You are on page 1of 13

See discussions, stats, and author profiles for this publication at: https://www.researchgate.

net/publication/319128833

Numerical Simulations of Flow Past a Submarine in Extraterrestrial, Cryogenic


Seas

Conference Paper · September 2017


DOI: 10.2514/6.2017-5204

CITATIONS READS
3 306

7 authors, including:

Shane Carberry Mogan Cyril Jerome Bernardo


New York University Northrop Grumman, Linthicum, United States
14 PUBLICATIONS   35 CITATIONS    2 PUBLICATIONS   6 CITATIONS   

SEE PROFILE SEE PROFILE

Iskender Sahin Jason Hartwig


New York University NASA
40 PUBLICATIONS   223 CITATIONS    96 PUBLICATIONS   725 CITATIONS   

SEE PROFILE SEE PROFILE

Some of the authors of this publication are also working on these related projects:

SPH Sloshing Group View project

LUNARPORT View project

All content following this page was uploaded by Shane Carberry Mogan on 18 September 2017.

The user has requested enhancement of the downloaded file.


AIAA 2017-5204
AIAA SPACE Forum
12 - 14 Sep 2017, Orlando, FL
AIAA SPACE and Astronautics Forum and Exposition

Numerical Simulations of Flow Past a Submarine in


Extraterrestrial, Cryogenic Seas

Shane R. Carberry Mogan1, Damon Chen1, Cyril Bernardo1, Iskender Sahin2, Angelantonio Tafuni2
NYU Tandon School of Engineering, Brooklyn, NY 11201

and

Jason W. Hartwig3, Steven R. Oleson4


NASA Glenn Research Center, Cleveland, OH 44135
Downloaded by NEW YORK UNIVERSITY on September 18, 2017 | http://arc.aiaa.org | DOI: 10.2514/6.2017-5204

A computational study is conducted to evaluate the performance of an autonomous


submarine to operate in the liquid hydrocarbon seas of Titan, Saturn’s largest moon,
to simulate the flow around the submarine and offer a prediction for the power
requirements. Computational Fluid Dynamics is used for deeply submerged and free-
surface conditions. In the deeply submerged case, mesh-based code is utilized to show
velocity and pressure contours, and to calculate the total drag force. Specific cases of
the submarine navigating the free-surface are analyzed using a particle-based method
which yields relevant information about wave generation and fluid forces. The goal is
to find the optimal navigation technique when the submarine is operating on the free-
surface and deeply submerged for different values of sinkage and depths in two of
Titan’s largest seas, Kraken and Ligeia Mare.

I. Introduction
ECENT findings made in the Cassini-Huygens mission have shown that Titan, Saturn’s largest moon, is the
R only known planetary object in our solar system, other than Earth, that possesses clear evidence of stable,
accessible bodies of surface liquid, depicted in Fig. 1. Data from this mission indicate that the Titan polar regions
sustains stable seas of variable concentrations of ethane, methane, and nitrogen, with a surface temperature around 93
K[1]. Since the finding of liquid hydrocarbon seas on Titan’s surface, NASA has supported research towards
developing the capabilities of conducting in-situ science within these liquid bodies. A conceptual design of an
autonomous submarine, depicted in Fig. 2, to navigate Titan’s cryogenic seas was funded by NASA’s Innovative
Advanced Concepts (NIAC) Phase I in 2014[2]. Concepts of vehicles capable of landing in these cryogenic seas have
been explored in the past as part of a NASA–ESA Flagship mission, Titan Saturn System Mission (TSSM), and a
NASA Discovery solicitation study, Titan Mare Explorer (TiME)[3]. The Huygens probe from the Cassini-Huygens
mission was also designed with the ability to float in the event that the landing was to occur in a liquid rather than on
land, a possibility still unknown during the design of the probe[4,5]. Floating probes would be capable of conducting
various free-surface experiments, such as measuring sea-surface meteorology, observing larger-scale weather activity
and investigating tidal waves. Conversely, the advantage of conducting science with a submarine versus the other
aforementioned vehicles from previous studies is that the former would be capable of submerging to various depths
whereas the latter would be limited to remaining on the surface. Conducting submerged science has the potential for
a broader range of results from the mission including mapping using side-looking sonar, imaging and spectroscopy of
the sea at all depths, as well as sampling of the sea bottom and shallow shoreline[2]. The submarine, however, would
be required to surface in order to communicate with Earth. Therefore, a mission for a Titan submarine requires both
surfaced and deeply submerged operations.
1
Graduate Student, Mechanical and Aerospace Engineering Department, 6 MetroTech Center, Brooklyn, NY 11201,
Student Member.
2
Professor, Mechanical and Aerospace Engineering Department, 6 MetroTech Center, Brooklyn, NY 11201.
3
Senior Research Aerospace Engineer, Fluids and Cryogenics Branch, 21000 Brookpark Road, MS 86-12, Cleveland,
OH 44135, Senior Member.
4
COMPASS Lead, Mission Analysis and Architecture Branch, 21000 Brookpark Road, MS 162-2, Cleveland, OH
44135, Senior Member.

