• 검색 결과가 없습니다.

Maneuvering simulation of an X-plane submarine using computational fluid dynamics

N/A
N/A
Protected

Academic year: 2021

Share "Maneuvering simulation of an X-plane submarine using computational fluid dynamics"

Copied!
13
0
0

로드 중.... (전체 텍스트 보기)

전체 글

(1)

Maneuvering simulation of an X-plane submarine using

computational

fluid dynamics

Yong Jae Cho

a

, Woochan Seok

b,*

, Ki-Hyeon Cheon

a,1

, Shin Hyung Rhee

a,b,** aDepartment of Naval Architecture and Ocean Engineering, Seoul National University, Seoul, South Korea

bResearch Institute of Marine Systems Engineering, Seoul National University, Seoul, South Korea

a r t i c l e i n f o

Article history: Received 16 April 2020 Received in revised form 30 August 2020 Accepted 6 October 2020

Keywords:

Computationalfluid dynamics X-plane submarine

Planar motion mechanism Captive model tests Maneuvering simulations

a b s t r a c t

X-plane submarines show better maneuverability as they have much longer span of control plane than that of cross plane submarines. In this study, captive model tests were conducted to evaluate the maneuverability of an X-plane submarine using Computational Fluid Dynamics (CFD) and a mathe-matical maneuvering model. For CFD analysis, SNUFOAM, CFD software specialized in naval hydrody-namics based on the open-source toolkit, OpenFOAM, was applied. A generic submarine Joubert BB2 was selected as a test model, which was modified by Maritime Research Institute Netherlands (MARIN). Captive model tests including propeller open water, resistance, self-propulsion, static drift, horizontal planar motion mechanism and vertical planar motion mechanism tests were carried out to obtain maneuvering coefficients of the submarine. Maneuvering simulations for turning circle tests were per-formed using the maneuvering coefficients obtained from the captive model tests. The simulated tra-jectory showed good agreement with that of free running model tests. From the results, it was proved that CFD simulations can be applicable to obtain reliable maneuvering coefficients for X-plane submarines.

© 2020 Production and hosting by Elsevier B.V. on behalf of Society of Naval Architects of Korea. This is an open access article under the CC BY-NC-ND license (http://creativecommons.org/licenses/by-nc-nd/4.0/).

1. Introduction

Submarines are operated both underwater and near the free surface to accomplish their missions. Though not frequent, there are accident reports such as collision and grounding when sub-marines move in a narrow waterway or canal. In addition, asym-metric force and moment arisen from the asymasym-metric flow adversely affect the submarine’s maneuverability. Hence, the suf-ficient maneuverability of submarines is required to prevent such kinds of accidents.

There are two kinds of submarines from the point of maneu-verability design: cross-plane submarines and X-plane submarines. A cross-plane submarine has two types of rudder blades depending on their function and an X-plane submarine has four rudder blades, as shown inFig. 1. The horizontal and vertical rudders of the

cross-plane submarine function as the vertical control system and hori-zontal control system, respectively. However, in the X-plane sub-marine, entire rudders are moved when the submarine is operated. The former has the orthogonal configuration of vertical and horizontal rudders, which is convenient for direct control of sub-marine’s positioning and widely adopted to traditional submarines. In cross-pane submarines, asymmetrical loads on rudders are observed, because the sizes of upper and lower rudders are different. The latter consists of the four rudders of the same size on the stern and each rudder can be independently operated unlike those for cross-plane submarines. As a result, even if one rudder fails to function, maneuverability can be secured by operating the others. In addition, it is well known that a submarine with X-plane rudders of higher aspect ratio demonstrates better maneuverability than one with the cross-plane rudders (Renilson, 2015). However, most studies on submarine maneuverability have been done for cross-plane submarines and those for X-plane submarines are hardly found in spite of their advantages in the viewpoint of maneuverability.Dubbioso et al. (2017) investigated the turning abilities of a submarine with two different rudder configurations using CFD. The results were allowed to analyze the maneuvering prediction of the submarine and the turning ability of the X-plane

* Corresponding author.

** Corresponding author. Department of Naval Architecture and Ocean Engineer-ing, Seoul National University, Seoul, South Korea.

E-mail addresses:swc@snu.ac.kr(W. Seok),shr@snu.ac.kr(S.H. Rhee). Peer review under responsibility of Society of Naval Architects of Korea. 1Currently, Defense Agency for Technology and Quality, Jin-ju, South Korea.

Contents lists available atScienceDirect

International Journal of Naval Architecture and Ocean Engineering

j o u r n a l h o m e p a g e : h t t p : / / w w w . j o u r n a l s . e l s e v i e r . c o m / i n t e r n a t i o n a l - j o u r n a l - o f - n a v a l - a r c h i t e c t u r e - a n d - o c e a n - e n g i n e e r i n g /

https://doi.org/10.1016/j.ijnaoe.2020.10.001

2092-6782/© 2020 Production and hosting by Elsevier B.V. on behalf of Society of Naval Architects of Korea. This is an open access article under the CC BY-NC-ND license (http://creativecommons.org/licenses/by-nc-nd/4.0/).

