Vous êtes sur la page 1sur 12

10th International Workshop on Ship and Marine Hydrodynamics

Keelung, Taiwan, November 5-8, 2017

Resistance and Hull Form Optimization for Vietnamese Fishing Vessels


Tran Dinh Tu, Jiahn-Horng Chen*
National Taiwan Ocean University, Department of Systems Engineering and Naval Architecture
2 Pei-Ning Road, Keelung 20224, Taiwan

Tran Gia Thai


Nha Trang Univeristy, Department of Naval Architecture
2 Nguyen Dinh Chieu Street, Nha Trang City, Vietnam

*Corresponding author, B005@ntou.edu.tw

ABSTRACT

In the present study, the resistance of a traditional Vietnamese fishing vessel was studied numerically.
A finite volume method was used to discretize the governing equations for incompressible viscous
flow around the ship. The Reynolds-Averaged Navier-Stokes equations were solved with the Realizable
k- model and SST k- model. Moreover, the volume of fluid method was employed to capture the free
surface. The SIMPLE algorithm was chosen to iterate the velocity and pressure fields. The results showed that
wave-making resistance was about twice as much as the frictional resistance. And we found that Realizable k-
 model is more realizable and accurate.

1. INTRODUCTION

Many methods to predict the resistance of fishing vessels have been developed. The experimental test
is the most important and reliable approach for performance predictions of a ship. Unfortunately, conducting
towing tank tests can be difficult, expensive, and, sometimes, time-consuming. Moreover, because of the
impossibility to fulfil the dynamic similarity in model tests, it is necessary to use the Froude method or
Prohaska method to obtain the total resistance of the full-scale ship from its model test data.
An alternative to the experimental tests is the computational fluid dynamics (CFD) approach which finds
the ship flow numerically. Since it is typically turbulent, the use of Reynolds-Averaged Navier-Stokes (RANS)
solvers is common in prediction. The turbulent free-surface flow computations around a ship have been studied
by many studies [1-7]. They provide important information in the preliminary ship design. In fact, CFD has
played an important role in marine hydrodynamics. Progress of this field has been well benchmarked in CFD
workshops for resistance, propulsion and seakeeping [8-9]. Two recent reviews [10-11] also provide us general
perspectives of viscous ship flow prediction. Furthermore, there are two different approaches to capture the
free surface wave: interface tracking and interface capturing. The volume of fluid (VOF) method [1] and the
level-set method [2] which belong to the latter one have been widely employed for complex flows due to their
robustness and efficiency [3].
In this work, the VOF method implemented in a RANS solver, ANSYS Fluent 14.5, was employed to
compute the flow fields of two typical Vietnamese fishing vessels. The flow around model ship at several
Froude numbers was computed firstly to verify computational techniques. Then, these approaches utilize to
simulate the flow past the full-scale ship. The resistance, wave pattern and nominal wake at propeller plane
of model and full-scale ship were interpreted and discussed. These data, especially, resistance components as
well as the information of wave patterns, provide a specific insight to improve the hull form for reducing the
resistance.
2. COMPUTATIONAL METHOD

2.1 Governing equations

The flow around the ship is governed by the conservations of mass and momentum (RANS) which can
be expressed in Cartesian tensor form as follows as
 (ui )
0 (1)
xi

   ui  
  ui u j    p     u u j  
t

xi xi x j
   i 
  x j x j
    
  uiuj  gi (2)
  x j
where ui is the mean (time-averaged) velocity and where  and  are the fluid density and fluid
dynamic viscosity, respectively; p is the time-averaged pressure; ui is the fluctuating velocity component;
g i is the gravity. The effect of turbulence on the flow is represented in Equation (2) by the Reynolds stresses
tensor Rij    uiuj , which is computed using a turbulence model.

2.2 Ship geometry

Table 1 and Figure 1 show the principal parameters and hull forms of the two typical fishing vessels.
Pattern 1 is the traditional fishing boat on the central coastal area in Vietnam. Pattern 2 is typical and popular
in the same area as that of Pattern 1.

Table 1: Principal parameters of the ship


Pattern 1 Pattern 2
Ship Model Ship Model
LOA (m) 21.90 3.13 16.50 3.13
B (m) 4.48 0.64 4.80 0.91
D (m) 2.51 0.36 2.60 0.72
T (m) 1.23 0.18 1.70 0.32
CB 0.59 0.59 0.67 0.67
CM 0.89 0.89 0.93 0.93
W (T) 60.46 0.18 86.41 0.59
S (m2) 100.40 2.05 99.70 3.59

(a) Pattern 1 (b) Pattern 2.


