Вы находитесь на странице: 1из 19

AIAA JOURNAL

Vol. 51, No. 12, December 2013

Thruster Plume Surface Interactions: Applications


for Spacecraft Landings on Planetary Bodies

Manish Mehta∗
NASA Marshall Space Flight Center, Huntsville, Alabama 35812
Anita Sengupta†
NASA Jet Propulsion Laboratory, Pasadena, California 91109
Nilton O. Renno‡
University of Michigan, Ann Arbor, Michigan 48109
John W. Van Norman§
Analytical Mechanics Associates, Inc., Hampton, Virginia 23666
Peter G. Huseman¶ and Douglas S. Gulick**
Lockheed Martin Space Systems, Denver, Colorado 80125
Downloaded by UNIVERSITY OF MICHIGAN on May 20, 2014 | http://arc.aiaa.org | DOI: 10.2514/1.J052408

and
Mark Pokora††
University of Michigan, Ann Arbor, Michigan 48109
DOI: 10.2514/1.J052408
Numerical and experimental investigations of supersonic jet interactions with a flat surface at various atmospheric
pressures are presented in this paper. These studies were done in assessing the landing hazards of both the NASA
Mars Science Laboratory and the Phoenix Mars spacecraft. Temporal and spatial ground pressure measurements in
conjunction with numerical solutions at altitudes of ∼35 nozzle exit diameters and jet expansion ratios e between 0.02
and 100 are used. This study shows that, for typical landing spacecraft engine parameters, thruster plumes exhausting
into Martian environments create the largest surface pressure loads and can occur at high spacecraft altitudes in
contrast to the jet interactions, which occur in terrestrial and lunar atmospheres. These differences are dependent on
the stability and dynamics of the plate shock, the length of the supersonic core, and plume decay due to shear layer
instability, all of which are functions of the jet expansion ratio. Theoretical, experimental, and analytical results show
that subscale supersonic cold gas jets adequately simulate the flowfield and loads due to rocket plume impingement,
provided important scaling parameters are in agreement. These studies indicate the critical importance of testing and
modeling plume–surface interactions for descent and ascent of spacecraft and launch vehicles.

Nomenclature q = dynamic pressure, Pa


A = nozzle cross-sectional area, m2 R = gas constant, J∕K · mol
cp = plume specific heat capacity at constant pressure, T = temperature, K
J∕kg · K t = time, s
cv = specific heat capacity at constant volume, J∕kg · K u, U = plume velocity, m∕s
d, D = diameter, m x = axial length along the jet/plume axis, m
e = jet expansion ratio β = plume coefficient of thermal expansion, 1∕K
F = force, N γ = cp ∕cv , specific heat ratio of exhaust plume
f = body force, N ε = nozzle expansion ratio
g = gravitational acceleration, m∕s2 λ = plume thermal conductivity, W∕m · K
h = engine pulse frequency, Hz μ = plume viscosity, Pa · s
m = plume mass flow rate, kg∕s ρ = jet or rocket plume density, kg∕m3
P = pressure, Pa τ = shear stress, Pa
φ = rate of viscous dissipation, m∕s3

Received 24 October 2012; revision received 15 March 2013; accepted for Subscripts
publication 22 March 2013; published online 16 August 2013. This material
is declared a work of the U.S. Government and is not subject to copyright c = instantaneous or nominal plenum pressure, Pa
protection in the United States. Copies of this paper may be made for personal e = nozzle exit conditions
or internal use, on condition that the copier pay the $10.00 per-copy fee to fb = free boundary conditions
the Copyright Clearance Center, Inc., 222 Rosewood Drive, Danvers, MA t = total
01923; include the code 1533-385X/13 and $10.00 in correspondence with ∞, b = ambient conditions
the CCC.
, s = solid surface boundary conditions
*Aerospace Engineer, Aerosciences Branch, EV33; manish.mehta@nasa
.gov. Member AIAA.

Project Manager, California Institute of Technology, Astronomy Physics
Directorate. Member AIAA. I. Introduction

Professor, Department of Atmospheric and Space Sciences. Member
AIAA.
§
Aerospace Engineer, Atmospheric Flight and Entry Systems Branch. S UPERSONIC jet interactions with the ground or a flat surface are
a complex fluid dynamics problem with many nonlinearities.
These nonlinearities arise from shock-wave surface interactions,
Member AIAA.

Senior Manager, Aerosciences Division. Associate Fellow AIAA. stagnation bubble formation, and the propagation of wall jets along
**Aeronautical Engineer, Aerosciences Division. Member AIAA. the surface. Lamont and Hunt [1] studied the flowfield and surface
††
M.S. Student, Department of Mechanical Engineering. interactions due to axisymmetric underexpanded supersonic nitrogen
2800
MEHTA ET AL. 2801

jets at distances between one and three nozzle exit diameters. Steady- II. Rocket Plume Structure in Various Atmospheric
state numerical simulations, conducted by Fujii et al. [2] at these dis- Environments
tances, show good agreement with Lamont and Hunt’s experimental
Plume–surface flow physics in tenuous to vacuum atmospheres
results [1]. Most of the literature concentrates on studies where the
due to the fact that steady supersonic jets have been studied by
nozzle pressure ratios (NPRs) are below 10 [1,3,4]. NPR is the ratio of
researchers in preparation for the Apollo missions [11]. The main
nozzle total chamber pressure to ambient pressure.
methods of characterizing these flows are to spatially and temporally
By applying schlieren imaging and ground pressure sensors along
determine the pressure fields and density gradients. There are three
with numerical simulations, important flow structures such as the
types of supersonic free-jets that exhaust from nozzles. The jet or
plate shock and stagnation bubble were identified. A plate shock is a
rocket plume can either be underexpanded, overexpanded, or
reflected and detached shock wave from a surface due to the
perfectly expanded [12]. The jet pressure tries to match the ambient
impingement of a supersonic jet [1,5]. A stagnation bubble
pressure, which leads to large differences in their shock structure.
(recirculation zones) can form below the plate shock and has only
For an underexpanded jet, the exit jet pressure Pe is larger than
recently received more attention due to its effects on acoustic noise
the ambient pressure Pb , and this leads to Prandtl–Meyer (PM)
production [6,7]. The third structure of importance is the propagating
expansion waves, initiated at the lip of the nozzle (Fig. 1a), which
wall jet, which can reach supersonic speeds, demonstrate compres-
reduce the jet pressure to match the ambient. Because of reflection of
sion and expansion regimes, and decay as a function of axial distance
the PM expansion waves in region 2, this leads to a jet pressure
from the impingement point [7]. These flow features are within the
smaller than the ambient in region 3. Hence, this results in the
far-field or shock-wave interaction regime and are considerably
reflection of these expansion waves from the ambient boundary,
different from the flow structures observed in the near-field regime.
Downloaded by UNIVERSITY OF MICHIGAN on May 20, 2014 | http://arc.aiaa.org | DOI: 10.2514/1.J052408

The near-field regime is within the first nozzle exit diameter from resulting in a series of compression waves. The coalescing of the
the nozzle exit plane. Near-field supersonic jet characteristics are compression waves leads to oblique shocks. This results in the
dependent on the nozzle chamber stagnation pressure, exhaust matching of the two competing pressures in region 4. When oblique
product specific heat ratio, nozzle area ratio, and atmospheric pres- shock reflection occurs, the jet pressure is larger than ambient in
sure. There are three types of jet flow characteristics that may be region 5, and the process starts again. These are known as “shock
observed in this regime: overexpansion, underexpansion, and perfect cells”, which form as repeatable trainlike structures, as shown in
expansion [8]. All three flow characteristics are observed during the Fig. 1a. They dissipate further downstream due to viscous losses and
ascent of a rocket into space and, as discussed in this paper, lead to turbulent mixing with the ambient atmosphere. The planar laser-
different far-field surface interactions. induced fluorescence (PLIF) image in Fig. 1b shows a contour of the
The main focus of this study is to investigate the flow physics of standard deviation of the mean velocity (standard-deviation image) in
plume–ground interactions from exhaust plumes of rocket engines Fig. 1c of an underexpanded jet [13]. The standard-deviation image
during planetary landings, specifically for environments of Mars. For correlates to turbulence intensity.
appropriate simulations, this requires different flow requirements For an overexpanded jet, the exit pressure is smaller than the
than those observed for past studies conducted by Lamont and Hunt ambient, and this leads to the formation of oblique shock waves at the
[1], Krothapalli et al. [9], Henderson et al. [6], and other researchers
focused on the acoustic nature of impinging jets. Rocket plumes
exhausting into near-vacuum planetary environments demonstrate
higher exit Mach numbers on the order Mach ∼5 with nozzle pressure
ratios greater than 1000, three orders of magnitude greater than those
observed in acoustic studies [9]. Rocket plumes can interact with the
surface at altitudes between ∼100d and ∼5d, where d is the nozzle
exit diameter [10], to decelerate the spacecraft and to ensure a
successful soft landing. In contrast to previous acoustic studies, all of
our tests were conducted at reduced atmospheric pressures, which
spanned from the Martian to terrestrial environments. The largest
difference between previous jet impingement studies is that the
engine mode can be either pulsed or steady during landings and
attitude corrections. Comparative studies between these two modes
are limited.
This paper will discuss numerical and experimental ground
interaction data for pulsed and steady underexpanded supersonic jets
exhausting from the simulated Phoenix rocket engine module (REM)
and Mars Science Laboratory (MSL) main landing engine (MLE)
nozzles. More importantly, we will focus on interactions associated
with jets exhausting from high altitudes of h > 20d. We will then
compare our experimental subscale temporal and spatial results with
numerical simulations at both subscale and full scale to provide
further insight into the complex flow physics and to ascertain the
reliability of nozzle scaling laws. This study was performed to reduce
mission risk for both of NASA’s recent Mars missions.
This paper is organized as follows. Section II discusses previous
literature on rocket plume structure in various atmospheric
environments. Section III derives scaling laws for rocket plume
flow and jet impingement physics. Section IV describes the landing
sequence of two NASA Mars missions in which the plume-
impingement dynamics are being investigated. Sections V and VI
describe the numerical and experimental methodologies behind the
ground tests conducted in support of the 2007 Phoenix Mars and Fig. 1 Plume shock structure: a) schematic of an underexpanded jet,
2011 MSL missions. The results and discussion of these two complex b) PLIF image, c) PLIF standard-deviation image [13], d) schematic of
test programs are described in Secs. VII and VIII, respectively. The an overexpanded jet, e) PLIF image, and f) PLIF standard-deviation
paper’s main findings are listed within Sec. IX. image [13].
2802 MEHTA ET AL.

nozzle lip, shown in region 1 in Fig. 1d. The same dynamics as for the steady-state conditions. One of our goals in this paper is to investigate
underexpanded jets are observed, but they are out of phase where it is the behavior of these structures in transient conditions.
now first compression and then expansion. The PLIF and standard-
deviation images are shown in Figs. 1e and 1f, respectively [13].
Turbulence and unsteadiness are much more pronounced for over-
III. Scaling Laws for Rocket Plume Flow Physics
expanded jets, which may lead to instability and faster dissipation of
the plume structure. According to Hagerman and Frey [14], the Theoretical scaling laws prove that subscale cold gas jets can
diffusion rate of the entrained flow into the turbulent mixing region simulate the shock structure and ground pressure profiles due to
(the interface region between the jet shock and the ambient atmo- full-scale rocket plume interactions. This is proven through first
sphere) is larger than the underexpanded case and increases axially. principles by normalizing the conservation equations. The governing
Because of overexpansion of the plume, the higher-density free compressible Navier–Stokes equations [Eqs. (1–3)] are presented in
atmospheric boundary pinches the jet inward, leading to an increase symbolic notation next [17]:
in the turbulent mixing region and attenuation of the inviscid core.