(2)

rudder configuration was superior to the cross-plane rudder configuration.Feng et al. (2015)studied the turning abilities of a submarine with cross-plane and X-plane rudders through numer-ical simulations. The X-plane rudder showed better turning ability than the cross-plane rudder due to a large effective rudder area.

Traditionally, there are two methods to evaluate a submarine’s maneuverability: free running model tests and simulation based on mathematical maneuvering models. Free running model tests are conducted by controlling the movement of submarine model with a remote controller in a large basin or lake. A series of free running tests are conducted to investigate the dynamic characteristics of a submarine and to validate its various functions. Mathematical maneuvering models are based on the Newton’s second law to describe the motion relative to the condition of the rudders and propeller.

The maneuverability of submarines in various situations has been evaluated from the simulations conducted by using mathe-matical maneuvering models. Gertler (1967) first proposed the mathematical maneuvering model for submarines, a set of equa-tions. The equations consist of linear and nonlinear coefficients, which can be obtained from the captive model tests. Feldman (1995)introduced the revised mathematical maneuvering model to account for unsteady and nonlinear motion characteristics by adopting a crossflow drag model.

Generally, captive model tests are adopted for estimating the maneuvering coefficient as well as evaluating maneuverability of submarines. Pan et al. (2012)simulated the captive model tests numerically using the Reynolds-averaged Navier-Stokes (RANS). They well-predicted theflow around a submarine and forces and moments in Planar Motion Mechanism (PMM) tests and an acceptable level of accuracy for the results was shown. Captive model tests were conducted byPark et al. (2017)in a towing tank and a wind tunnel to determine the maneuvering coefficients of a submarine. The maneuvering coefficients were analyzed using the least square method and Fourier integral method. Thus obtained maneuvering coefficients were useful for the maneuvering simu-lation of a submarine.

Captive model tests have been considered as the most accurate method to obtain maneuvering coefficients. However, captive model tests have many limitations stemming from struts- or sting mounted test model and complex measurement system. On the other hand, there is a method to obtain submarine’s maneuvering coefficients without performing experiments: Computational Fluid

Dynamics (CFD).

As computing power is being advanced rapidly, the number of studies using CFD to evaluate the maneuverability of a submarine has been increasing. Tyagi and Sen (2006) used the Reynolds-averaged Navier-Stokes (RANS) equations solver to determine the transverse hydrodynamic damping coefficients on two AUV ge-ometries. The results were compared with the available data on similar geometries and it was concluded that linear and nonlinear coefficients could be derived by the RANS solver. Phillips et al. (2007) showed that maneuvering coefficients of a submarine from RANS simulations with the k-ε model agreed with those from experiment. A full set of maneuvering coefficients could be found with relatively low computational cost at the initial design stage.

Nouri et al. (2016)simulated captive model tests, that is, resistance and Planar Motion Mechanism (PMM) tests for an AUV using a RANS solver with the realizable k-ε model. For numerical valida-tion, experimental tests for the same AUV in a water tunnel were carried out. It was shown that the RANS solver was utilizable to estimate the coefficients with reasonable accuracy comparing to the experimental data.Lee et al. (2009)showed that the maneu-vering coefficients calculated by a RANS solver were similar to those obtained from the Vertical Planar Motion Mechanism (VPMM) tests. Recently, advanced methods in CFD simulations, such as overset grid methodology and actuator disk model were applied to predict the maneuvering performance of ships or sub-marines.Martin et al. (2015)simulated maneuvers of a submarine model using a direct simulation of rudders and stern planes with a propeller. For the direct simulation, the overset grid methodology was used. The overset grid methodology uses a set of overlapping grids to discretize the domain and interpolation at appropriate points to couple the solution on the different grids. It is especially useful in cases involving large relative motion between geometries. Compared with the experimental data and CFD results with direct propeller simulation, comparable results were obtained for all maneuvers, in particular for motions and forces and moments.

Carrica et al. (2011)andCastro et al. (2011)used the overset grid methodology to simulate self-propulsion for surface ships. In addition, the self-propulsion results for forces, sinkage and trim were in good agreement with reported CFD results by other re-searchers.Jasak et al. (2019)andSezen et al. (2018)performed self-propulsion simulations using an actuator disk, which is the most efficient approach to model a propeller from the computational point of view. They confirmed that this approach well-predicted the

Fig. 1. Configurations of two types rudder.

(3)

self-propulsion performance of a ship.

In this study, captive model tests for an X-plane submarine were simulated using SNUFOAM. The submarine model was a generic submarine Joubert BB2 (hereafter BB2) model with X-plane rud-ders. Horizontal turning tests were performed with the maneu-vering coefficients obtained from CFD. The simulated results agreed well with the trajectory of the free running model tests presented by MARIN under the same conditions.