1
American Institue of Aeronautics and Astronautics
Copyright © 2017 by the American Institute of Aeronautics and Astronautics, Inc.
All rights reserved.
Downloaded by NEW YORK UNIVERSITY on September 18, 2017 | http://arc.aiaa.org | DOI: 10.2514/6.2017-5204

Figure 1. Cassini Captures Sunlight Glistening off of the “Mirror-Like” Seas of Titan[3].

(a) (b) (c)


Figure 2. (a) Front-, (b) Side- and (c) Back-Views of the Titan submarine[2].

One of the preferred methods for analyzing the interactions between a submarine and its environment is to simulate
submarine operations using numerical techniques. Since multiple bodies of liquid were found on the surface of Titan,
each containing varying quantities of ethane, methane and nitrogen, Computational Fluid Dynamics, CFD, software
is indeed a viable option to analyze the flow generated by the vehicle as it traverses seas with properties that change
with location and depth. Thus, CFD tools were used that can account for these different values correctly while still
producing accurate results. Simulation results can then be used to assess which body of liquid is more suitable to
conduct a more efficient mission meant to conduct in-situ science.
Past cryogenics CFD research includes simulating and quantifying “sloshing” of cryogenic liquid enclosed in the
tanks of space launchers to predict the temperature and pressure evolution. Other applications of this research have
been used to model pressurization within tanks, two-phase chilldown and transfer of fluids, evaporation and
condensation, as well as long-term storage[6-9]. To the best of the authors’ knowledge, there has never been an attempt
to simulate hydrodynamic interactions between a cryogenic liquid and a vehicle. Research on CFD analyses of
submarines typically includes steady and unsteady Reynolds–Averaged Navier–Stokes, RANS, simulations[10].
Therein, a mesh is generated to simulate motion and to extract the resultant hydrodynamic forces and moments acting
on the submarine. In this research, knowledge of cryogenic liquid dynamics and submarine hydrodynamics are used
to implement CFD simulations with the goal of assisting the design of an autonomous submarine capable of traversing
and conducting in-situ science in the cryogenic, liquid hydrocarbon seas on the surface of Titan. Results from this
work will be provided as feedback to list navigation techniques and ideal locations to conduct science in order to
increase the efficiency of the overall mission.

2
American Institue of Aeronautics and Astronautics
II. Methodology
The numerical study of the flow field around the Phase 1 design of the Titan submarine includes fully submerged
and surfaced operating conditions.
Existing empirical results from the Phase 1 Report used sea property values analogous to pure ethane, depicted in
Table 1, at a number of different velocities assuming only deeply submerged conditions. Furthermore, based on total
system efficiency, 51%, and power available for surfaced and deeply submerged operations, 100 W and 440 W,
respectively, maximum free-surface and deeply submerged velocities were quantified and are stated in Table 2[2].
Table 3 represents a component-by-component breakdown of individual drag results by the empirical calculations and
the percentage they contribute to the total drag acting on the submarine moving at a velocity = 1 m/s.
Numerical simulations for deeply submerged conditions using ethane properties were conducted to compare
against existing results. However, rather than analyzing the flow over isolated components, this study was able to
simulate flow over the whole submarine then disseminate the individual drag generated by the corresponding
components while still producing a qualitative agreement to the total drag. Simulation results were then attained
analyzing similar operating conditions in Titan’s seas, Kraken and Ligeia Mare, to determine the accuracy of those
Downloaded by NEW YORK UNIVERSITY on September 18, 2017 | http://arc.aiaa.org | DOI: 10.2514/6.2017-5204

