You are on page 1of 17

Hydrodynamic Investigation of the Ballast-Free Ship Concept

Miltiadis Kotinis, (AM), Assistant Professor, SUNY Maritime College


Michael G. Parsons, (FL), Arthur F. Thurnau Professor, University of Michigan

The hydrodynamic aspect of the Ballast-Free Ship concept is further investigated in order to analyze the water
discharge effect on the resistance and propulsion of a vessel. For this purpose, a Seaway-sized bulk carrier model
was tested in the towing tank of the University of Michigan Marine Hydrodynamics Laboratory. Additional results
were obtained with the aid of Computational Fluid Dynamics. The cost effectiveness of the concept is demonstrated
through a detailed economic analysis.

NOMENCLATURE water convention extends the biota control to sizes between 50


ACC - annual cargo capacity and 10 microns and adds specific requirements for selected
B - bias error smaller pathogens (i.e., toxicogenic Vibrio cholerae,
CP - pressure drag coefficient Escherichia coli, and intestinal enterococci). Even though the
CF - friction drag coefficient introduction of the requirements will not be in full effect until
CRF - Capital Recovery Factor 2014 for existing vessels; starting in 2009, ships without an
CT - total (friction plus pressure) drag coefficient approved ballast treatment system will have to conduct ballast
D - experimental form factor water exchange on the high seas, a method that has been
E - comparison error criticized regarding its effectiveness (Forsberg et al. 2005, Kent
k - form factor and Parsons 2004).
k-ε - turbulence closure model
k-ω - turbulence closure model A combination of a primary treatment system (e.g., filtration
P - precision error with backflush) to remove the larger organisms and a secondary
Qm - model-scale ballast trunk flow rate treatment system (UV irradiation, chemical treatment) has
Qs - full-scale ballast trunk flow rate proven to be effective in many cases. The installation of a
rG - grid refinement ratio ballast water treatment system, though, significantly increases
UD - experimental data uncertainty the capital and operating costs for trans-oceanic vessels. An
UG - grid uncertainty alternative to ballast treatment systems is the Ballast-Free Ship
UI - iterative uncertainty concept (U.S. Patent #6,694,908 2004), a new paradigm that
USN - simulation numerical uncertainty approaches ballast operation as the reduction of buoyancy rather
UT - total experimental uncertainty than the addition of weight to get the vessel to its required
UV - validation uncertainty ballast drafts. The feasibility and effectiveness of the Ballast-
y+ - non-dimensional distance to centroid of Free Ship concept has been demonstrated (Kotinis et al. 2004).
cell nearest wall This is a sequel to that earlier SNAME paper.
∆A - change in annual cost
In the Ballast-Free Ship concept, the traditional ballast tanks are
∆P - change in capital cost
replaced by longitudinal trunks that extend beneath the cargo
∆RFR - change in Required Freight Rate
region of the ship below the ballast draft. The trunks are
λ - geometric scale ratio connected to the sea through symmetrical with respect to the
ship centerplane plena; one at the bow and one at the stern. The
INTRODUCTION pressure developed on a moving ship is positive close to the
The detrimental environmental and socio-economic effects of bow stagnation point and negative (suction) near the stern.
the introduction of nonindigenous aquatic species into coastal Therefore, the hydrodynamic pressure differential between the
waters have motivated the international maritime regulatory bow and the stern region of a ship is theoretically able to induce
bodies to adopt strict regulations in order to control and a slow water flow inside the ballast trunks without the need for
eventually eliminate this phenomenon. The International pumps. While the ship is in the ballast condition, these trunks
Maritime Organization (IMO) has recently adopted new are constantly flooded with “local seawater,” thus, reducing the
regulations (IMO 2004) applicable to both new and existing buoyancy of the vessel. The net result is the elimination of the
vessels that conduct ballast operations. These regulations transportation of foreign ballast water and the nonindigenous
mandate the utilization of a ballast water treatment system and aquatic species it may contain.
prescribe strict limits regarding the concentration of viable
organisms in the discharged ballast water. The new IMO ballast