Figure 1: The geometry and lines of two patterns.

2
2.3 Grid generation and boundary condition

In the generation of grids for the case of a ship advancing on the free surface, some particular tricks are
necessary to obtain satisfactory results in free-surface shape as well as in resistance predictions. In the present
study, the multi-block hexahedral grids were generated. Figures 2 and 3 show the grids for Patterns 1 and 2,
global and local arrangement respectively.

(a) Pattern 1(Modern style) (b) Pattern 2 (Traditional style)

Figure: 2 Global gird arrangement

(a) Pattern 1 (left: bow region; right: stern region)

(b) Pattern 2 (left: bow region; right: stern region)


Figure 3: Local grid arrangment
Figure 4 shows the computational domain and the specifications of boundary conditions, based on the
recommendations of ITTC 2011[6]. All the multi-block structured grids are hexahedral. The bottom is located
at 2LOA below the water surface, while inflow and outflow boundaries were posed respectively at LOA and 5LOA
from the hull. Since the geometry is symmetric, only half of the hull was considered. Using the VOF model,
the region occupied by air has to be modelled as well, thus the domain was extended for 0.5LOA above the
design waterline with an additional block.

3
2L L Hull (No-Slip wall)
L
Free surface

5L

Inlet (Velocity) 0.5 L

Symmetry
Outlet 2L
(Pressure)
Bottom (Slip wall)
Side (Slip wall)

Figure 4: Computational domain and boundary condition


The boundary conditions were chosen as follows. A uniform velocity was specified at the inlet both for
air and water. The zero gradient condition was set at the outlet for which the hydrostatic pressure for air and
water has to be specified at the outlet domain. Due to symmetry of the hull form, only half of the ship form
flow has been computed. Therefore, a symmetric condition was set on the symmetry plane. The side, top, and
bottom boundaries are relatively far from the hull and they were set as rigid-slip walls. Finally, a no-slip
condition wall was applied on the hull surface, and defined as the wall with the logarithmic law for the velocity
distribution, which is a usual setup for the wall function in fully developed turbulent boundary layer.
For the computed free-surface viscous flow around the ship, three multi-block structured grid are
generated: coarse grid of 3.8 million cells, medium grid of 6.7 million cells and fine grid of 8.3 million cells.
The extremely fine grid of 21 million cells has been generated to validate the results of full-scale ship. The
mesh sensitivity study was made to observe the impact of the mesh density on the towing resistance at
Fn=0.345. This study had shown that mesh density has the quite large effect on the obtained results. The total
resistances in Table 2 is clear indication that the simulation error decreases when the grid is more refined. The
relative difference between CFD method and experimental data in the fined grid is 0.18%.
Table 2: Grid independence study
(CT )comp  (CT )data
Grid size (CT)comp (CT)exp CT   100
(CT )data
3.8M 8.21  104 7.60  104 8.01 %
6.7M 7.81  104 7.60  104 2.77 %
8.3M 7.62  104 7.60  104 0.18 %

2.4 Solution method

In this study, the commercial code Fluent 14.5 was employed for all computations. This code is based
on the finite volume method. We employed SIMPLE algorithm [4] for steady state flow iterations. The pressure
staggering (PRESTO) scheme is used for pressure interpolation while other terms are discretized using a
second order upwind method. Volume fraction equations solved by high resolution interface capturing (HRIC)
scheme [5]. The two turbulence models, realizable k- and SST k- models, were employed to simulate the
turbulent flow past the hull.

4
3. RESULTS AND DISCUSSION

3.1 Ship resistance

As a ship moves through calm water, many factors combine to form the total resistance force acting on
the hull. The principle factors affecting ship resistance are the friction and viscous effects of water acting on
the hull, the energy required to create and maintain the ship’s characteristic bow and stern waves.
In mathematical terms, total resistance can be written
RT  RW  RF , (3)
where RT is the total resistance, RW is the wave-making resistance and RF is the frictional resistance.
We can also write total resistance in terms of dimensionless coefficients
CT  CW  CF , (4)

where CT is the total resistance coefficient, CW is the wave-making resistance coefficient and CF is the
frictional resistance coefficient. Since the total hull resistance is a function of hull form, ship speed, and water
properties, the coefficient of total hull resistance is also a function of hull form, ship speed, and water properties.