power contraints. Data from Cassini has shown that Kraken and Ligeia Mare have reference densities of 643.56 kg/m3
and 523.10 kg/m3 and dynamic viscosities of 1014.8 µPa-s and 227.88 µPa-s, respectively[11]. This discrepancy in
values compared to the literature will differentiate results but a power assessment can still be made.
Within the surfaced configuration simulations, there was a comparative analysis done between two different drafts,
which is measured as the distance between the bottom of the submarine and the free-surface of the liquid. These
differing values determined which navigation technique resulted in less drag from the liquid on the submarine hull:
(1) 0.6 meters, where the free-surface aligned with the centerline of the ballast tanks and (2) 0.9 meters, where the
submarine is submerged to a depth so that only the communication array and mast is exposed to the atmosphere. The
more efficient of the two configurations will help determine which is the ideal navigation technique for the submarine
while operating above the free-surface. Futhermore, the Phase 1 Report did not take free-surface phenomena, such as
wave forces, into account when determining the speed limit on the surface. Thus, succeeding results will determine
the adequacy of ignoring such forces.

Table 1. Property values used in Empirical


Models and Simulations[2].

Table 2. Drag values, EHP and Power


requirements based on empirical results[2].

3
American Institue of Aeronautics and Astronautics
Table 3. Results from Empirical Models[2].
Downloaded by NEW YORK UNIVERSITY on September 18, 2017 | http://arc.aiaa.org | DOI: 10.2514/6.2017-5204

A. Deeply Submerged Case


For the fully submerged configuration, the lack of the free-surface facilitates the use of a grid based commercial
software, such as ANSYS© Fluent©[12]. A parasolid model of the geometry of submarine components is imported to
ANSYS and a domain of fluid surrounding the component is generated. The geometry model is then subtracted from
the enclosure, leaving only the fluid and the outer-shell of the submarine part inside. The mesh of the fluid domain is
obtained utilizing the ANSYS native mesh module. Simulations of coarse to high resolutions are performed. Fluid
flow and pressure forces on the hydrodynamic skin are subsequently presented in this section. For rapid computational
implementation within the framework of Fluent, and to increase the number of meshes at areas of interest, a solution-
based adaptive mesh is used to further refine the mesh when and where needed. Convergence is then determined
twofold: to determine the validity of a simulation using one mesh, and again while varying mesh sizes to ensure mesh
independence.
A finite volume method with a standard k– turbulence model and standard wall treatment function is chosen. The
semi-implicit method for pressure-linked equations (SIMPLE) scheme and second-order upwind scheme are used to
solve the RANS equations in Cartesian tensor form:

𝜕𝜌 𝜕
+ (𝜌𝑢𝑖 ) = 0 (1)
𝜕𝑡 𝜕𝑥𝑖

𝜕 𝜕 𝜕𝑝 𝜕 𝜕𝑢𝑖 𝜕𝑢𝑗 2 𝜕𝑢𝑙 𝜕


(𝜌𝑢𝑖 ) + (𝜌𝑢𝑖 𝑢𝑗 ) = − + [𝜇 ( + − 𝛿𝑖𝑗 )] + ̅̅̅̅̅̅
(−𝜌𝑢 ′ ′
𝑖 𝑢𝑗 ) (2)
𝜕𝑡 𝜕𝑥𝑗 𝜕𝑥𝑖 𝜕𝑥𝑗 𝜕𝑥𝑗 𝜕𝑥𝑖 3 𝜕𝑢𝑙 𝜕𝑥𝑗

where 𝑢𝑖 = 𝑢̅𝑖 + 𝑢𝑖′, 𝑢̅𝑖 and 𝑢𝑖′ are the mean and fluctuating velocity components, respectively, 𝜌 is the fluid density,
𝜇 is the fluid dynamic viscosity, p is pressure, t is time, and 𝛿𝑖𝑗 is the Kronecker delta. The stress tensor is estimated
using the standard k mode.
Boundary conditions at the domain edges include a 1 m/s velocity in the +x-direction at the inlet and isobaric
conditions at the outlet. Moreover, a no-slip condition is applied to all six faces of the submarine and a free slip wall

