimental inves­tigation of the structure of weak shock waves in noble gases. In RGD 10, pp. 5 1 9-28

Garrett, H. B. 1980. Spacecraft charging: a review. In Space Systems and Their Inter­action with the Space Environment. Pro­gress in Astronautics and Aeronautics, ed. H. B. Garrett, C P. Pike, 7 1 : 1 67-226. New York: AIAA

Gottesdiener, L 1984. Investigation of the Knudsen layer in the trailing edge region of a flat plate. In RGD 14, pp. 249-56

Grad, H. 1 958. See Table 2 Green, 8. D., Caledonia, G. E., Wilkersen,

RAREFIED GAS DYNAMICS 4 1 5

T. D . 1985. The shuttle environment gases, particulates, and glow. J. Spacecr. Rockets 22(5): 500-11

Gregory, 1. C., Peters, P. N. 1986. A measurement of the angular distribution of 5 eV atomic oxygen scattered off a solid surface in Earth's orbit. In RGD 15, pp. 644-56

Gross, E. P. 1960. Recent investigations of the Boltzmann equation. In RGD 1, pp. 139-50

Hagena, O. 1 984. Cluster formation in free jets: results of theory and experiment. In RGD 14, pp. 721-32

Hamel, B. B., Wachman, M. 1965. A discrete ordinate technique for the linearized Boltzmann equation with application to Couette flow. In RGD 4, pp. 370-93

Hamel, 8. B., Willis, D. R. 1966. Kinetic theory of source flow expansion with application to the free jet. Phys. Fluids 9: 829-41

Hasslacher, B. 1987. Discrete fluids. Los A/amos Sci. Spec. Iss. 15: 175-217

Hatakeyama, M. 1 984. Kinetic approach to supersonic condensation processes and bifurcation phenomena. In RGD 14, pp. 909-1 6

Hatakeyama, M . , Oguchi, H. 1979. Kinetic approaches to non-linear condensation of flowing vapor. In RGD 11, pp. 1293-1304

Hermans, L 1. F., Eggermont, G. E. 1., Her­mans, P. W., Beenakker, 1. 1. M. 1 977. Visco-magnetic heat flux: an experimental study in Burnett regime. in RGD 1U, pp. 695-702

Hermans, L 1. F., Eggermont, G. E. 1., Knaap, H . F. P. , Beenakker, J. 1. M . 1979. The use of a magnetic field in an experi­mental verification of transport theory for rarefied gases. In RGD 1 1 , pp. 799-806

Hermina, W. L. 1 987. Monte Carlo simula­tion of rarefied flow along a flat plate. AIAA Pap. No. 87-1547. 7 pp.

Hickman, R. S. 1 967. An experimental study of hypersonic rarefied flow over a 10° cone. In RGD 5, pp. 1067-86

Holtz, T., Muntz, E. P. 1984. Molecular ve­locity distribution functions in an argon normal shock wave at Mach number 7. Phys. Fluids 26(9): 2425-36

Holtz, T., Muntz, E. P., Yen, S.-M. 1971. Comparison of measured and predicted velocity distribution function in a shock wave. Phys. Fluids 14(3): 545-48

Hurlbut, F. C 1 985. Gas surface scattering models for satellite applications. AIAA Pap. No. 85-0997. 9 pp.

Hurlbut, F. C 1 987. Sensitivity of hyper­sonic flow over a flat plate to wall/gas interaction models using DSMC. AIAA Pap. No. 8 7-1545. 7 pp.

Hurlbut, F. C 1 989. Particle surface inter-

Ann

u. R

ev. F

luid

Mec

h. 1

989.

21:3

87-4

22. D

ownl

oade

d fr

om w

ww

.ann

ualr

evie

ws.

org

by W

est V

irgi

nia

Uni

vers

ity o

n 01

/22/

14. F

or p

erso

nal u

se o

nly.

Page 30: Rarefied Gas Dynamics

416 MUNTZ

actions in the orbital context: a survey. In RGD 16

Kantrowitz, A., Grey, J. 1 95 1 . High intensity source for the molecular beam. Part 1 , theoretical. Rev. Sci. Instrum. 22: 328-32

Karr, G. R., Gregory, 1. c., Peters, P. N. 1 986. Free molecule drag and lift deduced from shuttle flight experiments. In RGD 15, pp. 609- 1 7

Knight, C. 1. 1976. Evaporation from a cylindrical surface into a vacuum. J. Fluid Mech. 75(3): 469-86

Knuth, E. L. 1 980. Free-molecule normal­momentum transfer at satellite surfaces. AIAA J. 1 8 (5): 602-5

Kogan, M . N. 1 967. See Table 2 Kogan, M. N. 1 973. See Table 2 Kogan, M . N. 1 985. Transport processes in

Knudsen layers. In RGD 13, pp. 545-60 Kogan, M. N. 1986. Non-Navier-Stokesgas­

dynamics and thermal-stress phenomena. In RGD 15, pp. 1 5-24

Koppenwallner, G. 1 986. The free molecular orifice pressure probe for mass-spectro­scopic studies. In RGD 15, pp. 24--33