The coefficients are defined as follows.


RT
CT  (5)
0.5 SV 2
where S is the wetted surface area of the ship and V is ship speed. Similarly, the wake-making resistance and
frictional resistance coefficients are defined as
RW
CW  (6)
0.5 SV 2

RF
CF  (7)
0.5 SV 2
The total resistance coefficient is validated with available experimental data. The relative difference
between experimental data and computational results in percent was then determined by

(CT )comp  (CT )data


CT   100 (8)
(CT )data

The Froude numbers and Reynold numbers of flow field are defined as

V
Fn  (9)
gLWL

V LWL
Rn  (10)

where g is the gravitational acceleration and LWL is the length of water line.

The experimental data of total resistance of Pattern 1 is available [7]. Table 3 shows the results at 4
values of speed. It is obvious that they agree with each other quite well. The maximum errors is 2.88% at
Fn  0.202 while the minimum errors is 0.18% at Fn  0.345. This shows that the computational approach

5
can lead to reasonable accuracy. Furthermore, the total resistance coefficients at different Froude numbers are
shown in Figure 5.

Table 3: Total resistance and coefficient for Pattern 1.


VM (m/sec) Fn (RT)comp (N) (CT)comp (RT)exp (N) (CT)exp CT
1.758 0.345 31.960 1.0  10 2
31.902 1.0  102 0.18%
1.663 0.326 26.414 9.3  103 26.203 9.2  103 0.81%
1.544 0.303 20.800 8.5  10 3
20.464 8.4  103 1.64%
1.029 0.202 7.620 7.0  103 7.407 6.8  103 2.88%

1.2E-02

1.0E-02

8.0E-03

6.0E-03
CT

4.0E-03 CFD Experimental data

2.0E-03

0.0E+00
0.2 0.25 Fn 0.3 0.35
Figure 5: Comparison of the resistance coefficient.
After the computational method for model was verified, the resistance and flow for full-scale ship would
be conducted similarly. It can be carried out by two ways. The first one calculates the force and the flow field
directly for the same model by adjusting the water viscosity such that the Reynolds numbers for the model and
the ship are identical. It leads to the dynamic similarity between the model and the ship. The new value of
water viscosity is called artificial viscosity in the present study. With such a way, the geometry and dynamic
similarly between model and full-scale ship is fulfilled. It implies that the wave making resistance and frictional
resistance of model and ship are equal to those of the ship since the Reynolds number and Froude number of
model and ship satisfy simultaneously. This technique takes advantage of the CFD approach in ship resistance
prediction.
The second method is to compute the ship resistance directly by using the geometry of the real ship.
This technique leads to that the wake-making resistance of ship and model is the same but frictional resistance
is deferent due to the difference between Reynold number of model and full-size ship. The results of the two
methods could be compared to each other to verify the computational techniques.
Tables 4 and 5 shows the information of flow field such as Froude number, Reynold number as well as
the artificial and real viscosity of water in this study. The Reynoldd number of Pattern 1 model and full scale
ship are 4.658  106 and 8.627  107, respectively, while these values of Pattern 2 are 5.343  106 and 6.468
 107, respectively.

Table 4: Computational condition for Pattern 1


Turbulence
Grid Fn (Rn)Model (Rn)Ship Viscosity
model
8.3M Realizable k- 0.345 4.658  106 8.627  107 1.000  103
Model
8.3M SST k- 0.345 4.658  106 8.627  107 1.000  103
Artificial viscosity 8.3M Realizable k- 0.345 8.627  107 8.627  107 5.399  105
Real viscosity 8.3M Realizable k- 0.345 4.658  106 8.627  107 1.000  103
Ship
Real viscosity 21M SST k- 0.345 4.658  106 8.627  107 1.000  103
Real viscosity 21M Realizable k- 0.345 4.658  106 8.627  107 1.000  103

6
Table 5: Computational condition for Pattern 2
Turbulence
Grid Fn (Rn)Model (Rn)Ship Viscosity
model
7.2M Realizable k- 0.345 5.343  106 6.468  107 1.000  103
Model
7.2M SST k- 0.345 6.468  107 6.468  107 5.477  105
Artificial viscosity 7.2M Realizable k- 0.345 5.343  106 6.468  107 1.000  103
Ship
Real viscosity 7.2M Realizable k- 0.345 5.343  106 6.468  107 1.000  103

Table 6 presents the results of resistance components and their coefficient on model and full-scale ship
for the two patterns. It is obviously that the wave-making coefficients for model and full-scale ship are equal.
The result obtaining from finest grid (21M) use to estimates the errors of other cases. It is clear that the relative
errors of CW and CT are under 5%. This error can be reasonable since the simulation of model and full-scale
ship were performed on the same grid topology which is of different scale.