The paper is organized as follows. The test model for simulation is presented first, and followed by the numerical method and mathematical maneuvering model. The results of captive model tests and horizontal turning circle tests are described and dis-cussed. Finally, conclusions are provided.

2. Computational method

The target submarine studied in the present study is the generic submarine BB2 model which was introduced as an international benchmark for submarines. MARIN modified the original geometry for X-plane and sail plane rudders, as illustrated inFig. 2. The main particulars of BB2 are shown inTable 1.

The governing equations for theflow around the submarine are the continuity and the Reynolds-averaged Navier-Stokes (RANS) equations for incompressibleflow, expressed as

vui vxi¼ 0 (1)

r

vui vt þ

r

ujvuvxi j¼  vp vxiþ v vxj

m

vui vxj

r

ui0uj0 ! (2)

where uiand ui0represent the mean andfluctuating components of

thefluid velocity, respectively.

r

is the fluid density and p is the static pressure.

m

is the dynamic viscosity and

r

u0iuj0is the Reynolds

stress tensor. The RANS equations are not closed due to the Rey-nolds stress tensor. To close the equations, the k

u

Shear Stress Transport (SST) model was applied. The k

u

SST model is a combination of the original Wilcox k

u

model and the standard k ε model.

The Boussinesq approximation was adopted to model the Rey-nolds stress tensor in terms of the mean velocity gradients. The Reynolds stress tensor can be assumed as follows:



r

u0iuj0¼

m

t vuvxi jþ vuj vxi ! 2 3

rd

ijk (3)

where

m

trepresents the eddy viscosity and k is the turbulent kinetic

energy. The eddy viscosity is computed in the k

u

SST model as follows

m

r

k

u

(4)

where

u

is the specific turbulence frequency (or the turbulent dissipation rate).

The governing equations for turbulence quantities in the k

u

SST model are as follows

r

kÞ vt þ v

r

ujk vxj ¼ P 

b

*

ru

kþ v vxj " ð

m

þ

s

k

m

tÞ vk vxj # (5) vð

ru

Þ vt þ d

r

uj

u

 vxj ¼

g

vtP

bru

2þ v vxj " ð

m

þ

s

u

m

vxv

u

j # þ2ð1  F1Þ

rs

u 2

u

vxvkj v

u

vxj (6)

where P represents the production rate of turbulent kinetic energy by the meanflow. The variables are defined in the following way:

t

ijvuvxi j

t

ij¼

m

t  2Sij2 3 vuk vxk

d

ij  2 3

r

k

d

ij Sij¼ 1 2 vui vxjþ vuj vxi ! 4 ¼ F141þ ð1  F1Þ42 F1¼ tanh  arg41 arg1¼ min " max ffiffiffi k p

b

*

u

d; 500

n

d2

u

! ;4

rs

u2k CDkud2 # CDku¼ max 2

rs

u21

u

vxvkj v

u

vxj; 10 20 ! (7)

where

a

1¼ 0:31;

b

*¼ 0:09 and

s

u2¼ 0:856 are the same as that in

the present RANS model.

For temporal discretization, the Crank Nicolson scheme was applied with second-order accuracy. The second-order upwind scheme was used for spatial discretization. For pressure-velocity coupling, PIMPLE algorithm which is the merged semi-implicit method for pressure linked equations (SIMPLE) algorithm and pressure implicit with splitting of operator (PISO) algorithm, was adopted (Chen et al., 2014). The applied computational methods are summarized inTable 2.

The computational domain for the submarine is illustrated in

Fig. 3. The domain consisted of two regions. The inner region was defined as a sphere of which the radius was 1:5L from the sub-marine’s Center of Gravity (CG) and the outer region was the remaining region. The inner and outer regions were used to describe rotating and translating motions of the submarine in PMM simulations, respectively. In order to deal with the interface

(4)

between two regions, an arbitrary mesh interface was adopted. The inlet boundary was positioned at 3L upstream from submarine’s CG and outlet boundary was located 4:5L downstream. The Dirichlet boundary condition was applied to the inlet boundary and the zero gradient boundary condition was applied to the outlet boundary. Symmetry condition was applied to the side boundary 2:6Laway from midsection of submarine and no-slip boundary condition was applied to the submarine surface.

Two sets of orthogonal coordinate system were employed as

shown inFig. 4. The body-fixed coordinate system, o  xoyozo, was

adopted to measure external forces and moments of submarine. The origin of the body-fixed coordinate system was located at CG of submarine. The xo-axis and the yo-axis were positive pointing

up-stream and starboard, respectively. The zo-axis was determined

according to the right-hand rule. The earth-fixed coordinate sys-tem, O

xhz

, was applied for the global position of the submarine. The grid for BB2 was generated by means of trimmed mesh method in STAR-CCMþ as shown inFig. 5. The cells were more clustered around the submarine to precisely predict the velocity and pressure gradients in the boundary layer region.

Fig. 6represents a schematic diagram of PMM tests. To express the change in running attitude induced by PMM, translational motion was described by moving the whole domain and rotational motion was described by rotating the inner zone only.