Most rocket plumes exhausting from a descent engine on Mars or  −ρ∇ · u (1)
the moon are underexpanded, and this classification will be the prime Dt
focus of our studies. However, the plume structure and their effects on
the surface are considerably different between the two atmospheres. Du ↔
Flow structures due to impinging steady jets were separated into ρ  f − ∇p  ∇ · τ  (2)
three regimes as discussed by Donaldson et al. [15]: 1) free jet, Dt
Downloaded by UNIVERSITY OF MICHIGAN on May 20, 2014 | http://arc.aiaa.org | DOI: 10.2514/1.J052408

2) impingement zone, and 3) wall jet. These structures are mainly


characterized by pressure and density fields at steady state as done by DT Dp
Stitt [16] for preparation of the first landings on the moon. The flow ρcp  ∇ · λ∇T  βT ϕ (3)
Dt Dt
structures within the impingement zone and wall jet will be briefly
discussed due to their importance in ground `erosion [1]. The plate where T, p, ρ, v, λ, β, cp , τ, t, and ϕ are the plume temperature,
shock, tail shock, and stagnation bubble below the plate shock, as pressure, density, velocity, thermal conductivity, coefficient of
shown in Fig. 2a, are important flow structures, which are shown to thermal expansion, isobaric specific heat capacity, viscous stress
directly influence the ground pressure. Tail shocks are oblique shock tensor, time, and rate of viscous dissipation, respectively. The body
waves that are reflected from the plate shock and emanate from the force is denoted by f. Arrows above the parameter denote that these
triple point. The triple point is the region where the incident, plate, terms are vectors. These compressible Navier–Stokes equations are
and tail shocks converge (Fig. 2a). Because the total pressure loss is normalized with the following parameters [17]:
much greater for a plate shock than for oblique tail shocks, there is a
ring of relatively high surface pressure. A portion of the flow below x 0  x∕D; t 0  tUe ∕D; v 0  u∕Ue ;
the plate shock cannot overcome the relatively large pressure gradient
to the outboard of the plate shock, resulting in the recirculation of gas p 0  p − pe ∕ρe U2e ; T 0  T∕T e ; ρ 0  ρ∕ρe ;
(stagnation bubble) [6]. For comparison, the schematic and the cp0  cp ∕cpe ; λ 0  λ∕λe ; β 0  β∕βe ;
schlieren photograph (Fig. 2b) of an underexpanded impinging jet is
shown. Past studies predominantly investigated these structures at ∇ 0  D∇; ϕ 0  ϕD2 ∕μe U2e ; μ 0  μ∕μe (4)

All of the reference parameters are taken at the nozzle exit (denoted as
a subscript e) with a diameter of D. The length scale and plume
viscosity are denoted by D and μ, respectively. Upon substituting the
reference parameters [Eq. (4)] into Eqs. (1–3), the normalized
Navier–Stokes equations become

Dρ 0
 −ρ 0 ∇ 0 · v 0 (5)
Dt

Dv 0 1 1 0 ↔0
ρ0  2 ρ 0g − ∇ 0p 0  ∇ · τ  (6)
Dt 0 Fr Re

DT 0 1 U2 0 0 Dp 0 U2
ρ 0 cp0 0  ∇ 0 · λ 0 ∇ 0 T 0   βT 0  ϕ0
Dt RePr cpe T e Dt cpe T e Re
(7)

U2 c2
 Ma2 e  Ma2 γ − 1 (8)
cpe T e cpe T e

The speed of sound at the nozzle exit is denoted by ce. Substitute


Eq. (8) into Eq. (7) for compressible ideal gas flow, and Eq. (7)
becomes

DT 0 1
ρ 0 cp0  ∇ 0 · λ 0 ∇ 0 T 0 
Dt 0 Re Pr
Dp 0 Ma2 γ − 1 0
 Ma2 γ − 1β 0 T 0  ϕ (9)
Fig. 2 Plume shock structure within the impingement zone of an Dt 0 Re
underexpanded jet: a) schematic of flow structures, and b) schlieren
image distinctly showing a curved plate shock [1]. where the following nondimensional terms are defined as
MEHTA ET AL. 2803

ρe Ue D Ue Ue Pe
Re  ; Fr  p ; Ma  ; e (19)
μe gD ae P∞
μe cpe cpe
Pr  ; γ (10) The normalized boundary conditions at the solid surface are the
λe cve
following:
The variable R is the plume gas constant, which is dependent on
exhaust product molecular weights. Hence, the flow physics of the Ps − Pe
P0  (20)
rocket plume/supersonic jet is a function of the Reynolds number Re, ρe U2e
Froude number Fr, Mach number Ma, Prandtl number Pr, and
specific heat ratio γ. However, simplifications can be made due to the Ts
supersonic flow regime. For example, as Fr → ∞, the first term on T 0  (21)
Te
the right-hand side of Eq. (6) tends to zero. Also, as Re → ∞ and
applied to Eqs. (6) and (7), characteristic of supersonic flows, the
flow can be considered inviscid. This is valid for supersonic jet v0  0 (22)
interactions in all regions except at the viscous boundary layer.
Here, the viscosity of the jet plays a role especially in determining The dimensional term T s is the surface impingement temperature.
the heat transfer and thermal boundary layer. However, the ground According to the governing Navier–Stokes equations and boundary
total pressure has a relatively minor dependence on the jet viscosity conditions specific for supersonic jet interactions, there are five
for large Reynolds number, and this is known as the Barker effect
Downloaded by UNIVERSITY OF MICHIGAN on May 20, 2014 | http://arc.aiaa.org | DOI: 10.2514/1.J052408

nondimensional numbers that need to be satisfied to ensure dynamic


[17,18]. similarity: γ, Ma, Re, e, α. From Eq. (20) and the matching of these
Similar analyses need to be applied to the boundary conditions required nondimensional numbers, the ground pressure is theore-
imposed by these flows. There are two types of surface boundary tically simulated. The ground temperature may be calculated from
conditions for this application: 1) free boundary, and 2) solid this approach, but its large dependence on the viscous boundary layer
boundary. The free surface boundary is defined as the infinitesimal may lead to large errors. Next is a description of each of these
interface between the exhaust plumes and the ambient atmosphere parameters.
P∞ . At this interface region near the nozzle exit, the entrained flow The specific heat ratio for the gas mixture of hydrazine combustion
velocity and temperature are assumed continuous to the plume. The products at the nozzle exit is not analytically straightforward to
solid surface boundary, intuitively, is defined as the solid surface determine. This parameter was numerically calculated throughout
where the jets interact. At the free boundary [17], the nozzle using a two-dimensional method of characteristics (MOC)
code fully coupled with finite-rate kinetics. From these numerical
pfb  P∞  Pamb (11) calculations, the specific heat ratio between the rocket plume and N2
are within 3% of agreement, as tabulated in Table 1, and hence
vfb  Ue (12) nitrogen test gas was used for our experiments. Also, based upon the
MOC result, plume gas implementation within the OVERFLOW
computational-fluid-dynamics code [19] assumed a chemically
T fb  T e (13) frozen, single equivalent gas species of the representative hydrazine
exhaust products, with specific heat ratio of ∼1.4 [20].
Substituting Eqs. (11–13) into the normalized parameters shown in The exit Mach number between the two flows is matched by
Eq. (4), these parameters become simulating the nozzle expansion ratio, nozzle contour profile, and
specific heat ratio. This ensures that the compressibility effects are
0  P∞ − Pe simulated within the plume. This is also confirmed by numerical
pfb (14) simulations. The Mach similarity parameter is a function of the exit
ρe U2e
Mach number and specific heat ratio and is defined as the ratio of the
kinetic energy of the plume to the internal energy of the plume at the
0
vfb  1.0 (15) nozzle exit [21]:

k  γγ − 1M2 (23)


0
T fb  1.0 (16)
There are four flow regimes determined by the Reynolds number in
increasing order: 1) Stokes or creeping (<1), 2) laminar (100 –103 ),
Because the specific heat ratio and the nozzle area expansion ratio are
3) transition (103 –104 ), and 4) turbulent (>104 ) flows. For rocket
similar for full-scale and subscale systems, the dynamic pressure at
plumes and supersonic jets used in the experiments, the flow is fully
the nozzle exit, ρe U2e , will be similar to isentropic relations, provided
the stagnation pressure at the nozzle inlet, PC are matched. The flow turbulent, as shown in Table 2.
approximation is one-dimensional isentropic at the nozzle exit. By The other parameter, as derived previously, used to scale the test is
manipulating Eq. (14), another nondimensional term named the jet the jet expansion ratio of the exhaust plume at the nozzle exit. This
expansion ratio e, developed from the boundary conditions, needs to term has important physical interpretations as well. The pressure
be satisfied: force ratio of the rocket plume at the nozzle exit (Fe ) relative to the
atmosphere (F∞ ) is important in determining both the jet expansion
angle of the plume with respect to the centerline [10] and shock
P∞ − Pe P∞ 1 − P∞e  P∞ 1 − e
P
0 
pfb   (17) structure of the plume. This parameter described the effects of the
ρe U2e ρe U2e ρe U2e ambient atmosphere on the exhaust plume. Its far-field effects are
described in more detail in this paper. This, in turn, is determined by
PC needs to be matched for both full-scale and subscale systems for the ratio of nozzle exit pressure Pe to ambient pressure P∞ [22]:
Eq. (17) to be satisfied, and this is defined as the nozzle pressure ratio
(NPR) α. PC−max is defined as the stagnation chamber pressure when Fe PA P
maximum engine thrust is reached. The NPR and jet expansion ratio e  e e  e (24)
F∞ P∞ Ae P∞
are defined as follows:
The final scaling parameter used is the nozzle pressure ratio
Pc−max
α (18) [Eq. (18)]. NPR is an important parameter in ensuring that the test
P∞ nozzles have ground pressure profiles similar to the full-scale case of
2804 MEHTA ET AL.