1
In this paper, the hydrodynamic aspect of the concept is further condition corresponds to heavy ballast used under extreme
investigated in an attempt to better understand the effect of the weather conditions.
water discharge on the flow around the ship hull and the
propeller operation. The current research focuses on ships A significant parameter in determining the geometric scale
deballasting in the Great Lakes; however, this does not limit the factor of the Ballast-Free bulk carrier model was the size of the
validity or applicability of the results. A scaled model of a available stock propellers. The No. 23 stock propeller at the
typical Seaway-sized bulk carrier was constructed and utilized MHL was the available propeller providing the highest
during resistance and propulsion tests performed at the propulsive efficiency and, at the same time, satisfying the hull
University of Michigan Marine Hydrodynamics Laboratory clearance requirements, assuming a full-scale propeller diameter
(MHL). The experimental results were supplemented by of 6.0 m. The corresponding geometric scale ratio λ = 37.92.
theoretical (numerical) results obtained through Computational Constraints relative to the avoidance of the blockage and
Fluid Dynamics (CFD) simulations performed using the shallow water effects were also taken into account during this
commercial CFD software FLUENT® (Fluent 2006). The selection process. The propeller characteristics for the No. 23
numerical investigation was performed before the experimental model propeller are shown in Table 2. The non-dimensional
investigation. First, the flow around the hull without any water thrust and torque coefficients plotted versus the coefficient of
suction or discharge was analyzed (initial case). Subsequently, advance (Kt, Kq – J) of the No. 23 model propeller are shown in
an analysis was attempted for the effect of the water suction on Fig.1. The main particulars of the Ballast-Free bulk carrier
the flow at the bow and the effect of the water discharge on the model in the ballast condition are listed in Table 3. The bow
flow at the stern (modified case). Finally, the numerical results and the stern of the constructed model are shown in Figs. 2 and
were compared with the experimental results, and a validation 3, respectively.
procedure was performed for the integral variables of the initial
case. Table 2. Characteristics of the MHL No. 23 Propeller

BALLAST-FREE BULK CARRIER MODEL Number of blades 4


The initial investigation of the Ballast-Free Ship concept Diameter (m) 0.158
demonstrated the general feasibility of the concept (Kotinis et al. Hub diameter (m) 0.031
2004, Kotinis 2005), with an emphasis on the invasive aquatic Pitch-diameter ratio 1.08
species problem as encountered in the Great Lakes. The Expanded area ratio 0.55
majority of the transoceanic vessels entering the Great Lakes
through the St. Lawrence Seaway are Handymax bulk carriers
(Farley 1996). The current research attempts to analyze the
impact of the ballast trunk discharge on the flow around the hull
of a typical Seaway-sized bulk carrier in addition to its effect on
the propeller operation.

For this purpose, a typical Seaway-sized Handymax bulk carrier


was designed and utilized in towing tank experiments and a
numerical hydrodynamic investigation. The bulk carrier was
designed using data from similar vessels. These vessels were
the Isa owned by the Polish Steamship Company and a bulk
carrier designed by the Jiangnan shipyard in China. The vessel
was designed using the commercial ship design software
Maxsurf® (Formation Design Systems 2006). The main
particulars of the vessel are presented in Table 1.
Figure 1. Propeller Coefficients versus Advance Coefficient
Table 1. Main Particulars of the Ballast-Free Bulk Carrier

Waterline length (m) 195.5


Table 3. Characteristics of the Ballast-Free Bulk
Maximum beam (m) 23.76
Carrier Model in Ballast Condition
Depth to main deck 16.00
Full-load draft (m) 10.70 Waterline length (m) 5.00
Block coefficient 0.835 Maximum beam (m) 0.627
F.P. draft @ 40% DWL (m) 0.113
The service speed is assumed to be 14.5 knots and the speed of A.P. draft @ 70% DWL (m) 0.198
the vessel while at ballast drafts 15.5 knots. During ballast
Wetted surface area (m2) 5.34
operation, the ship drafts at the forward and aft perpendiculars
Geometric scale ratio λ 37.92
are 40% and 70% of the design draft, respectively. This

2
The friction and pressure drag coefficients on the hull and the
nominal wake fraction at the propeller plane were monitored to
ensure the convergence of the integral variables. The finest grid
was utilized for the hydrodynamic investigation and the
selection of the inlet/outlet positions. The post-processing of the
CFD results was aided by the utilization of Tecplot 360®
(Tecplot 2006).