Table 6: Resistance components and its coefficient for model and ship
CW CT
RW (N) RF (N) RT (N) CW CF CT
(%) (%)
Pattern 1 Model 20.440 11.680 32.120 6.866  103 3.924  103 1.079  102 4.03 -
Artificial
Ship 21.055 7.084 28.139 7.073  103 2.380  103 9.453  103 1.13 2.81
viscosity
Real viscosity 7306.3 2414.9 9721.3 7.154  103 2.365  103 9.519  103 - -
Pattern 2 Model 54.408 14.686 69.094 1.29  102 3.490  103 1.640  102 4.88 -
Artificial
Ship 2.75 3.05
viscosity 50.432 8.600 59.032 1.20  102 2.050  103 1.40  102
Real viscosity 10002.5 1737.9 13267.5 1.23  102 2.140  103 1.450  102 - -

Figure 6 illustrates the resistance coefficient in the bar graph. It can be seen from the graph that the
wave-making resistance contributes a larger percentage in total resistance for both model and ship. Particularly,
for the full-scale ship of Pattern1, the wave-making resistance accounts for around 70% while Pattern 2 80%
in total resistance. It implies that reducing wake-making resistance makes significant contribution to improve
the hull form in optimization design.

2.00E-02
1.20E-02
1.00E-02 1.50E-02
8.00E-03 Cw Cw
1.00E-02
6.00E-03 Cf Cf
4.00E-03 Ct 5.00E-03 Ct
2.00E-03
0.00E+00
0.00E+00
Model Ship Ship
Model Ship Ship

(a) Pattern 1. (b) Pattern 2.


Figure 6: Resistance coefficient components.

Due to the differences of ratio LOA/B, B/T and CB, the total resistance of Pattern 2 is significantly higher
than that of Pattern 1. However, the total resistance per unit displacement for Pattern 2 is smaller than that of
Pattern 1. Generally, in the current loading conditions, the propulsive performance of Pattern 2 is more
effective that of Pattern 1.

Table 7: Total resistance of ship for two Patterns


Speed (knots) Fn (RT)P1 (N) (RT)P2 (N) (RT)P1/1 (N/Ton) (RT)P2/2 (N/Ton)
9.05 0.345 9721.3 - 160.8 -
7.80 0.345 - 13267.5 - 153.5

7
3.2 Wave pattern and the wake distribution

The wave pattern contours are shown in Figure 7 for both Patterns and two computational techniques
which are used to calculate resistance at Fn = 0.345. It is obvious that the typical Kelvin wave pattern [19],
including transverse and diverging waves, is clearly observed. The angles of diverging waves from stern found
by the two computational techniques are the same. It indicates that although we employed two techniques to
compute full-scale ship, we could obtain the same value of CW.
We compare the wave profiles on the hull of the two patterns. It is clear in Figure 8 that the wave crest
and trough on the hull of Pattern 2 is larger than those on the hull of the Pattern 1. It also shows that the bow
wave is twice as high as stern wave.

(a) Pattern 1 for model. (b) Pattern 2 for model.

(c) Pattern 1 for ship with artificial viscosity. (d) Pattern 2 for ship with artificial viscosity.

(e) Pattern 1 for ship with real viscosity. (f) Pattern 2 for ship with real viscosity.
Figure 7: Wave pattern contour and wave profile at Fn = 0.326.

Pattern 1
Pattern 2
Wave crest
Wave trough

Figure 8: Wave profile on the hull for two methods at Fn = 0.345.