Table 1 Dimensional parameters


Parameter Symbol Phoenix REM: 100% throttle MSL MLE: 100% throttle
Half-scale Full-scale Quarter-scale Full-scale
Min area, m2 A 3.83 × 10−5 15.3 × 10−5 6.9 × 10−5 110.3 × 10−5
Exit area, m2 Ae 0.00079 0.00318 0.00193 0.03089
Expansion ratio Ae ∕A 20.7 20.7 28.0 28.0
Mass flow rate, kg∕s m_ 0.11 0.16 0.28 1.52
Chamber temperature, K Tc 300 1114 300.0 1218.2
Chamber pressure, kPa Pc 1240 1240 1765 1765
Exit density, kg∕m3 ρe 0.194 0.026 0.202 0.023
Exit pressure, Pa Pe 3091 3241 2837 2927
Exit temperature, K Te 54 217 47 202
Exit Mach number Me 4.8 4.7 5.1 5.1
Exit velocity, m∕s ve 713 1929 721 2123
Thrust, N F 80 321 206 3313
Time, s T 0.75,1.5,3.0 <2.0 1 ∼6
Pulse frequency, Hz F 10 10 N/A N/A
Pulse width, ms PW ∼65 55 N/A N/A
Altitude h∕d 8.4–25 8.4–80 34.5 34.5–50.0
Slope, deg 0 ∼0 0.0, 22.5 0.0, 22.5
Downloaded by UNIVERSITY OF MICHIGAN on May 20, 2014 | http://arc.aiaa.org | DOI: 10.2514/1.J052408

Specific heat ratio Γ 1.40 1.38 1.40 1.38

Table 2 Scaling parameters


Parameter Symbol MSL MLE Phoenix REM
Quarter-scale Full-scale Half-scale Full-scale
Hypersonic similarity k 14.8 14.0 12.7 11.4
Jet expansion ratio e 2.9–2.1a 6.8–2.2c ∼4.4a 3.8c
Jet expansion ratio e 3.5b 6.8–4.1b 4.5b 4.7b
Reynolds number Re 24.5 − 14.7 × 105 b 8.4 − 5.0 × 105 c 12.7 × 105 3.4 × 105
Reynolds number Re 23.3 × 105 b — — — — — —
Mach number Me 5.14 5.08 4.77 4.67
Strouhal Number St 0 0 4.4 × 10−4 3.3 × 10−4
a
Experiment.
b
Numerical simulation.
c
Spaceflight conditions.

the real size and performance of the rocket motor. Simulation of IV. Entry, Descent, and Landing Sequences
the thruster plume temperature is not critical in understanding the for NASA Mars Spacecraft
force loads on the surface [11]. This is further corroborated by For these studies, the main area of focus during the entry, descent,
normalization of the conservation of energy [Eq. (9)], which is only a and landing (EDL) sequence is the terminal descent phase. During
function of the Mach similarity parameter, Reynolds number, and this phase, the spacecraft is descending at a constant velocity until
Prandtl number for supersonic turbulent flows. touchdown onto the Martian surface. This requires the ignition of
The Strouhal number, which is the ratio of the inertial force the descent rocket engines because the required payload mass is
associated with unsteady flow characteristics to the inertial force due exceeded for the use of landing airbags as the desired landing
to the velocity gradient [Eq. (25)]. This nondimensional number is technique used for the Mars Exploration Rovers.
applied strictly for pulsed jets [23]. The dimensional quantities used
to calculate these scaling parameters are tabulated in Table 1. We will
A. 2011 Mars Science Laboratory
briefly discuss how these various parameters, in particular the jet
expansion ratio and Strouhal number, change the flow physics at the For this study, the main area of focus is the sky crane landing phase,
far-field/interaction regime: as shown in Fig. 3. After successful entry and descent events such as
the survival of the heat shield upon Mars guided entry, deployment of
hD the supersonic parachute, ground radar locking, and backshell
St  (25) separation, the powered descent phase is initiated. This starts at an
U
altitude of 1.3 km, where all eight engines are firing. At 14 m above
By matching these nondimensional numbers and the geometric the surface, the sky crane maneuver is initiated where only four of the
length scaling with respect to the nozzle diameter, the ground canted descent engines are firing to slow the descent stage to a
pressure profiles and the plume and impingement flow structures constant velocity of ∼0.75 m∕s. During the sky crane maneuver, the
produced by the test nozzles theoretically simulate those produced by rover is let down to the Martian surface by umbilical cords. Upon the
the rocket exhaust. This is further confirmed by comparing full-scale rover touchdown, the umbilical cords are cut, and the descent stage
and subscale numerical analyses as described in this paper. Table 2 increases its throttle, increases both vertical and horizontal velocities,
quantitatively compares nondimensional plume parameters between and crashes away from the landing site.
the experimental setup and full scale. Based on the similarity of the
nondimensional parameters shown in Table 2, the subscale tests B. 2007 Phoenix Mars Lander
theoretically simulate the interactions of the Phoenix and MSL rocket For this study, the main area of focus is the constant-velocity
plume with impermeable surfaces. The rocket exhaust parameters terminal descent phase, as shown in Fig. 4. L, E, r, and γ are defined in
may vary slightly depending on the extent of ammonia disassociation this figure as the touchdown time, spacecraft atmosphere entry time,
during the hydrazine decomposition reaction. distance from surface to spacecraft, and flight path angle, respectively.
MEHTA ET AL. 2805
Downloaded by UNIVERSITY OF MICHIGAN on May 20, 2014 | http://arc.aiaa.org | DOI: 10.2514/1.J052408

Fig. 3 NASA MSL EDL sequence and Sky Crane landing technique; courtesy of NASA/Jet Propulsion Laboratory/California Institute of Technology.

Fig. 4 NASA Phoenix Mars Lander EDL sequence. Courtesy of NASA.

This figure is developed from reconstructed data obtained from the ∼1.7 s. Experimental tests and numerical simulations investigate these
200 Hz sampled inertial measurement units onboard the spacecraft. interactions.
Figure 4 shows the time to spacecraft touchdown (L) and passing of the
atmospheric entry point (E), altitude (shown in bold), and descent
velocity. After backshell separation and the lander is momentarily V. Experimental Methodology
falling, the 12 pulsed descent engines are firing to ensure a soft landing. A. Jet–Ground Interactions Due to the Subscale Phoenix
The spacecraft is in constant-velocity terminal descent mode for the Thruster Plumes
last ∼20 s of the landing sequence. This is where the engines are firing A half-scale cold flow (nonheated jet) testbed (CFTB) was
at a 10 Hz frequency with a ∼50% duty cycle. However, from previous developed to study the impingement of supersonic pulsed jets on a
studies [24], plume–surface interaction is only observed for the last flat surface at Mars ambient pressure. The thruster firing frequency,
2806 MEHTA ET AL.

the duration of the pressure pulse, and the chamber pressure PC were MLE hot-fire tests. A schematic of the experimental setup is shown
adjustable. Dry compressed nitrogen gas at room temperature was in Fig. 5.
used to simulate hydrazine decomposition products because it has a A converging–diverging nozzle with an area ratio of 28 was used to
similar specific heat capacity ratio. The importance of this parameter match the nozzle contours on the MLEs of the MSL descent stage
in scaling is discussed in the following section. Fast-response spacecraft [27]. The subscale nozzle was canted to 22.5 deg to
absolute microelectromechanical system (MEMS) pressure sensors accurately simulate the nozzle orientation and configuration on the
were placed radially across the impingement plate at a spacing MSL descent stage (Fig. 5). The experimental nozzle is a quarter-
distance of 27.5 mm between sensors. A gauge pressure transducer scale of the MLE flight system with a nominal exit Mach number of 5.
was placed at the nozzle inlet, which measured the stagnation The varied parameters are the NPR or the jet expansion ratio e and
pressure of the incoming flow. Both transducer and MEMS sensors ground slope. The NPR ranged from 12 to 1200, and the jet expansion
each had a response time of 1 ms. One thermocouple was also placed ratio spanned between 0.02 and 3.00. The MLE simulations are
at the plate’s centerline. One half-scale thruster with a similar nozzle operated at an altitude of 35d with ground slopes of 0 and 22.5 deg. The
contour profile as the Phoenix MR-107 descent engine nozzle was test matrix is included in Table 3. This simulates the lowest altitude
horizontally mounted inside a thermal-vacuum chamber, which was the MSL descent stage approaches the surface before separation of the
set to an ambient pressure of 690 Pa and ambient temperature of MSL rover [27]. Other important nozzle and test specifications, such as
290 K. The ambient pressure within the thermal-vacuum chamber area ratio, exit diameter, etc., are tabulated in Table 1. The exhausting
was generated by a mechanical pump and monitored by a pressure jet impinges onto a 1.2 × 1.8 m aluminum plate (iplate) of which 24
transducer with a response time of 1 ms at a maximum sensitivity of surficial fast responsive piezoresistive pressure sensors are located
1000 Pa. During the constant-velocity descent phase of the Phoenix (Fig. 5). These sensors are then recorded at 1 kHz by a simultaneous-
Downloaded by UNIVERSITY OF MICHIGAN on May 20, 2014 | http://arc.aiaa.org | DOI: 10.2514/1.J052408