4
American Institue of Aeronautics and Astronautics
condition is applied to the four remaining outer faces, not including the inlet and outlet of this 6-sided prism produced
within Fluent. Future investigations will include varying the velocity in the +x-direction at the inlet and quantifying
the resultant differences.

B. Surfaced Case
The simulation of the flow generated by the submarine in the vicinity of the free-surface is carried out using
Smoothed Particle Hydrodynamics, SPH, a purely Lagrangian method, well-established in the literature for use in
hydrodynamics problems, via the computational tool DualSPHysics[1317]. The computational domain is regarded as a
set of moving particles, each having individual material properties and complying with the continuity and momentum
equations, written in SPH notation as follows:
𝑁
𝐷𝜌𝑘
= ∑ 𝑚𝑙 𝒖𝑘,𝑙 ∗ ∇𝑘 𝑊𝑘,𝑙 (3)
𝐷𝑡
𝑙=1
Downloaded by NEW YORK UNIVERSITY on September 18, 2017 | http://arc.aiaa.org | DOI: 10.2514/6.2017-5204

𝑁
𝐷𝒖𝑘 𝑝𝑙 𝑝𝑘
= ∑ 𝑚𝑙 ( 2 + 2 + 𝑇𝑘,𝑙 ) ∗ ∇𝑘 𝑊𝑘,𝑙 + 𝒈 (4)
𝐷𝑡 𝜌𝑙 𝜌𝑘
𝑙=1

where 𝒖𝑘,𝑙 = 𝒖𝑘 − 𝒖𝒍 , uk and ul represent the velocity vector of two random particles in the spatial domain, N is the
total number of particles within the support domain of particle k, ml, and 𝜌𝑙 are the mass and density of the interpolating
particle l, Tk,l represents the viscous stresses, and 𝑊𝑘,𝑙 = 𝑊(𝑥𝑘 − 𝑥𝑙 , ℎ) is a smoothing function that defines the
interpolation domain around each SPH particle.
SPH simulations are carried out using the open source code DualSPHysics v4.0[17]. Instead of focusing on a purely
incompressible problem, DualSPHysics utilizes the assumption of a small artificial compressibility for the fluid under
investigation, relating the pressure and the density via an equation of state given by:

𝜌 𝛾
𝑝 = 𝐵 (( ) − 1) (5)
𝜌0

where 𝜌0 is the fluid reference density, 𝛾 is a coefficient that depends on the fluid phase, typically equal to 7 for
liquids, and B is a coefficient that controls the relative density fluctuations. The value of B is chosen by scaling the
density fluctuations with the square of the Mach number of the flow and imposing a proper value of the particles’
speed of sound to obtain reasonable computational time[18].
The fluid properties chosen in this preliminary study are those of Kraken and Ligeia Mare. The gravitational
acceleration vector g  (0i, 0j, -1.352k) is used based on data for Titan available in the literature[2]. The initial particle
spacing is set to 1/160th of the submarine’s overall length, L, ~6 meters. This value represents a good compromise
between accurate solutions and reasonable computational time. The fluid domain has dimensions of 7  L × 3.5  L ×
(2/3)  L, and the physical simulation time is 25 s. The simulations consists of the submarine accelerating at a constant
rate of U/4 m/s2 for 4 s until reaching and maintaining a constant velocity, U, of 1 m/s. The total number of SPH
particles, including boundary particles, is about 6x107.