Koura, K. 1 973. Nonequilibrium velocity distributions and reaction rates in fast highly exothermic reactions. J. Chern. Phys. 59: 691-97

Krook, M. 1 955. On the solution of equa­tions of transfer I. Astrophys. I 1 22: 488-97

Krylov, A. A., Panov, B. F., Starikov, B. B. 1985. Experimental study of normal and tangential stresses on the surface of simple shape bodies. In RGD 13, pp. 495-502

Kunc, J. 1 989 . Review of efficient semi­classical methods for determination of atomic and molecular interactions. In RGD 16

Lewis, J. 1 989. Laser diagnostics in rarefied flows. In RGD 16

Liepmann, H. W., Narasimha, R., Chahine, M . T. 1 962. Structure of a plane shock layer. Phys. Fluids 5(1 1 ) : 1 3 1 3-24

Liu, S.-M., Sharma, P. K . , Knuth, E. L. 1 979. Satellite drag coefficients calculated from measured distribution of reflected helium atoms. A 1AA J. 17: 1 3 14-19

Marrone, P. V. 1 967. Temperature and den­sity measurements in free jets and shock waves. Phys. Fluids 1 0: 521-38

Mende, S. R, Garriott, O. K., Banks, P. M . 1 983. Observations of optical emissions on STS-4. Geophys. Res. Lett. 10: 1 22-25

Miller, D . R., Andres, R. P. 1 967. Rotational relaxation of molecular nitrogen. J. Chem. Phys. 46: 341 8-23

Miller, D. R., Toennies, J. P., Winkelmann, K. 1974. Quantum effects in highly ex­panded helium nozzle beams. In RGD 9, pp. C. 9. 1-9

Molodtsov, V. K., Ryabov, V. V. 1 985.

Investigation of the structural features of rarefied gas flows about a sphere using Navier-Stokes equations. In RGD 13, pp. 535-44

Moss, J. N. 1 986. Direct simulation of hyper­sonic transitional flow. In RGD 15, pp. 384--99

Moss, J. N. 1987. Non-equilibrium effects for hypersonic transitional flows. A1AA Pap. No. 87-0404. 17 pp.

Moss, J. N., Bird, G., Dogra, V. K. 1988. Non-equilibrium thermal radiation for an aeroassist flight experiment. A1AA Pap. No. 88-0081. 18 pp.

Mott-Smith, H. M . 1 95 1 . The solution of the Boltzmann equation for a shock wave. Phys. Rev. 82: 855-92

Muckenfuss, C. 1 962. Some aspects of shock structure according to the bimodal model. Phys. Fluids 5(1 1) : 1 325-36

Muntz, E. P. 1 967. Measurements of aniso­tropic molecular velocity distribution functions in rapid radial expansions. In RGD 5, pp. 1 257-86

Muntz, E. P. 1 969. The electron beam fluo­rescence technique. AGA RDograph No. 132. I I I pp.

Muntz, E. P. 1981a . Measurement of den­sity by analysis of electron beam excited radiation. In Methods of Experimental Physics, ed. R. 1. Emrich, pp. 434-55. New York: Academic

Muntz, E. P. 1 98 1 b. Measurement of tem­perature by analysis of electron beam excited emissions. In Methods of Experi­mental Physics, ed. R. J. Emrich, pp. 489-97. New York: Academic

Muntz, E. P., Harnett, L. N. 1 969. Molec­ular velocity distribution function measure­ments in a normal shock wave. Phys. Fluids 1 2: 2027-35

Muntz, E. P., Orme, M. 1 987. Charac­teristics, control, and uses of liquid streams in space. A1AA J. 25(5): 746-56

Muntz, E. P., Qian, S.-S., Dixon, M. 1 984. The injection of liquid into high vacuum for RGD and space applications. In RGD 14, pp. 91 9-35

Muntz, E. P., Kunc, 1., Erwin, D. A. 1987. A pulsed electron-photon fluorescence technique for temperature and specic con­centration measurements at points in rela­tively dense, unseeded air flows. AIAA Pap. No. 87-1526. 23 pp.

Nadiga, B. T., Broadwell, J. E., Sturtevant, B. 1 989. Study of a multi-speed cellular automaton. In RGD 16

Nanbu, K. 1 986. Theoretical basis of the direct simulation Monte Carlo method. In RGD 15, pp. 369-83

Nordsieck, A., Hicks, B. L. 1 967. Monte Carlo evaluation of the Boltzmann col­lision integral. In RGD 5, pp. 695-710

Ann

u. R

ev. F

luid

Mec

h. 1

989.

21:3

87-4

22. D

ownl

oade

d fr

om w

ww

.ann

ualr

evie

ws.

org

by W

est V

irgi

nia

Uni

vers

ity o

n 01

/22/

14. F

or p

erso

nal u

se o

nly.

Page 31: Rarefied Gas Dynamics

Oguchi, H. 1 98 1 . One-dimensional, steady supersonic condensation. In RGD 12, pp. 321-29

Patterson, G. N. 1956. Theory of free­molecule, orifice-type pressure probes in isentropic and non-isentropic flows. UTIA Rep. No. 41, Inst. Aerophys., Univ. Toronto, Can. 1 4 pp.