The 6 Degree of Freedom (DOF) equations of motion for sub-marine with the body-fixed reference frame are shown in Eq.(8)to Eq.(13), based on that ofWatt (2007).

mð _u  vr þ wqÞ ¼ X (8) mð _v  wp þ urÞ ¼ Y (9) mð _w uq þ vpÞ ¼ Z (10) Ix_p þ  Iz Iy  qr Ixzð _r þ pqÞ þ Iyz  r2 q2þ I xyðpr  _qÞ ¼ K (11) Iy_q þ ðIx IzÞrp  Iyxð _p þ qrÞ þ Izx  p2 r2þ I yzðqp  _rÞ ¼ M (12) Table 1

Main particulars of BB2 submairne

Description Symbol Unit Full scale Model scale

Length L m 70.2 3.8260

Beam B m 9.6 0.5232

Draft to deck Dd m 10.6 0.5777

Draft to sail top Ds m 16.2 0.8829

Displacement D m3 4440 0.7012

Longitudinal center of gravity (from nose) XCG m 32.31 1.761

Vertical center of gravity (from center of pressure hull) ZCG m 0.0443 0.0024

Propeller diameter Dp 5 0.273

Scale ratio l e 18.348

Table 2

Summary of computational methods.

Parameter Values

Turbulentflow model RANS

Temporal discretization Crank Nicolson scheme Spatial discretization scheme Second order upwind Pressure velocity coupling algorithm PIMPLE

Turbulence model kuSST model

Fig. 3. Computational domain.

Fig. 4. Inertial and body-fixed coordinates. Fig. 5. Grid distribution around the submarine. 846

(5)

Iz_r þ  Iy Ix  pq Izyð _q þ rpÞ þ Ixy  q2 p2þ I zxðrq  _pÞ ¼ N (13)

where m means the mass of the submarine and Ix; Iy; Iz are the

moments of inertia with respect to each axis. xG; yG and zG are

submarine’s CG of submarine. The terms on the right hand side of Eq.(8)to Eq.(13)represent the external forces and moments acting on the submarine.

The forces and moments acting on the submarine can be decoupled to four components as follows:

F¼ ½X Y Z K M NT¼ F

HSþ FCþ FHþ Fprop (14)

where FH is the hull force related to the body shape of the

sub-marine, FHSis the restoring force related to the gravity and

buoy-ancy, Fprop and FCare the forces acting on the propeller and control

plane, respectively.

Forces and moments exerted on the submarine and its ap-pendages can be expressed as Eq.(15) to (41)

1) Hydrostatic forces

XHS¼  ðW  BÞsin

q

(15)

YHS¼ ðW  BÞcos

q

sin4 (16)

ZHS¼ ðW  BÞcos

q

cos4 (17)

KHS¼  yBBcos

q

cos4 þ zBBcos

q

sin4 (18)

MHS¼  xBBcos

q

cos4 þ zBBsin

q

sin4 (19)

NHS¼  xBBcos

q

sin4 þ yBBsin

q

(20)

2) Forces and moments acting on control plane

XC¼ Xdbdb

d

b2þ Xdrdr

d

r2þ Xdsds

d

s2 (21) YC¼ Ydr

d

r (22) ZC¼ Zdb

d

bþ Zds

d

s (23) KC¼ Kdr

d

r (24) MC¼ Mdb

d

bþ Mds

d

s (25) NC¼ Ndr

d

r (26) 3) Hull forces XH¼ 1 2

r

L 2X0 uuu2þ Xvvv2  (27) YH¼ 1 2

r

L 2Y v0uv þ Yvv0vjvjþ21

r

L3ðY_v0_vÞ (28) ZH¼ 1 2

r

L 2Z0 *u2þ Zw0uwþ Zww0 wjwj  (29) KH¼12

r

L3  Kv0uv þ K0 vvvjvj  þ1 2

r

L 4K _v0_v (30) MH¼12

r

L3  M*0u2þ M0 wuwþ Mww0 wjwj  (31) NH¼12

r

L3  N0vuv þ N0 vvvjvj  þ1 2

r

L 4N _v0_v (32) 4) Propeller force Xprop¼ KTðJÞ

r

n2d4 (33) Kprop¼ KQðJÞ

r

n2d5 (34)

5) Summation of forces and moments

(6)

X¼ XHSþ XCþ XXPþ X_u_u þ Xvvv2þ Xvvvvv4þ Xrrr2þ Xrprp þ Xwww2þ Xwjwjþ Xwwwww4þ Xqqq2 (35) XXP¼ Xpropþ

r

2L2 h Xuu0 u2i (36) Y¼ YHSþ YCþ Y_v_v þ Y_p_p þ Y_r_r þ Yvv þ Yvjvjvjvj þ Ypp þ Ypjpjpjpj þ Yrrþ Yrjrjrjrj þ Yvjrjvjrj þ Ywpwpþ Ypqpq þ Yqrqr (37) Z¼ ZHSþ ZCþ Zw_w_þ Z_q_q þ Z0þ Zvvv2þ Zrrr2þ Zww þ Zwww2þ Zwjwjwjwj þ Zwwwww4þ Zqqþ Zqqq2 þ Zqjqjqjqj þ Zvpvp þ Zwjqjwjqj þ Zprpr (38) Table 3

Verification test for virtual captive model tests.