spacecraft, the rocket plumes are pulsed at a 10 Hz frequency, with a sampling data-acquisition system with an accuracy of 0.1% of full
55–45 ms pulse width, a maximum chamber pressure PC−max of scale. These pressure sensors are scattered in high concentration near
1.24 MPa, and a chamber pressure PC rate of change during engine the expected impingement point, with lower concentration outward
startup/shutdown cycles of approximately 152 MPa∕s. Our CFTB from this point. The ground pressure contour plots use linear
system generally met all of these requirements as shown in Table 1, interpolation between sensor data points. The stagnation pressure
but it took many design iterations to obtain the required performance. within the inlet nozzle chamber (Pc ) is recorded by a fast responsive
The thruster chamber pressure and the ground impingement pressure transducer at 1 kHz with an accuracy of 0.25% of full scale.
pressures (Ps ) were measured at 1 kHz with an accuracy of 0.1% of Two transparent baffle planes at 90 deg as seen in Fig. 6
(Plexiglass) are used in our experiments to simulate the effects of the
full-scale for the ground pressure sensors and 0.25% for the thrust
outboard thrusters shown in Fig. 6 as done for the Phoenix-based
chamber pressure transducer. The thruster altitude (i.e., the distance
experiments [28,29]. Only one MLE per quadrant is firing during
of the thruster exit plane from the impingement plate) can be adjusted
descent. The jets from symmetrically opposing thrusters stagnate at
from 0.25 m (scaled touchdown altitude) to 1 m above the surface.
the midline between the two engines, and a similar testing
For the full-scale and subscale Phoenix REM nozzles, the jets were
methodology was also used in the subscale NASA Phoenix and
pulsed with similar initial conditions described in Table 1. Although
Viking studies [24,28]. These planes were also used to minimize the
during Phoenix descent, where 12 engines were firing, a single nozzle
complexity of the experiment.
was used to validate the numerical solutions at the University of
Michigan. The plume interactions with the surface at a 0 deg slope for
the Phoenix spacecraft lasted less than 2 s, which decreased in VI. Numerical Methodology
altitude from ∼60d to 8.4d, touchdown altitude [25]. The varied
It should be clarified that the main objective of these studies was
parameters were the jet expansion ratio and the altitude from 8.4d to not for validation but to characterize the thruster-plume–surface
25d. A more detailed experimental setup for the subscale Phoenix interactions for spacecraft landings. Mainly qualitative flow features,
cases is presented in Plemmons et al. [26]. general trends, and its effects on the surface were addressed in this
paper. Hence, significant details in grid and numerical convergence
B. Jet–Ground Interactions Due to the Subscale Mars Science have not been described in this paper. All numerical simulations are
Laboratory Thruster Plumes limited to continuum, frozen, perfect gas, and Reynolds-averaged
These experiments were conducted at NASA Ames Research Navier–Stokes (RANS) turbulence model assumptions. Continuum
Center in a 4000 m3 vacuum test chamber at the Planetary Aeolian flow assumption is valid for these applications because the Knudsen
Laboratory, which is directed by Arizona State University. The number is significantly below 1 for the spacecraft’s last 20 m of
vacuum chamber has a height of 30 m and a diameter of ∼15 m, and it terminal descent.
can be evacuated to 350 Pa by a steam ejector driven vacuum system.
The chamber was backfilled with air for all tests conducted. Its A. Jet–Ground Interactions Due to Mars Science Laboratory
average temperature was approximately 280 K with an average Thruster Plumes
relative humidity of 5–10%. For our tests, the atmospheric pressure Computations were carried out using the OVERFLOW 2.1 [19]
within the vacuum chamber was varied between 0.7 and 101.3 kPa flow solver, a three-dimensional, time-marching implicit code that
(1% of full scale). Because of the steam ejectors and the large uses structured overset grid systems. The full RANS equations were
chamber volume, atmospheric pressure and temperature were solved over the entire domain using Menter’s shear-stress-transport
constant during the entire test duration. turbulence model with compressibility correction and a first-order
Dry nitrogen flows from the 14 MPa high-pressure supply time-stepping scheme. The grid system was composed of four
cylinder, which is controlled by a regulator to two reservoir tanks in structured blocks: three of cylindrical topology to model the nozzle
parallel that are set to a specified total pressure dependent on the interior, nozzle enclosure, and jet plume, and a large rectangular
throttle levels to be simulated (Fig. 5). The total pressure in the block to encompass the entire test domain, measuring 98 in. vertically
reservoir tanks is set to ∼1850 kPa for 100% MLE throttle setting. and 108 in:2 horizontally. Grids at all solid wall boundaries were
The other two settings used are 60 and 30% throttle. To minimize the modeled with viscous spacing and a y of unity. The grid block
loss of total pressure as a function of time, the regulator supplies the encompassing the plume contained 6.4 million points, and the far
needed flow rate to prevent premature choking. The solenoid valve field grid million 8 points, and they were interfaced together so as
controls had remote operations of the jet such as duration, pulse to finely resolve the plume impingement and ground jet regions.
width, pulse frequency, and total pressure within the nozzle chamber Characteristic boundary conditions were used at the domain free
(Pc ). The flow then passes through a solenoid valve, which is boundary extents, while a total-pressure, total-temperature-inflow
remotely activated, and into a converging–diverging nozzle. The boundary condition was used to satisfy chamber conditions at the
solenoid valve and propellant system performance was tested to nozzle plenum. The simulations were advanced in time until the free
ensure a relatively constant Pc that simulates the profiles from the and ground plumes had reached a quasi-steady state.
MEHTA ET AL. 2807
Downloaded by UNIVERSITY OF MICHIGAN on May 20, 2014 | http://arc.aiaa.org | DOI: 10.2514/1.J052408

Fig. 5 Views (top) and general plumbing and electrical schematic (bottom) of MSL MLE plume impingement test setup at the NASA Ames Research
Center.

B. Jet–Ground Interactions Due to the Phoenix Thruster Plumes motor tests and cold flow experimental tests. Grid independence
Two flow solvers were used to obtain numerical solutions for both is applied to both axisymmetric and 3-D meshes. Second-order
full-scale and subscale flow physics of the interaction of the spatial and first-order temporal convergence was obtained. Internal
underexpanded supersonic thruster plumes from the Phoenix REM nozzle flow solution was calculated as well. For the 3-D models, a
nozzle with the flat surface. Transient and steady-state solutions were symmetry plane is developed between the pair of thrusters to obtain
developed from these numerical solvers. Both three-dimensional the solution for a 60 deg wedge of the spacecraft [28,31]. This was
(3-D) and axisymmetric solutions were developed. The two numerical done because 180 and 60 deg wedges showed very similar flowfields
codes used were Aerosoft GASP [30] and ANSYS FLUENT. GASP and ground pressure results due to the development of stagnation
was used to model both full-scale and subscale cases. planes formed by the exhaust plume interactions of the two pairs of
Transient and steady-state RANS equations were solved by the descent thrusters [32]. This also decreased the complexity of the flow
GASP code, using both the axisymmetric and 3-D density-based domain. Four million grid cells were generated for the fine-mesh 3-D
solver. To resolve the shock waves, the Van Leer flux splitting scheme models.
is used, which is dissipative and leads to the smearing of shocks Another RANS numerical solver, ANSYS FLUENT was applied
(Walters et al. [30]). The laminar model was used for these cases. To to further confirm the results of the experimental test data. The
obtain time-accurate results, a dual implicit time-stepping derived turbulence model used was the renormalization group of k-ε [32], and
solution is selected. A single species frozen flow model is assumed. to confirm that numerical dissipation did not significantly [33,34]
Total pressure and temperature are the inlet boundary conditions, and affect the shock wave profiles, an inviscid case was also run.
the outlet is a Riemann subsonic inflow/outflow, which takes into Transient axisymmetric and 3-D solutions with a 1 μs dual implicit
consideration the potential entrainment of exhaust gases near the time-stepping scheme with adaptive grid meshing was applied to
nozzle. The pressure inlet boundary condition for the transient resolve the shock waves in the flow domain. Pressure inlet and outlet
simulations was forced by using test data from both the hot-fire rocket were the applied boundary conditions. Convergence was observed
2808 MEHTA ET AL.

Table 3 Test matrix A. Experimental Results of Subscale Mars Science


Laboratory Thruster Plumes
Test Atmospheric pressure, Ground, Height, Throttle level,
Pa deg m % For the temporal ground pressure profiles, the ground pressure rise
and settling times, maximum and steady-state pressure values,
1 800 0.0 1.73 60.00
2 733 0.0 1.73 60.00 atmospheric pressure (vacuum chamber), and Pc profiles were
3 933 0.0 1.73 100.00 recorded. All of these steady tests were performed for 1 s duration.
4 1,470 0.0 1.73 100.00 The transient ground overpressures showed an increase on the order
5 1,530 22.5 1.73 96.67 of 100 to 30% from its steady-state values for e > 2 (Fig. 7).
6 1,610 22.5 1.73 91.67 However, no repetitive ground shock frequencies were observed for
7 1,730 22.5 1.73 86.67 these cases. These overpressures span from 0.1 to ∼0.3 s during the
8 1,730 22.5 1.73 80.00 Pc rise due to engine startup as observed in Fig. 7. There is a steady-
9 1,800 22.5 1.73 73.33
10 1,840 22.5 1.73 68.33
state ground pressure regime that lasts from 0.3 to ∼1.0 s. After this
11 1,870 22.5 1.73 63.16 point, there is a simultaneous rapid ground pressure decline with a
12 1,880 22.5 1.73 56.84 sudden decrease in the Pc due to engine shutdown. For e < 2, there
13 1,930 22.5 1.73 53.68 are no characteristic overpressure, and it exhibits relatively steady
14 2,000 22.5 1.73 47.37 and much smaller pressure amplitudes, as seen from Figs. 7a and 7b.
15 2,000 22.5 1.73 44.21 At e  0.02 (Earth atmospheric pressure), the ground pressure
16 2,030 22.5 1.73 41.05 maximum is well below 0.001.
17 2,070 22.5 1.73 37.89
For the spatial ground pressure profiles, ground pressure contour
Downloaded by UNIVERSITY OF MICHIGAN on May 20, 2014 | http://arc.aiaa.org | DOI: 10.2514/1.J052408

18 2,110 22.5 1.73 34.74


19 2,470 1.9 1.73 31.58 maps were developed at the maximum overpressure value at t 
20 2,530 1.9 1.73 28.42 0.25 s (Fig. 8) and steady-state values at t  0.45 s. From these
21 101,325 22.5 1.73 100.00 contour maps at e  2.09 and e  2.83, there is observed to be a
22 101,325 0.0 1.73 100.00 ∼2d-in.-diam radial pressure footprint, which is bounded by the
sensors (Figs. 8c and 8d). There is a large normalized pressure
gradient ∇ [Eq. (26)] of 0.017 at e  2.83 determined from the
periphery of the footprint. For e  0.02, these large pressure
for second-order upwind discretization schemes for all state gradients (∇ < 0.001) are not observed, and there is a modest increase
parameters. Grid independence was also satisfied for these in pressure that spans a distance of ∼7d, as seen in Fig. 8a. For
simulations. e  0.93 and less, the pressure gradient is much smaller in magnitude
∇ < 0.0025 and more gradual, typical of a Gaussian distribution.
The normalized pressure gradient is defined as follows:
VII. Results
ΔPg d
Spatial and temporal ground pressure profiles along with Mach ∇ (26)
contours are used to analyze the flow physics of supersonic Pc Δx
impinging jets by mainly varying the jet expansion ratios and
Strouhal numbers. Correlation of this data with temporal Pc profiles Figure 9 shows how the jet expansion ratio affects normalized
and other initial conditions such as nozzle exit pressure, mass flow impingement pressure Pg ∕Pc from our experiments. There is a
rate, and thrust added insight into the physics. These values are maximum for overpressure (Pg ∕Pc  0.005) and steady ground
tabulated in Table 1. All ground pressure values Pg are normalized by pressure (Pg ∕Pc  0.015) values for e > 2. This translates to a
Pc , and physical dimensions are normalized by the nozzle exit plume–ground pressure of 27 kPa (∼4 psia) within the Mars
diameter d. atmospheric regime. The overpressure is defined as the difference