III. Results and Discussion


The results of several numerical simulations are presented in this section and are organized as follows: first,
pressure and velocity contours around the submarine and individual components are shown in the deeply submerged
condition coupled with a drag breakdown compared to the percentages in Table 2; next, drag values and corresponding
power requirements are determined with a submarine velocity = 1 m/s and compared to the values in Table 2; then,
velocity and free-surface elevation contours are illustrated in surfaced conditions for the submarine with varying
drafts; finally, for the same cruising conditions, drag forces and power requirements are attained and compared to the
values in Table 2.

A. Deeply Submerged Case


Figure 3 depicts pressure contours around the submarine and individual components, including the mast assembly
which is comprised of the surface imager, meteorology sensor and omni-antenna, and an individual stabilizer and
thruster. As expected, values of the stagnation pressure, expressed in gage units, converge to the value of the dynamic

5
American Institue of Aeronautics and Astronautics
pressure of the undisturbed flow. Areas of low pressure can be observed particularly behind the meteorology sensor
and other areas of accelerated flow, such as behind the stagnation point at the submarine bow. Observing the velocity
contours in Fig. 4, flow velocity gradients occur as expected from external components. Using the k-ℇ model for
tubulence formulation, no significant eddys are seen. Vorticity trails are seen following the top of the camera mast
and rear fins. At 1.5 m/s viscous forces account for 30% of the forces on the surface of the submarine.
Downloaded by NEW YORK UNIVERSITY on September 18, 2017 | http://arc.aiaa.org | DOI: 10.2514/6.2017-5204

(b)

(a)

(c)
Figure 3. Pressure Contours on (a) the Phase 1 Submarine, (b) an individual thruster and stabilizer (c) and
mast assembly.

Figure 4. Velocity contours on thePhase 1 Submarine moving at v = 1.5 m/s.

6
American Institue of Aeronautics and Astronautics
The drag force results calculated for the individual components are listed in Table 4. Overall, a good agreement is
seen between simulations and hand calculations. Some higher discrepancies are observed for certain components,
including the side-scan sonar array, duct, fins and camera housing. They are likely due to the interference of flow
generated by the submarine hull, which is not taken into account for drag calculation in Table 2.

Table 4. Component drag breakdown using pure ethane properties at v = 1 m/s.


% Difference
Phase 1 Report Fluent
between Phase 1
Component
Drag Drag Report and
% of total % of total Fluent
[lbf] [lbf]
Bare Hull 3.28 23.50% 7.39 50.62% 125%
Fins 0.42 3.00% -0.13 -0.92% 132%
Duct 0.17 1.20% 2.63 18.04% 1450%
Downloaded by NEW YORK UNIVERSITY on September 18, 2017 | http://arc.aiaa.org | DOI: 10.2514/6.2017-5204

Antenna 0.95 6.80% 1.47 10.07% 55%


Struts 1.95 14.00% 0.40 2.71% 80%
EGVs 0.00 0.00% 0.00 0.00% 0%
Pods 1.72 12.30% 0.00 0.00% 0%
DS 0.79 5.70% 0.12 0.80% 85%
Bottom Sampler 1.84 13.20% 0.79 5.39% 57%
P3 Sensor 0.16 1.10% 0.02 0.12% 89%
Side scan sonar 0.10 0.70% 0.25 1.69% 147%
Sonar transducer 0.04 0.30% 0.04 0.25% 10%
Camera Housing 0.13 0.90% 0.88 6.03% 578%
Camera mast 0.17 1.20% 0.09 0.62% 47%
Met. Sensor 0.36 2.60% 0.19 1.29% 48%
Subsurface illumination 1.48 10.60% 0.26 1.79% 82%
Doppler Velocimeter 0.19 1.40% 0.16 1.08% 17%
Omni-antenna 0.21 1.50% 0.06 0.43% 70%
Total drag, lbf 13.97 100.00% 14.60 100.00% 5%

Table 5 lists the drag forces and power requirements corresponding to a velocity = 1 m/s in both Kraken and Ligeia.
Based on these results and the power available for deeply submerged operations, the submarine is capable of achieving
and maintaining such velocities, confirming the results in the Phase 1 report despite the difference in properties.