Fine Medium Coarse p/RE

Total cell count 6,658,833 2,788,222 1,211,204 Grid refinement ratio, r 1.3204 1.3367 e

X-directional force [N] 19.4021 19.3951 19.3738 4.0816/19.418

ε 0.00036 0.0011 e

GCI 0.0001 0.0006 e

Fig. 7. Resistance at various towing speed.

Fig. 8. Thrust and torque coefficients of POW tests. Fig. 9. Towing force at various propeller revolution speed. 848

(7)

K¼ KHSþ KCþ K_v_v þ K_p_p þ K_r_r þ Kvv þ Kvjvjvjvj þ Kppþ Krr þ Krjrjrjrj þ Kvwvw þ Kwpwpþ Kwrwrþ Kvqvq þ Kqrqr þ Kpqpqþ Kprop (39) M¼ MHSþ MCþ Mw_w_þ M_q_q þ M0þ Mvvv2þ Mvvvvv4 þ Mrrr2þ Mwwþ Mwww2þ Mwjwjwjwj þ Mqqþ Mqqq2 þ Mqjqjqjqj þ Mvrvr þ Mvpvp þ Mjwjqjwjq þ Mrprp (40)

Fig. 10. Definition ofaandbangles.

Fig. 11. Forces and moments at various angles of attack.

(8)

N¼ NHSþ NCþ N_v_v þ N_p_p þ N_r_r þ Nvv þ Nvjvjvjvj þ Npp

þ Nrrþ Nrjrjrjrj þ Njvjrjvjr þ Nvqvq þ Nwpwpþ Npqpq

þ Nqrqr

(41)

where xB; yBand zBare submarine’s center of buoyancy and W and

B are gravity and buoyancy forces acting on submarine, respectively.

The X-plane rudders consist of four rudders moving indepen-dently. The effective rudder angle (

d

rÞ and sternplane angle (

d

sÞ can

be written as

Fig. 13. Instantaneous pressure contours and streamlines.

Fig. 14. Control surface efficiency test results for sail plane.

(9)

d

d

up port

d

down portþ

d

up starboard

d

down starboard

4 (42)

d

d

up port

þ

d

down port

d

up starboard

d

down starboard

4 (43)

A verification study was conducted using the Grid Convergence

Index (GCI) method to determine the appropriate cell count. The GCI method which was proposed byCelik and Karatekin (1997)was used. The total resistance of the submarine was selected as a key variable in the verification study. The results of the verification study are summarized inTable 3. The order of accuracy was esti-mated as

Fig. 15. Control surface efficiency test results X-plane: Vertical motion.

(10)

p¼ln h ð∅medium ∅coarseÞ . ∅fine ∅medium i lnðrÞ (44)

where∅coarse,∅medium and∅fineare solutions on the coarse,

me-dium and fine grids, respectively. The Richardson extrapolated value (RE) and the GCI were calculated as follows:

RE¼ ∅fineþ  ∅fine ∅medium  rp 1 (45) GCI¼rpjεj 1 (46)

whereε is obtained from Eq.(47).

Table 4

PMM test conditions.

Test Unit Condition

HPMM Pure sway Maximum sway amplitude m 0.475

Frequency Hz 0.88, 1.02, 1.25, 1.76

Pure yaw Maximum sway amplitude m 0.475

Maximum yaw amplitude  6.38, 9.02, 11.05, 12.76, 14.27

Frequency Hz 0.14, 0.20, 0.25, 0.28, 0.32

VPMM Pure heave Maximum sway amplitude m 0.475

Frequency Hz 0.72, 0.88, 1.02, 1.25, 1.44

Pure pitch Maximum sway amplitude m 0.475

Maximum pitch amplitude  6.38, 7.81, 9.02, 11.05, 12.76

Frequency Hz 0.14, 0.17, 0.20, 0.25, 0.28

Fig. 17. Y-directional force and Z-directional moment in pure sway tests.

Fig. 18. Y-directional force and Z-directional moment in pure yaw tests.

(11)

jεj ¼ ∅fine ∅medium

∅fine

(47)

The grid refinement ratio, r, is the effective mesh refinement ratio of  Nfine Nmedium 1 D ¼  Nmedium Ncoarse 1 D

, with the total cell count, N, and the

number of dimensions, D. The results show good grid convergence from the corresponding RE of less than 0.1%. The medium grid was chosen for the results presentation.