Fig. 6 Top-down schematic of the MSL descent stage with superimposed baffles to simulate outboard thrusters as implemented in experimental testbed
(left), and picture of the MSL Descent Stage (right); courtesy of NASA.
MEHTA ET AL. 2809
Downloaded by UNIVERSITY OF MICHIGAN on May 20, 2014 | http://arc.aiaa.org | DOI: 10.2514/1.J052408

Fig. 7 Temporal maximum ground pressure (peak) and PC profiles of steady impinging jets at an altitude of ∼35d for varying jet expansion ratios of
a) test 22, b) test 11, c) test 1, d) and test 2.

between maximum and steady-state ground pressure values. This is The computed Mach contours for the MLE plumes have a
an increase by a factor of 5 compared to amplitudes at e < 2. maximum Mach number of ∼10 with a shock cell length of 14d.
These amplitudes are relatively constant between e  2 to e  0.25 Approximately 2.5 shock cells are formed within the plume. As can
with a slight increase in the steady ground pressure magnitude be seen from Fig. 11, the plumes are highly collimated even to
Pg ∕Pc . A minimum in steady and overpressures values at e  0.02 is distances greater than 35d. This is a very important characteristic that
observed. will be discussed in a later section. The observed characteristic flow
Normalized ground pressure rise rates are how quickly the features are a plate shock with a diameter of ∼2d and wall jet as
normalized ground pressure values rise to the maximum amplitude observed by Lamont and Hunt [1]. Because of an oblique jet
due to engine startup. Figure 10 shows the rise rates (1∕s) as a interaction, the wall jet predominantly propagates in a x direction,
function of the jet expansion ratio. A loosely linear increase in the as shown in Fig. 11.
pressure rise rate was observed with increasing jet expansion ratio for All solutions shown are instantaneous, and hence only a snapshot
e > 0.5. The largest rise rates are observed for e > 2 with an increase of the ground spatial pressure profiles are recorded (Fig. 12). This
by a factor of ∼4 from the values compared to e < 1.5. The exhibits a radial pressure footprint with high pressure gradients of
normalized settling rates, which determine how quickly the ground 0.026 and a diameter of 1.75d. There is an asymmetry in the profile,
overpressure values settle to its steady-state values, do not show a and the highest pressure regions reach a nondimensional value of
characteristic trend with respect to the jet expansion ratio. 0.028, which is approximately 0.5d from the plume centerline. The
average ground pressure value within the footprint area is within the
B. Numerical Results of Mars Science Laboratory range of 0.018.
Thruster Plumes From Fig. 13, good agreement (within 7.5%) is observed
Numerical results are first used to compare subscale ground between numerical solutions and experimental measurements for
pressure spatial and temporal profiles and near-field plume structure normalized spatial ground pressure profiles at e ∼ 3. The pressure
with measured quantities. Once in similar agreement, Mach contours footprint diameter, normalized maximum ground pressure values,
can be used in conjunction with ground pressure profiles to and pressure gradients are similar for both simulation and
understand the flow physics. This approach was used to obtain an measurements. The numerical solution shows a slightly smaller
understanding of our observations. pressure footprint (see Sec. VII). There are some minor discrepancies
2810 MEHTA ET AL.
Downloaded by UNIVERSITY OF MICHIGAN on May 20, 2014 | http://arc.aiaa.org | DOI: 10.2514/1.J052408

Fig. 8 Ground spatial pressure profiles of steady impinging jets at an altitude of 35d and t  0.22 s for varying jet expansion ratios for test cases
presented in Fig. 5. The color bar depicts the normalized ground pressure values.

in the features such as the lack of capturing the high-pressure expansion ratios are tested: e ∼ 4.5–3.8 (moderately underexpanded)
asymmetry region, which is due to lower measurement resolution and e  0.02 (highly overexpanded) at an altitude of 8.4d. The most
than available in the numerical simulation. interesting feature in the temporal pressure profiles analysis at e 
4.4 are the transient overpressures observed during engine startup and
C. Numerical and Experimental Results of Phoenix shutdown phases [26]. These peaks demonstrated normalized rise
Thruster Plumes rates on the order of 6.0. These overpressures were repeatable and did
For experimental results in understanding pulsed jet effects not demonstrate hysteresis.
on ground pressure (nonzero Strouhal numbers), two different jet The spatial ground pressure profile for e  4.4 were also radial
with a pressure footprint diameter of ∼3.2d. As noted for the MSL

Fig. 9 Normalized maximum ground pressure vs jet expansion ratio


curve at an altitude of ∼35d. Measurement uncertainty of P g ∕Pc is Fig. 10 Rise rate vs jet expansion ratio curve at an altitude of ∼35d.
2.5% of the value. Measurement uncertainty of the rate is 2.5% of the value.
MEHTA ET AL. 2811
Downloaded by UNIVERSITY OF MICHIGAN on May 20, 2014 | http://arc.aiaa.org | DOI: 10.2514/1.J052408

Fig. 11 Full-scale and subscale MLE plume shock structure and axial static pressure profiles: a, b) numerical solutions of the Mach contour at an altitude
of ∼35d for MLE plumes at 100% throttle, and c, d) numerical solutions of Mach and static pressure profiles as a function of x∕de (axial distance along the
plume).

Fig. 12 Numerical solutions between subscale (Figs. 12b and 12d) and full-scale (Figs. 12a and 12c) spatial ground pressure profiles of MLE plume
interactions at 35d at 100% throttle. Numerical solutions of normalized ground pressure profiles (Pg ∕P c ) as a function of x∕de along the dotted lines
shown in Figs. 12a and 12b.
2812 MEHTA ET AL.

magnitude, footprint area, and pressure gradients. Good agreement is


also observed between numerical simulation and shadowgraph
imaging in the plume shock structure at the near-field regime
(Figs. 14b and 14c). The numerical simulation exhibits a plume
expansion angle from the nozzle exit plane of approximately 25 deg.
This also shows good agreement with the findings from Clark [10].
This value is somewhat less for our shadowgraph images, which
show an expansion angle of 22 deg.

VIII. Discussion
The first part of this section discusses how the plume and ground
pressure are affected by these three parameters: 1) jet expansion ratio,
Fig. 13 Numerical solution compared to measured quantities 2) supersonic core length, 3) plate shock dynamics. Then this section
at t  0.1 s and t  0.45 s for N2 jet impingement at e ∼ 3 at an discusses the comparisons between full-scale flight and subscale test
altitude of 35d (test 3). Measurement uncertainty of Pg ∕Pc is 2.5% of conditions and the applicability of the generated scaling laws. The
the value. last part focuses on how plume effects are important in understanding
plume–plume interactions, site alteration, and dust lifting.
Downloaded by UNIVERSITY OF MICHIGAN on May 20, 2014 | http://arc.aiaa.org | DOI: 10.2514/1.J052408

A. Jet Expansion Ratio


experimental measurements, for highly overexpanded jets, the The jet expansion ratio is one of the most important factors in
pressure footprint was more diffuse with a smaller magnitude and a determining ground pressure profiles at high altitudes (h > 5d). The
Gaussian distribution and did not indicate large pressure gradients as jet expansion ratio influences the near-field and far-field plume
observed for moderately underexpanded jets. The spatial pressure structure. For e > 1 (underexpanded jets), the expansion fans form at
profile for moderately underexpanded jets is non-Gaussian, and there the lip of the nozzle, causing the plume to expand outward with
are minor pressure peaks observed at 1.6d due to the effects of respect to the normal increasing the plume expansion angle
oblique tail shocks. (θ > 0 deg). For e < 1 (overexpanded jets) leads to reflected oblique
Transient and steady-state solutions were developed for the pulsed shock waves that cause a decrease in the plume expansion angle with
supersonic jets impinging at the surface at various altitudes. respect to the normal (θ < 0 deg). As a result of the low planetary
Simulations at two altitudes are presented here: 8.4d (Fig. 14) and atmospheric pressure on Mars and the moon, most of the thruster jets
25d (Fig. 15). Numerical simulations show that an underexpanded observed during spacecraft landings on these celestial bodies is
supersonic jet with an e  4.4 at an altitude of 25d results in the underexpanded.
development of a normal plate shock with a diameter of ∼2d and wall From Fig. 9, it can be seen that the normalized ground pressure
jets that propagate in the x direction (Fig. 15). Two shock cells are value increases by a factor of 5 for jet expansion ratio greater than 2. It
observed with a length of ∼12d within the plume structure at an can also be seen that for very low e on the order of 0.02, the ground
altitude of 25d. Here, once again, the plume structure is collimated as pressure was minimal. Numerical simulations that were validated by
observed for the MLE thruster plume numerical simulations. experimental tests show in Fig. 15 that the Mach contours and plume
According to Figs. 14a and 14b, the numerical simulations and structure are also considerably different between e ∼ 4.5 and
experimental measurements at an altitude of 8.4d show good e  0.02. For moderately underexpanded jets (e ∼ 4.5), the plume
agreement (within an average of 10%) in both temporal and spatial structure has compressed, and collimated shock cells that are formed
ground pressure profiles at the far-field and near-field regimes for until a downstream distance is reached when plume and ambient
underexpanded jet at e ∼ 4.4. They both show ground overpressures static pressure are in equilibrium. This downstream distance,
at a 20 Hz frequency and similar quasi-steady ground pressure supersonic core length x, is considerably larger for moderately under-

Fig. 14 Comparison of axisymmetric numerical simulations and experimental measurements of pulsed supersonic N2 jet interactions at h∕De  8.4:
a) temporal centerline ground pressure (dashed lines) and Pc (solid line) profiles, and b) spatial ground pressure profiles and Mach contour (Gulick [32]).
Shadowgraph image of a near-field underexpanded jet (e ∼ 4.4) at <5 ms during engine startup (Fig. 14c, part 1) and during full-throttle (point 2) and
numerical solution during full-throttle (point 3). Measurement uncertainty of P∕Pc−max is 2.5% of the value.
MEHTA ET AL. 2813
Downloaded by UNIVERSITY OF MICHIGAN on May 20, 2014 | http://arc.aiaa.org | DOI: 10.2514/1.J052408

Fig. 15 Numerical solution of velocity contour comparison between steady underexpanded (Mars) and highly overexpanded (Earth) supersonic
jets at an altitude of 25d (left), and normalized centerline ground pressure vs jet expansion ratio for nitrogen jets and rocket plumes at an altitude of ∼35d
(right).