Peters, P. N., Linton, R. c., Miller, E. R. 1983. Results of apparent atomic oxygen reaction on Ag, C, and as exposed during the shuttle STS-4 orbits. Geophys. Res. Lett. 1 0: 569-7 1

Ramohalli, K., Dowler, W., French, J., Ash, R. 1 987. Some aspects of space propulsion with extraterrestial resources. J. Spacecr. Rockets 24(3): 236-44

Rebrov, A. K. 1977. Experimental study of relaxing low-density flows. In RGD 10, pp. 8 1 1-48

Rebrov, A. K. 1 985. Free jet as an object of non-equilibrium processes investigation. In RGD 13, pp. 805-78

Reynolds, M. A., Smolderen, J. J . , Wendt, J. F. 1 974. Velocity profile measurements in the Knudsen layer for the Kramers problem. In RGD 9, pp. A.2 1 . l -14

Robben, F., Talbot, L. 1966. Measurements of rotational temperatures in a low-den­sity wind tunnel. Phys. Fluids 9(4): 644-62

Schaaf, S. 1958. See Table 2 Schilder, R., Adomeit, G., Wortberg, G.

1985. Theoretical and experimental inves­tigations of the Knudsen layer produced by intensive evaporation of iodine. In RGD 13, pp. 577-84

Schmidt, B. 1969. Electron beam density measurements in shock waves in argon. J. Fluid Mech. 39: 361-73

Schumacher, B. W., Gadamer, E . O. 1 958. Electron beam fluorescence probe for measuring the local gas density in a wide field of observation. Can. 1. Phys. 36: 659-7 1

Sherman, F. S . 1 963. A survey o f experi­mental results and methods for the tran­sition regime of rarefied gas dynamics. In RGD 3, 2: 246-49

Sherman, F. S. 1 969. See Table 2 Sherman, F. S., Talbot, L. 1960. Experiment

versus kinetic theory for rarefied gases. In RGD 1, pp. 1 6 1-91

Shidlovskii, V. P. 1 965. See Tahle 2 Simon, C. E., Foch, J. D. 1977. Numerical

integration of the Burnett equations for

RAREFIED GAS DYNAMICS 4 1 7 shock structure i n a Maxwell gas. I n RGD 10, pp. 493-500

Sone, Y., Aoki, K., Yamashita, 1. 1 986. A study of an unsteady strong condensation on a plane condensed phase with special interest in formation of steady profile. In RGD 15, pp. 323-33

Steinheil, E., Sherber, W., Seidl, M., Rieger, H. 1977. Investigation on the interaction of gases and well-defined solid surfaces with respect to possibilities for reduction of aerodynamic friction and aerothermal healing. In RGD 10, pp. 589-602

Struminskii, V. V. 1 980. Towards the theory of systems of identical particles. Dokl. Akad. Nauk SSSR 252: 1 345-55

Sturtevant, 8., Steinhilper, E. A. 1974. Inter­molecular potentials from shock structure experiments. In RGD 8, pp. 1 59-66

Takens, W. 8., Mischke, W., Korving, J., Beenakker, J. J. M . 1 984. A spectroscopic study of free evaporation of sodium. In RGD 14, pp. 967-74

Thomas, L. 8. 1 985. Accommodations of gases on solid surfaces-description of experimental methods and of some recently completed investigations. In RGD 13, pp. 621-36

Tsien, H. S. 1 946. Superaerodynamics, mechanics of rarefied gases. J. Aeronaut. Sci. 1 3 : 653-64

Velikodnyi, V. Yu., Korolokov, G. A., Orlov, A. V. 1 985. Structure of a shock wave in a rarefied gas. In RGD 13, pp. 253-58

Vincenti, W. G., Kruger, C. H. 1 965. See Table 2

Whitaker, A. 1 983. LEO elTeds on space­craft materials. AIAA Pap. No. 83-2632. 1 5 pp.

Yen, S. M. 1 97 1 . Monte Carlo solutions of nonlinear Boltzmann equation for prob­lems of heat transfer in rarefied gases. Tnt. J. Heat Mass Transfer 1 4: 1 865-69

Yen, S. M. 1984. Numerical solution of the nonlinear Boltzmann equation for non­equilibrium gas flow problems. Ann. Rev. Fluid Mech. 1 6: 67-97

Ytrehus, T. 1 977. Theory and experiments on ga� kinetics in evaporation. In RGD 10, pp. 1 197-1 2 1 2

Ytrehus, T . , Aukrust, T. 1984. Flow break­down in condensation Knudsen layers. In RGD 14, pp. 901-8

Zahm, A. F. 1934. Superaerodynamics. 1. Franklin Inst. 2 1 7: 1 53-66

Ann

u. R

ev. F

luid

Mec

h. 1

989.

21:3

87-4

22. D

ownl

oade

d fr

om w

ww

.ann

ualr

evie

ws.

org

by W

est V

irgi

nia

Uni

vers

ity o

n 01

/22/

14. F

or p

erso

nal u

se o

nly.