Firstly, the resistance tests were conducted at various towing speeds. The resistance values are shown inFig. 7(a). However, it is limited to perform the verification of SNUFOAM as there are no experimental results for BB2. Therefore, resistance tests were per-formed for the DARPA Suboff submarine and compared to the experimental results (Crook, 1990). The DARPA Suboff hull geom-etry is publicly available, which has been used as a representative of the benchmark model for CFD predictions since 1989. Extensive model tests were carried out, providing data for resistance, velocity and pressurefields. As shown inFig. 7(b), the resistance values agreed well with experimental results, where the discrepancy is less than 1%.

Before performing self-propulsion tests, Propeller Open Water (POW) tests were simulated for MARIN 7371R propeller. MARIN 7371R, a six-bladed stock propeller, was employed to compare with the experimental results (Pontarelli et al., 2017) as shown inFig. 8.

Fig. 19. Z-directional force and Y- directional moment in pure heave tests.

Fig. 20. Z-directional force and Y- directional moment in pure pitch tests.

Table 5

Comparison of maneuvering coefficients of BB2.

SNUFOAM MARIN (CFD) Yv0 0.081 0.080 Nv0 0.0151 0.0176 Yr0 0.0104 0.0057 Nr0 0.0078 0.0060 Y_v0 0.0265 e N_v0 0.00015609 e Y_r0 0.00040765 e N_r0 0.0011 e

(12)

The thrust coefficient 

KT¼rnT2D4



and torque coefficient (KQ ¼ T

rn2D5) were compared with those of experiment as a function of

advance ratio (J¼ V

nD), where ndenotes the rotation per second and

D denotes the diameter of the propeller, respectively. The present results showed good agreement with the experimental data within a maximum discrepancy of 3% for KTand 10KQ, respectively. The

averaged quantities were validated against the experimental data at various advance ratios.

Fig. 9shows the towing force in self-propulsion tests obtained from the CFD and experiments (Kim et al., 2018) for the given propeller revolution speed (rpm) at the design speed. The obtained self-propulsion point was in good agreement with the experi-mental results, with difference less than 2.4%. From the self-propulsion test results, the submarine’s wake fraction (w) and thrust deduction (t) were obtained as 0.3312 and 0.1666, respectively.

3. Results and discussion

Static drift tests consist of static angle of attack and drift tests. Here,

a

and

b

represents the angle of attack as shown inFig. 10. Static drift tests using angles ranging between10andþ10were

conducted at the design speed.

The obtained forces and moments in the static angle of attack, static drift, pure yaw, pure pitch tests were compared with those predicted byBettle (2014)using CFD, as shown inFigs. 11 and 12. Total average errors were less than 6% for forces and moments, which are in an acceptable range (Gao et al., 2018).

FromFigs. 11 and 12, discrepancies in forces and moments are larger when the

a

and

b

increase. To confirm what caused these discrepancies, the instantaneous non-dimensional pressure, Cp ¼

p=ð0:5 

r

 U2Þ, contours and streamlines in the static angle of

attack and drift tests are shown inFig. 13, respectively. At

a

and

b

¼ 3, stagnation points were observed at both nose and leading edge of coning tower and theflow was attached to the BB2 surface in the static angle of attack and drift tests. On the other hand, for static angle of attack test at

a

¼ 10

, theflow was separated from the leeward side of the hull where a relatively low pressure region was observed. Further downstream, theflow was fully separated and it formed the vortex pair. For static drift test at

b

¼ 10, theflow was separated from the upper surface of the BB2 with low pressure. This created the large swirling streamlines on the leeward side of the coning tower. For these reasons, there can be some asymmetric effect at

a

and

b

¼ 10

due to the separatedflow at the leeward side of coning tower.

The static rudder tests were conducted to obtain the forces and moments acting on the submarine body in the presence of a control surface deflection angle. The test results were used to evaluate effectiveness of the control system of the submarine, comprising the sail planes and X-planes. The sail plane tests were performed using rudder deflection angles ranging from 30 to 30at zero

hull incidence. Because a submarine is bilaterally symmetrical, the X-plane rudder tests were conducted using rudder deflection an-gles ranging from 0to 21at zero hull incidence.Figs. 14e16show the forces and moments obtained from the sail plane and X-plane rudder efficiency tests. The coefficients related to the control sur-faces, such as Xdbdb; Xdrdrand Ndrcould be determined by the control surface efficiency tests. The lift of the sail plane was smaller than that of the other control surfaces. No roll moment was observed in the X-plane rudder efficiency tests as shown inFig. 16(c) because the geometry of the rudder is symmetric.

PMM tests were performed to determine maneuvering co-efficients of the submarine associated with variations in sway/yaw frequency. The maximum sway amplitude, yaw amplitude and the

Fig. 21. Comparison of trajectories in turning circle test. Table 6

Comparison of turning circle test results.