expanded jets as compared to highly overexpanded jets (e ∼ 0.02), as (e < 0.5) and underexpanded jets (e ≫ 5). Hence, to obtain accurate
shown in Fig. 15. This classification of jets may develop a stable plate risk assessment of spacecraft landings on Mars and the moon without
shock at the surface and a propagating wall jet. changing the thrust conditions, it is critical to study these interactions
The relatively larger ambient pressure for highly overexpanded at the appropriate atmospheric environments so that the jet expansion
jets (e ≪ 1) leads to the formation of reflected oblique and normal ratio is accurately simulated.
shock waves, which may occur at the diverging section of the nozzle. This analysis of jet expansion ratio is quite important to consider
Boundary-layer separation at the diverging section can lead to because these values are dependent on the propulsion system and
attenuation of the jet [35]. Once a strong normal shock wave forms at atmospheric density of the planets for our application. The effects of
the diverging section (Fig. 15), this results in the propagation of weak thruster plumes on the surface can be decreased by changing the
shock waves, which leads to flow separation and shock-wave propulsion system requirements or by landing at a different latitude
instability. The shock wave is further attenuated by the interaction and longitude on planets. For example, the atmospheric pressure on
and mixing of the shock with the dense shear layer at the jet boundary, Mars can change from ∼350 to 1000 Pa depending on the location
which leads to Kelvin–Helmholtz instabilities. Because of over- and time of day.
expansion, the atmospheric free surface boundary pinches the jet
inward, leading to an increase in the turbulent mixing layer with axial B. Supersonic Core Length
distance. This decreases x and results in the rapid decay of the plume Supersonic core length can be inferred from ground pressure
structure to a fully turbulent subsonic jet with a linear spreading profiles (Fig. 9). An indirect approach in measuring the supersonic
profile [30]. As a result of the spreading profile, large altitudes, and core length is by varying the distance of the engine from a flat surface
subsonic flowfield, normal plate shocks are not developed above the and monitoring the ground pressure. Supersonic core that propagates
surface. This prevents the formation of large pressure gradients at the to the flat surface results in the formation of a plate shock, which leads
surface as well as the formation of supersonic wall jets as shown to steep pressure gradients and relatively large pressure magnitudes
in Fig. 9. as observed in Fig. 8. We conclude from the previous section that
For e ≫ 1, highly underexpanded jets shown in Fig. 15, there is a there are large differences in the supersonic core length and spatial
decrease in normalized ground pressure profiles, and this is due to a ground pressure profiles for highly overexpanded and moderately
large plume expansion angle. The shock propagation of a large underexpanded jets at an altitude of 35d. In contrast to moderately
expansion plume results in a large areal plate shock as observed by underexpanded jets, a very diffuse and Gaussian pressure profile is
Clark [10], which significantly reduces the normalized ground observed for highly overexpanded jets (e  0.02), as depicted in
pressure because pressure is inversely dependent on area. Another Fig. 8, and this leads to rapid decay of the plume structure, which
mechanism may be due to the increased pressure losses due to a demonstrates a core length of less than ∼5d, as shown in Fig. 15. The
normal shock wave or a Mach disk formed within the near-field experiments indicate that the supersonic core length developed by
regime at high e as opposed to the unsteady oblique shock waves MLE plumes with the Mars atmosphere (e ∼ 3) propagates to at least
developed in the far-field regime at lower jet expansion ratios. a distance of 37d. This is also supported by Mach contours generated
We previously presented data of the comparison of supersonic by the numerical simulations (Fig. 11). Numerical solutions show
nitrogen jets and rocket plumes from monopropellant and bipro- that REM plumes (e ∼ 4.4) have a supersonic core length of at least
pellant rocket motors at an altitude of ∼35d (Fig. 15). All nitrogen jets 25d (Fig. 16). This is further supported by Inman et al. [13], who
and rocket plumes demonstrate an exit Mach number of ∼5 show with PLIF imaging that a collimated moderately under-
[10,28,36]. The monopropellant rocket motors use hydrazine as the expanded (e  5.4) turbulent nitrogen jet at Mach 2.6 has a
fuel, which combusts at a T C of greater than 1000 K, releasing N2 , H2 , supersonic core length to a distance of 31d. Scroggs and Settles [8]
and NH3 as low-density exhaust species. The bipropellant uses show with schlieren imaging that the supersonic core length increases
methyl hydrazine as the fuel and nitric oxides as the oxidizer [35]. with jet expansion ratio and Mach number. They recorded supersonic
Good agreement (within 12.5%) is shown in the trends between core lengths on the order of 35d for e  4 and a Mach number of 2.2.
nitrogen jets and rocket plumes in which the highest normalized Although the numerical simulations do not capture the eddy
ground pressure values have jet expansion ratios between 2 and 5, transport, which may decrease the supersonic core length, the
as observed in our studies (Fig. 15). Rocket plumes also exhibit experiment observations show good agreement with the RANS-
minimal normalized ground pressure values for highly overexpanded based simulations.
2814 MEHTA ET AL.

For a fixed plume gas and chamber temperature, the length of the due to the MLE plumes are smaller than observed for the Phoenix
shock cell xs is determined to be a function of the jet expansion ratio. cases and virtually absent during the engine shutdown phase, and this
An increase in the shock cell length is seen at higher NPR than at may be due to the development of a much weaker normal shock at the
lower corresponding values [10]. The shock cell length increases surface. This could be attributed to a larger axial distance the shock
logarithmically with increasing jet expansion ratio and increases waves need to travel and the slope of the inclined surface. This could
linearly with nozzle exit diameter. Shock cell length may have a weak also be attributed to slower stagnation pressure rise and fall rates.
dependence on Me . Large instabilities in the plate shock may delay the formation of a
fully developed shock. Most importantly, the characteristic ground
C. Plate Shock Dynamics shock frequency observed for pulsed jets was entirely absent in these
steady jet cases. This is mainly the result of engine cycling. However,
Plate shock dynamics lead to large ground pressure fluctuations
during plate shock formation and collapse, high instability due to
and gradients at the surface. Figure 17 shows a numerical solution for
axial plate shock oscillations occurs. This instability is further
the formation of a plate shock at the surface from a Mach 4.7 supported by the highly variable settling trend for e > 1.5. As a
underexpanded jet (e ∼ 4.4). Before initial jet impact, shock waves result, plate shock formation is also a function of the jet expansion
accelerate toward the surface, which initiates coalescing of the plume ratio and altitude.
density and gas compression. After jet impingement, an unstable From observations (Figs. 8 and 9), a stable plate shock may not
normal shock wave (plate shock) is formed above the surface and a form for supersonic jets at e below 1.75 at altitudes of 35d and greater.
transient high plume density is observed below the plate shock. This As we approach a jet expansion ratio of 1, shock cells within the
concentrated and localized plume density at high velocity results in plume disappear and lead to greater shock-wave and static pressure
Downloaded by UNIVERSITY OF MICHIGAN on May 20, 2014 | http://arc.aiaa.org | DOI: 10.2514/1.J052408

larger ground pressures relative to subsonic flowfields. Hence, attenuation at these large distances. This is also supported by Inman
the large differences in ground pressure observed with varying jet et al. [13], which also shows a significant drop in the ground pressure
expansion ratios. Overpressure due to plate shock formation is at jet expansion ratios near and below unity.
observed for both steady (MLE, Fig. 7) and pulsed (REM, Fig. 14) A significant decrease in normalized ground pressure magnitude
exhaust plumes. Overpressure occurs when the shock wave first for e < 2 may be attributed to the plate shock within the expansion
impinges on the surface and the overpressure settles to quasi-steady- regime of the shock cell. Large pressure fluctuations occur depending
state value upon stable formation of the plate shock. The non- on whether the plate shock is within the expansion or compression
Gaussian and high-pressure gradients are characteristic of plate shock regimes of the shock cell [1]. This is unlikely the result of our study
formation, and the footprint is similar to the shock diameter. due to the fact that successive points with decreasing jet expansion
There are some major differences in the shock dynamics between ratios ranging from e  1.90 to e  0.30 (changing shock structure
pulsed and steady (MSL) descent jets. The initial overpressure peak profiles) lead to relatively constant normalized ground pressure

Fig. 16 Numerical solutions of the Mach contours of the Phoenix REM plumes at an altitude of 25d [32] (left), and normalized plate shock diameter and
shock cell length as a function of the jet expansion ratio at an altitude of ∼35d (right). Lines are polynomial fits to the data points.

Fig. 17 Numerical solutions of) Mach, and b) density contours of a pre- and postnormal shock-wave interaction. The black bar depicts the length of the
nozzle exit diameter.
MEHTA ET AL. 2815

magnitudes. Hence, the significant decrease in ground pressure The ground pressure magnitude for the MSL cold gas jets is greater
(Figs. 9 and 15) is mainly due to the inability for a plate shock to than the rocket plumes by a factor of 0.25. The plumes at subscale
develop at these expansion ratios and large downstream distances. conditions are more compacted and collimated than for the full-scale
From Fig. 16, the plate shock diameter is independent of both NPR case. The subscale jets also exhibit more frequent and shorter shock
and Me but shows logarithmic dependence on the jet expansion ratio cells within the collimated plume. This is attributed to a difference in
and linear dependence on the nozzle exit diameter. These length-scale the jet expansion ratio (Table 2).
profiles were determined by both numerical and experimental Subscale experimental spatial and temporal ground pressure
observations. measurements show relatively good agreement with the full-scale
numerical solutions for rocket plume impingement. The Phoenix and
D. Comparisons Between Subscale and Full-Scale Tests MSL test cases show few discrepancies but overall show similar
As discussed in Sec. VII, numerical simulations and experimental pressure footprint area, normalized magnitude, and pressure gradients.
data show good qualitative and quantitative agreement between the Because of limited pressure sensors and a decrease in resolution, the
MSL and Phoenix subscale test programs and flight conditions. To test data were not able to capture the asymmetry. Because of limitations
ensure similar plume structure and ground pressure profiles between in obtaining low vacuum in the chamber, the test was not able to
our subscale tests and exhaust plumes from both MSL and Phoenix achieve exactly the same jet expansion ratio for the MLE cases as seen
rocket engines, scaling parameters derived previously were used. for the full-scale numerical simulations (Table 2). As a result of this
There, parameters govern the flow regime, specific plume energy extensive study, these scaling parameters applied to cold gas jets are
critical in properly simulating rocket plume impingement effects.
density, compressibility, and unsteady effects. Our goal is to
Further controlled experimental studies between full-scale rocket
Downloaded by UNIVERSITY OF MICHIGAN on May 20, 2014 | http://arc.aiaa.org | DOI: 10.2514/1.J052408