8
Figure 9 shows the contours of the axial and tangential velocity vectors which present the nominal wake
distribution on the propeller plane. By comparing wake distribution of Pattern 1 to that of Pattern 2, it is
obviously clear that the stern geometry leads to a significant influence on the wake distribution. For Pattern
2, the velocity distribution on the propeller plane is obviously much more uniform, compared to that of Pattern
1. The tangential velocity vector also demonstrates a similar trend. The velocity uniformity is an important
factor for propeller operation because non-uniformity is the source of propeller noise and vibration. In addition,
a nominal wake is critical because it gives first indications on the propeller inflow.

(a) Pattern 1. (b) Pattern 2.


Figure 9: The axial velocity (left) and tangential velocity vector (right).
Figure 10 presents the streamlines of the flow around the stern for two Patterns. The streamline
formation shows that there are no recirculating regions at the stern so that the hull shape is reasonable widely
used in fishery in the central coastal area of Vietnam.

(a) Pattern 1. (b) Pattern 2.


Figure 10: The streamlines around the stern of model.
We also computed the turbulent flow around the ship by employing two turbulence models which were
widely and popular using in industry and hydrodynamics research community. There were Realizable k- and
SST k-. The results of 21M grid size in Table 8 indicate that the total resistance, which is computed by SST
k- is smaller Realizable k- and the deviation of CW is 0.3% while deviation of CF is 1.9%.

Table 8: Resistance components and its coefficient for Realizable k- and SST k-
Turbulence CW CF
RW (N) RF (N) RT (N) CW CF CT
model
Realizable k- 7328.9 2499.8 9828.6 7.176  103 2.448  103 9.624  103 - -
SST k- 7306.4 2414.9 9721.3 7.154  103 2.365  103 9.519  103 0.3% 1.9%

Due to the complexity of hull surface, we attempt to reduce the first layer of grid near wall as much as
possible to achieve that y+ is small enough. The turbulence models, which coupled frictional wall laws, have
restrictions on the value of y+. For instance, the k- model requires this value be between approximately 100
and 300. Figure 11 presents the value of y+ along the hull on waterline. At the midship location, this value is
around 50. Figures 12, 13 and 14 show the wave patterns, wave profile and wake distribution of two patterns.
These turbulent flow characteristics is quite similar. However, we found that the Realizable k- turbulence

9
model is considered more reliable and simple.

Figure 11: The y+ on the wall of 21M grid size of full-scale ship.

(a) Realizable k-, Pattern 1. (b) SST k-, Pattern 1.

(a) Realizable k-, Pattern 2. (b) SST k-, Pattern 2.

Figure 12: Wave pattern of full scale ship

(a) Pattern 1. (b) Pattern 2.

Figure 13: The wave profile on the hull of Pattern 1

10
(a) Realizable k-, Pattern 1. (b) SST k- Pattern 1.

(b) Realizable k-, Pattern 2. (b) SST k-, Pattern 2.


Figure 14: Nominal wake at plane x/LWL = 0.45m of Pattern 1.

4. CONCLUSIONS

The viscous free-surface flow around the Vietnam fishing boats has been carried out by computations
with RANS equations. The resistance and flow around the hull have been computed for model and full scale
ship. The realizable k- coupled with standard wall function turbulence model is efficient and quite accurate
when simulating the viscous flow past to the hull with free surface. The total resistance of experimental data
use to validate computational results of Pattern 1. The deviation was 0.18% when we employed a mesh with
8.3M computational cells at the Froude number 0.345. Since there exists no experimental data for the full scale
ship resistance, computational data on an extremely fined grid which has 21M computational cells are utilized
to validate the resistance for full size ship. The deviation was 3%, compared to the results computed with the
grid of 8.3M cells.
We compared the total resistance, wave pattern, wake distribution on the propeller plane, and
streamlines of two typical forms of fishing vessel. The stern shapes of two hull forms in this work are different.
Pattern 1 has a stern bulb while Pattern 2 has a straight one. The results show that the stern form of Pattern 2
induces a more homogeneous nominal wake. Furthermore, although the block coefficient of Pattern 2 (CB =
0.67) is bigger than that of Pattern 1 (CB = 0.59), the total resistance per unit displacement for Pattern 2 is
smaller. Therefore, its energy consumption per unit displacement is evidently smaller. To sum up, these results
show why the Pattern 2 is currently widely used in the fishery of Vietnam.
The results in this paper shows that the CFD approach has made possible significant capabilities and
advancement when solving the turbulent flow with free surface past complex geometry such as ship hull. In
addition, If there is no significant flow separation, the boundary layer is completely turbulent (artificial

11
turbulence generation by wires in towing tank tests), and for such case the realizable k- model is found to be
quite accurate and reliable.