Unit SNUFOAM Overpelt et al. (2015) Difference

Transfer M 5.735 4.706 5.56%

Tactical diameter M 11.202 10.712 4.42%

Fig. 22. Time series of the motion of turning circle tests: advance speed, sway velocity and yaw rate. 854

(13)

frequency are denoted by y0,

j

0and

u

respectively. The sway and

yaw motions induced by the PMM device are as follows:

y¼ y0sin

u

t

j

¼

j

0cos

u

t (48)

The sway velocity v; sway acceleration _v, yaw rate r and yaw angular acceleration r_ can be expressed as follows:

v¼ ðy0

u

Þcos

u

t_v ¼   y0

u

2  sin

u

t r¼  ð

j

0

u

Þsin

u

t r_¼ 

j

0

u

2  cos

u

t (49)

PMM tests include Horizontal Planar Motion Mechanism (HPMM) and VPMM tests. HPMM tests consist of pure sway, pure yaw tests. The derivatives related to sway and yaw acceleration were determined by the pure sway and pure yaw tests, respectively. VPMM tests consist of pure heave and pure pitch tests. The de-rivatives related to vertical motion can be determined by VPMM test results. The test conditions are given inTable 4.

The results of both HPMM and VPMM tests are shown in

Figs. 17e20. The forces (X, Y and Z) and moments (M and N) were non-dimensionalized using the submarine length (L), design speed (U), and the fluid density (

r

Þ. The results were following linear trend and the maneuvering coefficients were determined using the least square method on the gradient.

The estimated maneuvering coefficients using SNUFOAM are shown and compared with the MARIN’s CFD results (Toxopeus, 2017) inTable 5. They presented a similar tendency of maneu-vering coefficients. The comparisons of the estimated maneuvering coefficients and those of MARIN indicated that SNUFOAM yields acceptable results.

Turning circle tests using the maneuvering coefficients obtained from the computational captive model tests were carried out and trajectories are shown inFig. 21. The trajectory was compared with that of free running model tests (Overpelt et al., 2015) when the rudder deflection angle was 30as shown inFig. 21. The tactical and

steady turning diameters of the simulation were less than 5% compared to those of the free running model tests, as summarized inTable 6. The discrepancy is in an acceptable range (≪7.26%). (Kim et al., 2018). Fig. 22shows the time series of the motion in the turning circle test simulation. The advance speed, sway velocity and yaw rate from SNUFOAM were similar with that of free running model test, as shown inFig. 22(a)e(c), which proves the validity of SNUFOAM.

4. Conclusions

In the present work, captive model tests of the Joubert BB2 submarine model with X-plane rudders were conducted by CFD simulations and its maneuvering performance was evaluated by a mathematical maneuvering model. The simulation model tests including resistance, static, HPMM and VPMM tests to obtain maneuvering coefficients were conducted using SNUFOAM, an in-house CFD solver specialized in naval hydrodynamics based on the open-source OpenFOAM toolkit. From the comparison between the simulated trajectory results utilizing the numerically evaluated maneuvering coefficients and those obtained by the experimental free running tests, it is confirmed that the simulation results show good agreements with the experimental results within the differ-ence of 5%.

Declaration of competing interest

The authors declare that they have no known competing

financial interests or personal relationships that could have appeared to influence the work reported in this paper.

Acknowledgements

This research was supported by the National Research Founda-tion of Korea through a grant funded by the Korean government (NRF-2017K1A3A1A19071629, NRF-2020R1I1A2074369) and the Institute of Engineering Research at Seoul National University.

References

Bettle, M.C., 2014. Validating design methods for sizing submarine Tailfins. In: Proceedings of Warship 2014. Naval Submarines& UUV’s, Bath, UK. June 18-19. Carrica, P.M., Fu, H., Stern, F., 2011. Computations of self-propulsion free to sink and trim and of motions in head waves of the KRISO Container Ship (KCS) model. Appl. Ocean Res. 33 (4), 309e320.

Castro, A.M., Carrica, P.M., Stern, F., 2011. Full scale self-propulsion computations using discretized propeller for the KRISO container ship KCS. Comput. Fluids 51 (1), 35e47.

Celik, I., Karatekin, O., 1997. Numerical experiments on application of Richardson extrapolation with non-uniform grids. J. Fluid Eng. 119 (3), 584e590. Chen, L.F., Zang, J., Hillis, A.J., Morgan, G.C.J., Plummer, A.R., 2014.“Numerical

investigation of waveestructure interaction using OpenFOAM. Ocean Eng. 88, 91e109.

Crook, B., 1990. Resistance for DARPA Suboff as represented by model 5470. In: DTRC Report No. SHD-1298-07. David Taylor Research Center.

Dubbioso, G., Broglia, R., Zaghi, S., 2017. CFD analysis of turning abilities of a sub-marine model. Ocean Eng. 129, 459e479.

Feldman, J., 1995. Method of Performing Captive-Model Experiments to Predict the Stability and Control Characteristics of Submarines. CRDKNSWC Report No. HD-0393-25. Carderock Division Naval Surface Warfare Center.

Feng, D., Wang, X., Jiang, F., Zhang, Z., 2015. Large eddy simulation of DARPA SUBOFF for Re¼ 2.65 107. J. Coast Res. (73), 687e691.