Table 5. Drag results and power requirements corresponding to a velocity = 1 m/s deeply submerged in
Kraken and Ligeia Mare.
Sea Drag [N] EHP [W] Power [W]

Kraken 51.8 51.8 101.6

Ligeia 37.8 37.8 74.1

B. Surfaced Case
A series of preliminary results is presented for DualSPHysics simulations of the submarine navigating the free-
surface in the two sinkage conditions with drafts = 0.6 and 0.9 meters. Figures 5a and 5b show the velocity contours
for both these conditions along the submarine centerplane for a cruising speed of 1 m/s. In both cases a significantly
long wake can be observed behind the submarine, with different features depending on the depth of submergence.
When only the bottom thrusters are submerged, the interaction between the fluid and the cowlings yields an oscillatory

7
American Institue of Aeronautics and Astronautics
flow pattern typical of flow past a solid object. This effect is less evident when all thrusters are submerged, with a
possible cause being the interference between the two patterns from the top and bottom thrusters. Furthermore, with
a draft = 0.6 meters it is possible to observe a slight ventilation of the rudder, although small enough to be considered
negligible in terms of hydrodynamic pressure change at the stern. Clearly this effect does not occur in the larger draft
since the rudder is fully submerged.
Downloaded by NEW YORK UNIVERSITY on September 18, 2017 | http://arc.aiaa.org | DOI: 10.2514/6.2017-5204

(a)

(b)
Figure 5. Velocity contours obtained via DualSPHysics at the submarine center plane for a velocity = 1 m/s
and (a) draft = 0.6 meters and (b) draft = 0.9 meters.

A top-view of the free-surface elevation in the vicinity of the submarine is shown in Fig. 6. A nondimensional
ratio between free-surface elevation, z, and the overall height of the submarine, H, is used to determine resultant wave
heights. Bow and stern waves propagates in a direction which is at an angle with respect to the path of the submarine.
The features of these waves are similar to those seen on Earth for regular vessels [15,19,20]. Larger values of the free-
surface deformation are observed with a smaller draft since a larger bow area is above the liquid interface, resulting
in a larger free-surface flow. In that condition the “deck” of the submarine is covered with a layer of liquid resulting
from run-off of the bow wave. Due to the relatively low speed of the submarine during navigation at the free-surface,
part of the liquid is able to rest on top of the submarine at this draft. This does not occur as the submarine submerges
since the hull is sunk past the depth of the “deck”, and only the communication array and mast are exposed to the
atmosphere. An interesting result can be seen in Fig. 6a as the waves generated at the stern do not have the typical
form of transverse waves of ships, i.e. normal to the direction of motion, but travel at an angle with it, as for the bow
waves, but with a significantly smaller amplitude. After the main wave, a group of secondary waves excited by the

8
American Institue of Aeronautics and Astronautics
submarine with a draft of 0.9 meters present a higher amplitude than with a draft of 0.6 meters. Further analyses
comparing the wake patterns resulting from both conditions will be a part of future investigations.
Downloaded by NEW YORK UNIVERSITY on September 18, 2017 | http://arc.aiaa.org | DOI: 10.2514/6.2017-5204

(a)

(b)
Figure 6. Side-view of free-surfaced elevation obtained with DualSPHysics for (a) a draft = 0.6 meters and (b)
a draft = 0.9 meters.

The total hydrodynamic force acting on the hull opposing motion in the forward direction as the submarine operates
in Kraken and Ligeia at drafts = 0.6 and 0.9 meters are plotted in Figures 7 and 8, respectively. Due to Kraken’s higher
density and viscosity, the differences in overall forces acting on the submarine in the forward motion are equivalent
to the difference in those properties, ~20%. At a draft = 0.6 m, the total hydrodynamic drag convergences at around
the values of ~200 N and ~250 N after about 15 seconds of simulated physical time in Kraken and Ligeia, respectively.
Results from the same simulation with a submarine draft = 0.9 m show drag values converging to ~250 N and ~300
N over the same time interval. Therefore, when the submarine has a larger draft, more drag force acting on the
submarine ensues. This is a result of more of the submarine coming into contact with the liquid which results in
increaesd viscous drag. However, as the submarine submerges to larger drafts, the wave heights, and resulting wave
forces, decrease, illustrated by the bow waves in Figure 6. This relationship coupled with the results shown in Figures
7 and 8, determine the increase in viscous forces are greater than the change in wave forces corresponding to depth
since the larger draft attains more total drag.