determine whether these theoretical scaling parameters accurately


motors and supersonic cold gas jets are needed to confirm these results.
determine the plume flow dynamics from strictly a momentum
Similarity in ground pressure profiles between subscale test data
perspective between cold gas jets and rocket plumes. These
and numerical results for the flight conditions is critical to adequately
parameters are approximately similar between subscale and full-scale
simulate the plume-induced landing site alteration caused by the
cases for both test programs (Table 2): Phoenix REM and MSL MLE.
Phoenix and MSL descent engines. Hence, this paper shows that this
Experimental measurements and numerical solutions show that the
condition and the required scaling laws are satisfied, and as a result,
Mach number, shock structure, and spatial and temporal pressure
subscale site-alteration tests were conducted for the Phoenix [24] and
profiles are in good agreement (within 12.5%) between cold flow
MSL spacecraft. This was one of the main reasons for the analyses
subscale and full-scale systems, Phoenix rocket exhaust plumes, at
shown in this paper.
altitudes of 25d and 8.4d. This is observed for both single- and dual-
First, full-scale descent engines were numerically modeled, and
thruster systems. For example, Fig. 18 shows a numerical solution of
good agreement is observed with data from subscale experiments.
a full-scale temporal and spatial ground pressure profiles for a full- This was an important result because it partially confirmed the
scale rocket plume interaction from a single Phoenix descent engine. validity of our theoretical scaling laws. Full-scale rocket test firing at
Note the similarity in comparison with the subscale case shown in simulated Martian conditions that record both temporal and spatial
Fig. 14. The numerical solutions of the Mach contours between the ground pressures are needed to fully validate these laws.
subscale and full-scale MLE plumes show relatively good agreement Once numerical and experimental results showed relatively good
(within 40%) in both magnitude and shock structure (Fig. 11). The temporal and spatial agreement, the numerical code GASP was used
numerical solutions and experimental observations of the spatial to model full-scale 3-D cases for adjacent thruster plumes impinging
ground pressure profiles between the MLE subscale and full-scale at the surface. The computational domain spans a 60 deg wedge
cases also show relatively good agreement (within 40%) in where two engines are modeled and is bounded by symmetry planes
pressure footprint area, normalized magnitude, asymmetry, and [32]. This domain is then extrapolated to obtain the Mach and
pressure gradients (Fig. 12). pressure contour profile of the full 360 deg Phoenix Lander, as shown
Numerical simulations were used to characterize the effects of in Fig. 19. Because of adjacent plume interactions and nonlinear
rocket plumes on the ground at various altitudes. There are no large shock–shock interactions as described in the previous section, the
discrepancies observed for the numerical solutions of the Phoenix plate shock demonstrates noncoplanarity and oscillates in three axes,
test cases, but the MSL test cases did show some minor differences. leading to both asymmetric high-pressure regions (Fig. 19a) and
ground pressure fluctuations (Fig. 19b) during the quasi-steady
regime. Most importantly, characteristic overpressure peaks are
observed during rapid engine startup and shutdown, suggesting the
mechanism of plate shock formation and collapse. These numerical
simulations show that the spatial ground pressure profiles between
the full-scale (Fig. 19) and subscale (Fig. 14) systems show good
agreement in trends and further confirm the use of these scaling laws.

E. Asymmetries Due to Adjacent Plume Interactions


Although most of the studies presented thus far are primarily
concerned with single jet interactions, we will briefly discuss the
ground pressure and shear stress (τ) behavior due to two adjacent
underexpanded N2 jets with a nozzle spacing of x∕De  0.1, similar
in geometric configuration as adjacent Phoenix descent engines. A 3-
D steady-state numerical simulation was developed to understand
possible asymmetries at the surface. In contrast to the radial
symmetry of single subsonic and supersonic jet interactions [26],
adjacent jets develop large asymmetries in both ground pressure
and wall shear-stress parameters, as shown in Fig. 20. There are no
symmetric ringlike contour profiles but rather asymmetric semi-
circular high-pressure and shear-stress regions. This may be
Fig. 18 Full-scale (singe Phoenix REM plume interaction) numerical attributed to shock–shock Riemann interactions [36], which develop
results of the spatial (t  0.036 s) and temporal normalized ground a merged plate shock with higher strength at both the near-field and
pressure profiles at an altitude of 8.4d at Mars atmospheric environ- far-field regimes. Hence, the plate shock is almost twice in diameter
ment [32]. as observed for a single jet case. This merging could lead to a
2816 MEHTA ET AL.

Fig. 19 3-D numerical simulation of full-scale interacting REM plumes at Mars atmosphere: a) Mach and normalized pressure contours of steady REM
plumes interacting at the surface at an altitude of h∕De  8.4 and 0 deg slope [32], and b) centerline ground pressure and P c temporal profiles due to 0.1 s
jet pulse at an altitude of h∕De  21.9 and 0 deg slope. The dashed line is the centerline ground pressure solution for steady-state numerical simulation
[32].
Downloaded by UNIVERSITY OF MICHIGAN on May 20, 2014 | http://arc.aiaa.org | DOI: 10.2514/1.J052408

Fig. 20 3-D numerical simulation of two adjacent underexpanded N2 jets interacting at the surface at h∕De  25. Contours of a) wall shear stress
normalized by ambient pressure, and b) normalized ground pressure; c, d) τ and ground pressure profiles along the x and y centerlines; and e) Mach
contour.
MEHTA ET AL. 2817

noncoplanar development of the normal plate shock, resulting in


asymmetry in ground pressure. This may also affect both the tail
shock and the flow behavior of the supersonic wall jets, leading to
asymmetry in the wall shear stress. These wall jets are caused by flow
expansion, developing mainly from flow across the tail shock and
propagating along the surface (Fig. 20). Hence, these results
demonstrated the need for 3-D numerical simulations of the six pairs
of Phoenix REM plume interacting at the surface. As a result,
asymmetry due to adjacent plume interactions can lead to larger soil
erosion rates as compared to a single jet.

F. Site Alteration and Dust Lifting


Various mechanisms due to the rocket plume interactions can lead
to significant soil erosion and dust lifting. The overpressures
described previously and the high surface shear stress associated with
the supersonic wall jets can lead to soil erosion and dust lifting.
Large and rapid pressure fluctuations might cause soil liquefaction
[24]. Soil liquefaction is defined as the fluidlike state of granular
Downloaded by UNIVERSITY OF MICHIGAN on May 20, 2014 | http://arc.aiaa.org | DOI: 10.2514/1.J052408

media. Ground shock vibrations caused by these large transient


overpressures superimposed by the pulsing quasi-steady-state
overpressure regions may disrupt the soil and break the particle-to-
particle cohesive forces as observed for the Phoenix ground test data
[24]. This can decrease the bearing capacity, maximum average
contact pressure between the foundation and the soil to prevent shear
failure, and increase the fluidization of the soil, possibly leading to
lateral ground failure and crater formation [24]. The extent of ground
failure depends on the soil properties, surface impingement temporal
and spatial pressure profiles, ground shear stress, and the dynamic
interactions between the thruster plume and the ground [24]. The
minimum ground pressure threshold before erosion takes place at a Fig. 21 High variability of ground pressure vs altitude profiles due to
simulated Martian environment is approximately ∼2 kPa for soil shock-wave interactions: a) limited experimental results compared to NASA
White Sands Test Facility and NASA Langley Research Center studies
similar to dune sand or lunar nominal [28].
using a single nozzle [28]; and b) numerical result for full-scale Phoenix
Also, depending on whether the shock cell is within the com- REM plume interactions at both steady-state and transient conditions [32].
pression or expansion zone near the surface results in a nonlinear
behavior of ground pressure with altitude, which presents itself
differently than for the asymptotic profiles of subsonic jets [24]. These IX. Conclusions
pressures are relatively independent of altitude and oscillate around a This research investigation was ultimately undertaken to assess
mean value for small h∕De (<40). The overpressure and quasi-steady- landing site alteration due to rocket plume impingement during
state ground pressure can significantly change in magnitude with spacecraft landings. The first approach, presented here, is to provide
altitude, as can be seen for the two nondimensional altitude cases insight into plume shock structure and dynamics and their effect on
presented here: h∕De  8.4 and 25. This high variability is further ground pressure profiles. From extensive numerical and experi-
shown in Fig. 21. Good qualitative agreement is observed between mental analyses presented in this paper, moderately underexpanded
experimental observations of cold gas simulations and 3-D numerical jets (e between 2 and 5) are shown to demonstrate collimated shock
results of full-scale systems [32]. This variability may also lead to structures, compact radial pressure footprints, large supersonic core
further disturbance of soil during spacecraft descent. lengths, plate shock dynamics, and maximum pressure loads. For e
According to Mehta et al. [24] and Metzger et al. [37], there are less than 2 and greater than 10, we illustrate a significant decline in
five main mechanisms that lead to site alteration and dust lifting due the ground pressure loads by a factor of 4 with large Gaussian
to plume interactions: 1) bearing capacity failure, 2) viscous erosion, pressure footprints, which is mainly attributed to large changes in the
3) diffused gas erosion, 4) diffusion-driven flow, and 5) diffusive gas plume shock structure. The plate shock dynamics is shown to be
explosive erosion. Through extensively developed scaling laws that responsible for the following effects at the surface; it increases the
address these five mechanisms, scaling Earth-gravity-based tests for pressure gradients, fluctuations, and average magnitudes and
Mars conditions can be achieved. Diffusive gas explosive erosion develops pressure asymmetry and overpressures. There is sensitivity
played a significant role in site alteration and dust lifting due to in both shock structure and ground pressure dynamics due to both the
Phoenix’s pulsed thrust impact on the Martian polar surface [24]. jet expansion ratio and Strouhal number (pulsed or steady). Most
Because of the temporal and spatial pressure profiles (Figs. 8 and 9) importantly, the flow dynamic studies at Mars atmosphere (e ∼ 2–5)
and the projected soil properties on Mars, bearing capacity failure is show that rocket plumes possess large collimated shock cells greater
the most likely erosion mechanism to occur at the MSL landing site. than ∼10d and generate maximum ground pressure load at even high
Because of the importance of understanding ground pressure profiles spacecraft altitudes of ∼35d with respect to either lunar or terrestrial
of thruster plume interactions, further studies were conducted in atmospheric regimes. This leads to larger soil erosion for the Mars
support NASA’s new Mars mission. landing spacecraft relative to lunar and terrestrial landings. Hence,
The main parameter that is used to determine the extent of jet- extensive numerical and experimental investigations of plume
induced soil erosion is the surface pressure profile of the impinging interactions with large variations in the scaling parameters defined in
jet or rocket plume [24]. Soil pore pressure and soil properties are Table 2 are needed to reduce mission risk associated with landing
directly dependent on the surface pressure. However, this investiga- spacecrafts on Mars.
tion shows that, along with surface pressure magnitude and spatial Good agreement (within 12.5%) is observed between measure-
profiles, the temporal pressure profile is also a critical parameter in ments and numerical solutions for both subscale rocket engine
determining the flow physics of the exhaust gas within the granular module and main landing engine nozzle plumes at various altitudes.
media. To simulate surface pressure profiles, it is important to This provides a comparison with the numerical solvers and provides
accurately simulate both the thruster inlet stagnation pressure and the insight to the flow physics occurring at the surface. Relatively good
atmospheric pressure environments of planetary bodies. agreement in the numerical calculations is also observed in shock
2818 MEHTA ET AL.