REFERENCES

1. Pranzitelli, A., Nicola, C., and Miranda, S., (2011). Steady-state calculations of free surface flow around
ship hulls and resistance predictions. High Speed Marine Vehicles (IX HSMV), Naples, Italy.
2. Tahara, Y., Longo, J., and Stern, F.(2002). Comparison of CFD and EFD for the Series 60 CB = 0.6 in
steady drift motion. Journal of Marine Science and Technology, Vol. 7, No.1, 17-30.
3. Raven, H. C. and Starke, B.(2002). Efficient methods to compute steady ship viscous flow with free surface.
24th Symposium on Naval hydrodynamics, ACROS Fukuoka, Fukuoka, Japan.
4. Azcueta, R.(2004). Steady and unsteady RANSE simulations for Littoral Combat Ships. 25th Symposium
on Naval Hydrodynamics, St. John's, Newfoundland.
5. Ahmed, Y. and Soares, C. G.(2009). Simulation of free surface flow around a VLCC hull using viscous
and potential flow methods. Ocean Engineering, Vol. 36, No. 9, 691-696.
6. Wackers, J., Koren, B., Raven, H.C., van der Ploeg, A., Starke, A.R., Deng, G.B., Queutey, P., Visonneau,
M., Hino, T., and Ohashi, K. (2011). Free-Surface Viscous Flow Solution Methods for Ship
Hydrodynamics. Archives of Computational Methods in Engineering, Vol. 18, No.1, 1-41.
7. Banks, J., Phillips, A., and Turnock, S.(2010). Free surface CFD prediction of components of ship
resistance for KCS. 13th Numerical Towing Tank Symposium, Duisburg, DE.
8. Larsson, L., Stern, F., and Visonneau, M. (2011). CFD in ship hydrodynamics—results of the Gothenburg
2010 workshop, in MARINE 2011, IV International Conference on Computational Methods in Marine
Engineering, MARINE.
9. Hino, T., Gijutsu, K., and Kenkyūjo, A. (2005) The Proceedings of CFD Workshop, Tokyo.
10. Stern, F., Yang, J., Wang, Z., Sadat-Hosseini, H., Mousaviraad, M., Bhushan, S. and Xing, T. (2013).
Computational ship hydrodynamics: Nowadays and way forward, International Shipbuilding Progress,
Vol. 60, No.1-4, 3-105.
11. Raven, H. C., Van der Ploeg, A., Starke, A., and Eca, L.(2008). Towards a CFD-based prediction of ship
performance—progress in predicting full-scale resistance and scale effects. Proceedings of RINA-CFD,
London, UK.
12. Hirt, C. W. and Nichols, B. D. (1981). Volume of fluid (VOF) method for the dynamics of free boundaries.
Journal of Computational Physics, Vol. 39, No.1, 201-225.
13. Osher, S. and Sethian, J. A. (1988). Fronts propagating with curvature-dependent speed: Algorithms based
on Hamilton-Jacobi formulations. Journal of Computational Physics, Vol. 79, No.1, 12-49.
14. ITTC (2011). The specialist Committee on computational fluid dynamics. Proceedings of 26th
International Towing Tank Conference, Rio de Janeiro, Brazil.
15. Binh, N. T. (2015). Computational Fluid Dynamics study for experimental fishing vessel M1317A. Nha
Trang Univeristt, Maste thesis in Vietnamese.
16. Patankar, S. V. and Spalding, D. B. (1972). A calculation procedure for heat, mass and momentum transfer
in three-dimensional parabolic flows. International journal of heat and mass transfer, Vol. 15, No.10,
1787-1806.
17. Muzaferija, S., Peric, M., Sames, P., and Schellin, T.E. (1998). A two-fluid Navier-Stokes solver to
simulate water entry. Proceedings of the 22nd symposium on naval hydrodynamics, Washington DC,USA.
18. ITTC (2011). Practical Guidelines for Ship CFD Applications. ITTC – Recommended Procedures and
Guidelines.
19. Kelvin, L. (1906). On deep-water two-dimensional waves produced by any given initiating disturbance
Proceedings of the Royal Society of Edinburgh, Vol. 25, No. 1, 185-196

12

Vous aimerez peut-être aussi