Gao, T., Wang, Y., Pang, Y., Chen, Q., Tang, Y., 2018. A time-efficient CFD approach for hydrodynamic coefficient determination and model simplification of subma-rine. Ocean. Eng. 154, 16e26.

Gertler, M., 1967. The DTMB Planar-Motion-Mechanism System. Naval Ship Research and Development Center. NSRDE Report No. 2523.

Jasak, H., Vukcevic, V., Gatin, I., Lalovic, I., 2019. CFD validation and grid sensitivity studies of full scale ship self-propulsion. Int. J. Naval Architect. Ocean Eng. 11 (1), 33e43.

Kim, H., Ranmuthugala, D., Leong, Z.Q., Chin, C., 2018. Six-DOF simulations of an underwater vehicle undergoing straight line and steady turning manoeuvres. Ocean Eng. 150, 102e112.

Lee, S.K., Cheon, S.J., Joung, T.H., Jang, T.S., Lee, J.H., 2009. Evaluation of the added mass for a spheroid-type UUV by VPMM test. In: Proceedings of the Nineteenth International Offshore and Polar Engineering Conference, Osaka, Japan, June 21-26.

Martin, J.E., Michael, T., Carrica, 2015. Submarine maneuvers using direct overset simulation of appendages and propeller and coupled CFD/potentialflow pro-peller solver. J. Ship Res. 59 (1), 31e48.

Nouri, N.M., Mostafapour, K., Hassanpour, S.H., 2016. CFD modeling of wing and body of an AUV for estimation of hydrodynamic coefficients. J. Appl. Fluid Mech. 9 (6), 2717e2729.

Overpelt, B., Nienhuis, B., Anderson, B., 2015. Free running maneuvering model tests on a modern generic SSK class submarine (BB2). In: Proceedings of Pacific In-ternational Maritime Conference, Sydney, Australia, October 6-8.

Pan, Y.C., Zhang, H.X., Zhou, Q.D., 2012. Numerical prediction of submarine hydro-dynamic coefficients using CFD simulation. J. Hydrodyn. 24 (6), 840e847. Park, J.Y., Kim, N., Shin, Y.K., 2017. Experimental study on hydrodynamic coefficients

for high-incidence-angle maneuver of a submarine. Int. J. Naval Architect. Ocean Eng. 9 (1), 100e113.

Phillips, A.B., Furlong, M., Turnock, S.R., 2007. Virtual planar motion mechanism tests of the autonomous underwater vehicle autosub. In: Proceedings of STG-Conference, Hamburg, Germany, September 26.

Pontarelli, M., Martin, J.E., Carrica, P.M., 2017. Dynamic instabilities in propeller crashback. In: Proceeding of the 5th International Symposium on Marine Pro-pulsors, Espoo, Finland, June 12-15.

Renilson, M., 2015. Submarine Hydrodynamics. Springer, pp. 131e145.

Sezen, S., Dogrul, A., Delen, C., Bal, S., 2018. Investigation of self-propulsion of DARPA Suboff by RANS method. Ocean Eng. 150, 258e271.

Toxopeus, S.L., 2017. SHWG collaborative exercise: BB2. In: Tech. Report 29347-1-RD, Submarine Hydrodynamic Working Group.

Tyagi, A., Sen, D., 2006. Calculation of transverse hydrodynamic coefficients using computationalfluid dynamic approach. Ocean Eng. 33 (5e6), 798e809. Watt, G.D., 2007. Modelling and simulating unsteady six degrees-of-freedom

수치

Fig. 1. Configurations of two types rudder.
Fig. 3 . The domain consisted of two regions. The inner region was de fined as a sphere of which the radius was 1:5L from the  sub-marine ’s Center of Gravity (CG) and the outer region was the remaining region
Fig. 3. Computational domain.
Fig. 6. Schematic diagram of PMM tests.
+7

참조

관련 문서

~ depends on the relative importance of the turbulent transport terms For the turbulent jet motion, simulation of turbulence is important. For the horizontal motion in

The construction of Electric motor propulsion system are presumed arrangement of boat and compared The engine fuel of cost with battery charge of cost .The Model

Absolute and Relative Velocity in Plane Motion Instantaneous Center of Rotation in Plane Motion Absolute and Relative Acceleration in Plane Motion... Analysis of Plane

Hele-Shaw flow (potential flow) Van Dyke, Album of Fluid Motion (1982).. 2017

KEY WORDS: Computational fluid dynamics 전산유체역학, Rogue wave 로그파, Numerical wave tank 수치수조, OpenFOAM 오픈폼... 라이브러리

In the proposed method, the motion of focused object and the background is identified and then the motion vector information is extracted by using the 9

~ In steady flow, Lagrangian path lines are the same as the Eulerian streamlines, and both are the same as the streak lines, because the streamlines are then fixed in space

~ In steady flow, Lagrangian path lines are the same as the Eulerian streamlines, and both are the same as the streak lines, because the streamlines are then fixed in space