9
American Institue of Aeronautics and Astronautics
Downloaded by NEW YORK UNIVERSITY on September 18, 2017 | http://arc.aiaa.org | DOI: 10.2514/6.2017-5204

Figure 7. Total drag acting on the submarine at a draft = 0.6 m in Kraken and Ligeia Mare.

Figure 8. Total drag acting on the submarine at a draft = 0.9 m in Kraken and Ligeia Mare.

Finally, Table 6 lists the power requirements for the submarine to move at a velocity = 1 m/s on the free-surface
at the two drafts used in this study. Based on the power limitation of 100 W when operating on the free-surface, the
submarine is not capable of reaching such speeds, confirming the Phase 1 Report analysis. However, values attained
from DualSPHysics are much larger than the empirical results at this velocity on the surface which stems from the
free-surface phenomena being taken into account. It is recommended to avoid maneuvering on the surface as much as
possible and only navigate the seas when operating deeply submerged. Therefore, surfaced operations should only
consist of communicating with Earth and performing observational experiments.

Table 6. Total drag, EHP and power requirements for the submarine operating at drafts = 0.6 and 0.9 m at a
velocity = 1 m/s in Kraken and Ligeia Mare.
Draft [m] Sea Total Drag [N] EHP [W] Power [W]
Kraken 238.54 238.54 467.72
0.6
Ligeia 194.29 194.29 380.96
Kraken 292.26 292.26 537.06
0.9
Ligeia 237.99 237.99 466.65

10
American Institue of Aeronautics and Astronautics
IV. Conclusion
Numerical simulations of flow past a submarine operating in the liquid hydrocarbon seas of Titan, Saturn’s largest
moon, are conducted. To simulate the submarine operating while deeply submerged, mesh-based code is used. Results
attained include pressure and velocity contours as well as quantifying total drag forces acting on the submarine in
Kraken and Ligeia Mare. These solutions indicate pressure and velocity features as expected. Quantified drag results
in this study are in good agreement with existing empirical results. The initial power requirements confirm such speeds
are achievable for this mission. Specific cases of the submarine navigating at the free-surface are analyzed using a
particle-based method. Relevant information is obtained about wave generation and fluid forces. Comparitive analyses
between drafts for free-surface navigation are made. Observations from the free-surface elevation contours of draft
demonstrate features of the Kelvin waves developing for both operation modes. Preliminary results show that free-
surface navigation presents less drag when less of the hull comes into contact with the free-surface despite increased
wave forces. Drag acting on the submarine operating on the surface is far greater compared to deeply submerged
operations when free-surface phenomena is taken into account.
Downloaded by NEW YORK UNIVERSITY on September 18, 2017 | http://arc.aiaa.org | DOI: 10.2514/6.2017-5204

References
J.W., Colozza, A., Lorenz, R.D., Oleson, S., Landis, G. Schmitz, P., Paul, M., Walsh, J., 2016. “Exploring the depths
1Hartwig,

of Kraken Mare – Power, thermal analysis, and ballast control for the Saturn Titan submarine,” Cryogenics, 74:31–46.
2Oleson, S.R., Lorenz. R.D., Paul, M.V., 2015, “Phase 1 Final Report: Titan Submarine,” NASA Technical Reports Server,

20150014581.
3Oleson, S.R., Lorenz. R.D., Paul, M.V., 2015, “Titan Submarine: Exploring the Depths of Kraken Mare,” Proceedings of the

AIAA SPACE 2015 Conference and Exposition, August 31–September 2, Pasadena, CA.
4Lorenz, R.D., Mitton, J., 2002, “Lifting Titan’s Veil: Exploring the Giant Moon of Saturn,” Cambridge UP, Cambridge.
5Lorenz, R.D., Mitton, J., 2008, “Titan Unveiled: Saturn’s Mysterious Moon Explored,” Princeton UP, Princeton.
6Lacapere, J., Vieille, B., Legrand, B., 2009, “Experimental and Numerical Result of Sloshing with Cryogenic Fluids,” Progress

in Propulsion Physics, 1:267–278.