structure and ground pressure dynamics between subscale and full- [15] Donaldson, C. D., Snedeker, R. S., and Margolis, D. P., “A Study of Free
scale systems. Most importantly, these results show that the scaling Jet Impingement Part 2: Free Jet Turbulent Structure and Impingement
laws developed can simulate the ground interaction physics due to Heat Transfer,” Journal of Fluid Mechanics, Vol. 45, No. 3, 1971,
rocket exhaust plumes. pp. 477–512.
doi:10.1017/S0022112071000156
Now that extensive plume–ground pressure analysis has been [16] Stitt, L. E., “Interaction of Highly Underexpanded Jets with Simulated
conducted for the Mars Science Laboratory (MSL) descent stage Lunar Surfaces,” NASA TN-D-1095, 1963.
spacecraft, a site-alteration map will be created by various MSL [17] Schlichting, H., and Gersten, K., Boundary Layer Theory, Springer,
cameras of the postlanding site. More adequate analyses of the effects New York, 2001, pp. 145–155.
of plume–ground pressure loads on the Martian soil and environment [18] Anderson, J. D., Modern Compressible Flow, McGraw–Hill, New York,
can be characterized for this mission and future Mars missions. The 2008, pp. 209–210.
main focus of this research is to provide insight into plume–surface [19] Buning, P. G., Chan, W. M., Renze, K. Z., Sondak, D. L., Chiu, I.-T., and
interactions and to demonstrate the importance of modeling and Slotnick, J. P., “OVERFLOW User’s Manual,” Ver. 1.6ab, NASA Ames
testing of these complex flow physics. This is used to provide Research Center, Moffett Field, CA, Jan. 1993.
[20] Van Norman, J. W., and Novak, L., “CFD Support of MSL MLE Plume
accurate landing environment data to minimize potential future Simulations,” NASA CR-NNL04AA03Z, 2009.
hazards during spacecraft landings on planetary bodies. [21] Land, N., and Scholl, H., “Scaled LEM Jet Erosion Tests,” NASA
Working Paper LWP-252, 1966.
[22] Clark, L., and Conner, W., “Exploratory Study of Scaled Experiments to
Acknowledgments Investigate Site Alteration Problems for Martian Soft-Lander
This research was supported by the NASA Graduate Student Spacecraft,” NASA Working Paper LWP-765, 1969.
Downloaded by UNIVERSITY OF MICHIGAN on May 20, 2014 | http://arc.aiaa.org | DOI: 10.2514/1.J052408

Researchers Program grant NNX06AH56H. Special thanks to Ron [23] Choutapalli, I., Krothapalli, A., and Arakeri, J. H., “An Experimental
Study of an Axisymmetric Turbulent Pulsed Air Jet,” Journal of Fluid
Greeley of Arizona State University, Ken Smith of the Planetary
Mechanics, Vol. 631, 2009, pp. 23–63.
Aeolian Laboratory at NASA Ames Research Center and the Entry, doi:10.1017/S0022112009007009
Descent, And Landing and Advanced Technologies Division at the [24] Mehta, M., Renno, N. O., Marshall, J., Grover, M. R., Sengupta, A.,
NASA Jet Propulsion Laboratory. Rusche, N. A., Kok, J. F., Arvidson, R. E., Markiewicz, W. J., Lemmon,
M. T., and Smith, P. H., “Explosive Erosion During the Phoenix Landing
Exposes Subsurface Water on Mars,” Icarus, Vol. 211, No. 1, 2011,
References pp. 172–194.
[1] Lamont, P. J., and Hunt, B. L., “The Impingement of Underexpanded doi:10.1016/j.icarus.2010.10.003
Axisymetrical Jets on Perpendicular and Inclined Flat Plates,” Journal [25] Desai, P. N., Prince, J. L., Queen, E. M., Cruz, J. R., and Grover, M. R.,
of Fluid Mechanics, Vol. 100, No. 3, 1980, pp. 471–475. “Entry, Descent and Landing Performance of the Mars Phoenix Lander,”
doi:10.1017/S0022112080001255 AIAA/AAS Astrodynamics Specialist Conference. and Exhibit, AIAA
[2] Fujii, K., Tsuboi, N., and Fujimatsu, N., “Visualization of Jet Flows over Paper 2008-7346, Aug. 2008.
a Plate by Pressure-Sensitive Paint Experiments and Comparison with [26] Plemmons, D. H., Mehta, M., Clark, B. C., Kounaves, S. P., Peach, L. L.
CFD,” Annals of the New York Academy of Sciences, Vol. 972, No. 1, Jr., Renno, N. O., Tamppari, L., and Young, S. M. M., “Effects of the
2002, pp. 265–270. Phoenix Lander Descent Thruster Plume on the Martian Surface,” Journal
doi:10.1111/nyas.2002.972.issue-1 of Geophysical Research: Planets, Vol. 113, No. E3, 2008, Paper E00A11.
[3] Hunt, J. C. R., “Industrial and Environmental Fluid Mechanics,” Annual doi:10.1029/2007JE003059
Review of Fluid Mechanics, Vol. 23, 1991, pp. 1–42. [27] Sengupta, A., Kulleck, J., Sell, S., Van Norman, J., Mehta, M., and
doi:10.1146/annurev.fl.23.010191.000245 Pokora, M., “Mars Landing Engine Plume Impingement Environment
[4] Lamont, P. J., and Hunt, B. L., “The Impingement of Underexpanded of the Mars Science Laboratory,” IEEE/AIAA Aerospace Conference,
Axisymmetric Jets on Wedges,” Journal of Fluid Mechanics, Vol. 76, Paper 1349, IEEE Publ., Piscataway, NJ, 2009.
No. 2, 1976, pp. 307–336. [28] Romine, G. L., Reisert, T. D., and Gliozzi, J., “Site Alteration Effects
doi:10.1017/S0022112076000657 from Thruster Exhaust Impingement During a Simulated Viking Mars
[5] Donaldson, C. P., and Snedeker, R. S., “A Study of Free Jet Landing: Nozzle Development and Physical Site Alteration,” NASA
Impingement. Part 1. Mean Properties of Free and Impinging Jets,” CR-2252, 1973.
Journal of Fluid Mechanics, Vol. 45, No. 2, 1971, pp. 281–319. [29] Huseman, P. G., and Bomba, J., “CFD Analysis of Terminal Descent
doi:10.1017/S0022112071000053 Plume Impingement for Mars Landers,” 34th AIAA Thermophysics
[6] Henderson, B., Bridges, J., and Wernet, M., “An Experimental Study of Conf., AIAA Paper 2000-2501, 2000.
the Oscillatory Flow Structure of Tone-Producing Supersonic Impinging [30] Walters, R. W., Cinnella, P., and Slack, D. C., “Characteristic-Based
Jets,” Journal of Fluid Mechanics, Vol. 542, 2005, pp. 115–137. Algorithms for Flows in Thermo-Chemical Non-Equilibrium,” AIAA
doi:10.1017/S0022112005006385 Journal, Vol. 30, No. 5, 1993, pp. 1304–1313.
[7] Carling, J. C., and Hunt, B. L., “The Near Wall Jet of a Normally doi:10.2514/3.11065
Impinging, Uniform Axisymmetric, Supersonic Jet,” Journal of Fluid [31] Van Leer, B., “Flux-Vector Splitting for the Euler Equations,” Lecture
Mechanics, Vol. 66, No. 1, 1974, pp. 11–19. Notes in Physics, Vol. 170, 1982, pp. 507–512.
doi:10.1017/S0022112074000127 doi:10.1007/3-540-11948-5_66
[8] Scroggs, S. D., and Settles, G. S., “An Experimental Study of Supersonic [32] Gulick, D. S., “Phoenix Mars Lander Descent Thruster Plume/Ground
Microjets,” Experiments in Fluids, Vol. 21, No. 6, 1996, pp. 401–409. Interaction Assessment,” Lockheed Martin Interoffice Memo-WS-06-
doi:10.1007/BF00189042 002, Denver, CO, 2006.
[9] Krothopalli, A., Rajkuperan, E., Alvi, R., and Lourenco, L., “Flow Field [33] Nichols, R. H., Turbulence Models and Their Application to Complex
and Noise Characteristics of a Supersonic Impinging Jet,” Journal of Flows, Rev. 3.0, Univ. of Alabama at Birmingham, Birmingham, AL, 2000.
Fluid Mechanics, Vol. 392, 1999, pp. 174–188. [34] Papp, J. L., and Ghia, K. N., “Study of Turbulent Compressible Mixing
doi:10.1017/S0022112099005406 Layers Using Two-Models Including RNG k-ε Model,” 36th AIAA
[10] Clark, L. V., “Effect of Retrorocket Cant Angle on Ground Erosion—A Aerospace Sciences Meeting and Exhibit, AIAA Paper 1998-0320, 1998.
Scaled Viking Study,” NASA TM-X-2075, 1970. [35] Shimshi, E., Ben-Dor, G., and Levy, A., “Viscous Simulation of Shock-
[11] Roberts, B., Wallace, R. O., and Sims, J. L., “Plume Base Flow Reflection Hysteresis in Over-Expanded Planar Nozzles,” Journal of
Simulation Technology,” NASA TR-19820023542, 1982. Fluid Mechanics, Vol. 635, 2009, pp. 189–206.
[12] Sutton, G. P., and Biblarz, O., Rocket Propulsion Elements, 7th ed., doi:10.1017/S002211200900771X
Wiley, New York, 2001, pp. 639–652. [36] Latto, W. T., Jr., and Stitt, L. E., “Highly Underexpanded Exhaust
[13] Inman, J. A., Danehy, P. M., Nowak, R. J., and Alderfer, D. W., Jets Against Adjacent Surfaces,” NASA TR-7542, 1963.
“Fluorescence Imaging Study of Impinging Underexpanded Jets,” 46th [37] Metzger, P. T., Immer, C. D., Donahue, C. M., Vu, B. M., Latta, R. C. III,
AIAA Aerospace Sciences Meeting and Exhibit, AIAA Paper 2008-619, and Deyo-Svendsen, M., “Jet-Induced Cratering of a Granular Surface
Jan. 2008. with Applications to Lunar Spaceports,” Journal of Aerospace
[14] Hagermann, G., and Frey, M., “Shock Pattern in the Plume of Rocket Engineering, Vol. 22, No. 1, 2009, pp. 24–32.
Nozzles: Needs for Design Consideration,” Shock Waves, Vol. 17, No. 6,
2008, pp. 387–395. T. Jackson
doi:10.1007/978-3-540-85168-4_8 Associate Editor

Вам также может понравиться