7Hartwig, J., Vera, J., 2016 “Numerical Modeling of the Transient Chilldown of a Cryogenic Propellant Transfer Line,” Journal

of Thermophysics and Heat Transfer, 30(2):403–409.


8Kartuzova, O., Kassemi, M., 2011, “Modeling Interfacial Turbulent Heat Transfer during Ventless Pressurization of a Large
Scale Cryogenic Storage Tank in Microgravity,” Proceedings of the AIAA/ASME/SAE/ASEE 2011 Joint Propulsion Conference
& Exhibit, July 31–August 3, San Diego, CA.
9Kartuzova, O., Kassemi, M., 2016, “Modeling Droplet Heat and Mass Transfer during Spray Bar Pressure Control of the

Multipurpose Hydrogen Test Bed (MHTB) Tank in Normal Gravity,” Proceedings of the AIAA 2016 Propulsion and Energy
Forum, July 25–27, Salt Lake City, UT.
10Pan, Y., Zhang, H., Zhou, Q., 2012, “Numerical Prediction of Submarine Hydrodynamic Coefficients Using CFD

Simulation,” Journal of Hydrodynamics, 24(6):840–847.


11Hartwig, J.W., Meyerhofer, P., Lorenz, R., and Lemmon, E., 2017, “An Analytical Solubility Model for

Nitrogen/Ethane/Methane Ternary Mixtures,” Icarus.


12ANSYS© Academic Research, Release 17.1.
13Gingold, R.A, Monaghan, J.J., 1977. “Smoothed particle hydrodynamics – Theory and application to non-spherical stars,”

Monthly Notices of the Royal Astronomical Society, 181:375– 389.


14Tafuni,
A., Sahin, I., 2014. “Seafloor Pressure Signatures of a High-Speed Boat in Shallow Water with SPH,” Proceedings
of the ASME 2014 International Conference on Ocean, Offshore and Arctic Engineering, June 8–13, San Francisco, CA.
15Tafuni, A., Sahin, I., Hyman, M., 2015. “Numerical Investigation of Wave Elevation and Bottom Pressure Generated by a
Planing Hull in Finite–Depth Water” Applied Ocean Research, 58:281–291.

11
American Institue of Aeronautics and Astronautics
16Vacondio, R., Rogers, B.D., Stansby, P.K., Mignosa, P., Feldman, J., 2013. “Variable Resolution for SPH: A Dynamic

Particle Coalescing and Splitting Scheme,” Computer Methods in Applied Mechanics and Engineering, 256:132–148.
17Crespo A.J.C., Domínguez J.M., Rogers B.D., GómezGesteira M., Longshaw S., Canelas R., Vacondio R., Barreiro A.,

García-Feal O., 2015. “DualSPHysics: Open-Source Parallel CFD Solver on Smoothed Particle Hydrodynamics (SPH),” Computer
Physics Communications, 187:204–216.
18Tafuni, A., Sahin, I., 2015. “Non-linear Hydrodynamics of Thin Laminae Undergoing Large Harmonic Oscillations in a

Viscous Fluid,” Journal of Fluids and Structures, 52:101–117.


19Rabaud, M., Moisy, F., 2013. “Ship Wakes: Kelvin or Mach Angle?,” Physical Review Letters, 110:214503.
20Sahin, I., Hyman, M., 2001. “Simulation of ThreeDimensional Finite-Depth Wave Phenomenon for Moving Pressure

Distributions,” Ocean Engineering, 187:1621–1630.


Downloaded by NEW YORK UNIVERSITY on September 18, 2017 | http://arc.aiaa.org | DOI: 10.2514/6.2017-5204

12
American Institue of Aeronautics and Astronautics

View publication stats

You might also like