In the modified case, the water inlet and discharge locations


were modeled and their effect on the hull resistance was
investigated. An attempt was also made to analyze the effect on
propulsion in a qualitative manner.
Figure 2. Bow View of the Seaway-sized Bulk Carrier Model
Description of the Numerical Solver
The numerical solver of FLUENT is based on a finite volume
method with the flow properties calculated at the cell centers.
The mesh, in 3-D problems, can be either structured or
unstructured, single-block or multi-block, with cells of various
topologies (hexahedral, tetrahedral, pyramidal, and prismatic).
For incompressible flows, a pressure-based segregated solver is
typically utilized. The fluid velocity is obtained by the solution
of the Reynolds-Averaged Navier-Stokes (RANS) equations.
The pressure-velocity coupling is achieved through an iterative
procedure where a pressure correction equation is utilized to
ensure conservation of mass and eventually compute the
Figure 3. Stern View of the Seaway-sized Bulk Carrier pressure field. For steady-state flows, either the SIMPLE or
Model SIMPLEC algorithms are usually employed. The diffusion
terms in the RANS equations are discretized with a central
NUMERICAL HYDRODYNAMIC differencing scheme. The convection terms are discretized
INVESTIGATION using a higher-order upwind scheme to minimize numerical
The commercial CFD software FLUENT® (Fluent 2006) was diffusion; for unstructured meshes with tetrahedral elements a
utilized to study the external flow around the bulk carrier model. second-order upwind scheme is typically utilized. The
An IGES file of the model-scale design was imported into discretized equations are solved using the Gauss-Seidel iterative
Gridgen (Gridgen 2007) and the flow domain was meshed
® algorithm. The solution convergence is accelerated through the
using a multi-block hybrid grid, which is shown in Fig. 4. A utilization of an algebraic multi-grid method. Further details of
'double-body' flow model, which does not take the free-surface the numerical methods can be found in (Mathur et al. 1997) and
flow into account, was adopted considering the free surface as a (Kim et al. 1998).
plane of symmetry. In addition to this, only half of the hull
surface was considered assuming the flow to be symmetric with Turbulence and Near-Wall Modeling
respect to the centerplane. The shape of the flow domain was a The turbulence model utilized in the present computations was
quarter-cylinder; the outer cylindrical boundary, which was the shear-stress transport (SST) model (Menter 1994). This
treated as a slip wall, has a radius equal to half a ship length. At model implements a blending function in order to apply the
the velocity inlet, which was placed at a distance equal to half a standard k - ω model close to solid boundaries (ship hull) and a
ship length upstream of the forward stagnation point, a uniform transformed version of the k - ε model in the far field. The SST
velocity boundary condition was imposed. The flow domain model has been shown to provide quite accurate results for ship
was truncated downstream of the hull stern at a distance flows (Kim and Rhee 2002, Duvigneau et al. 2002).
approximately equal to one ship length. At this location, the
flow had essentially regained its parallel character; thus, it was A mesh capable of fully resolving the viscosity-affected near-
being unaffected by the presence of the ship hull. At the outlet wall region was constructed. This was deemed necessary,
of the flow domain, a Neumann boundary condition was especially in the modified case where the interaction of the
imposed and the flow properties were obtained by extrapolating water discharge with the boundary layer flow needs to be
the values from the interior of the domain. adequately modeled. In addition to this, a low-Reynolds-
number version of the k - ω model, which applies a damping
In order to obtain a converged solution for the initial case, three coefficient to the turbulent viscosity, was employed.
different grids were created using a systematic refinement
method and a verification procedure was applied to the results.

3
Figure 4. Computational Domain for the Ballast-Free Bulk Carrier

Numerical Results
Computations were performed for the Ballast-Free bulk carrier Table 4. Grid Dimensions and y+ Values
model in the ballast condition, where the model-scale speed is
1.295 m/s and the corresponding Reynolds number (in fresh Grid Total Number of y+
water at 15°C) is 6.10e+6. The corresponding friction Cells
coefficient based on the ITTC model-ship correlation line is Coarse 510,613 4.6
3.28e-3. The stopping criterion for the simulations was the Medium 715,410 3.3
convergence of the total resistance coefficient and the reduction Fine 1,019,973 2.4
of the solution residuals by four orders of magnitude. Prior to
analyzing the numerical results, a grid convergence study was
performed. The dimensions of the three grids utilized for this Table 5. Grid Convergence Study
grid study are listed in Table 4 along with the y+ values. The y+
value is the non-dimensional distance between the centroid of Grid Coarse Medium Fine
the cell adjacent to the solid boundary and the solid boundary. CP 1.167e-3 1.155e-3 1.154e-3
A systematic refinement was achieved by adjusting the grid Change (%) -1.0 -0.1
spacing on the hull surface and in the flow domain around the CF 2.932e-3 2.975e-3 2.999e-3
hull so that the resulting grid refinement ratio between Change (%) 1.5 0.8
consecutive grids was rG = 2 . CT 4.099e-3 4.130e-3 4.153e-3
Change (%) 0.8 0.6
Verification Procedure. A verification study was performed Form Factor k 0.250 0.259 0.266
using the methodology developed in (Stern et al. 2001) in order Change (%) 3.8 2.6
to assess the simulation numerical uncertainty. The computed
values of the pressure, friction, and total drag coefficients along The total drag coefficient listed above is not the overall drag
with the corresponding form factor are listed in Table 5. coefficient as it does not include the contribution of the wave
drag nor its interaction with the viscous drag. On the other
hand, the 'double body' flow model is ideal for the estimation of
the form factor. This is why the latter was selected as the
integral variable to utilize in the verification and validation
procedure.

4
The results in Table 5 show that the friction drag coefficient CF
is monotonically convergent and the pressure drag coefficient
CP is grid-independent. The physics behind the friction and
pressure drag coefficients are different; a fact that partially
justifies the different convergence behavior. The total drag
coefficient CT was monotonically convergent with a change of
only 0.6% between the medium and fine grid. The convergence
ratio is defined as the solution change between the medium and
fine grid over the solution change between the coarse and
medium grid. The convergence ratio for the total drag
coefficient is 0.74, a value that demonstrates monotonic
convergence. The form factor k is computed by dividing the Figure 5. Pressure Coefficient Contours at the Bow of the
computed total 'double-body' drag coefficient by the ITTC Ballast-Free Bulk Carrier Model – Initial Case
friction coefficient value. The value of the form factor and its
corresponding uncertainty was also determined during the The corresponding pressure coefficient contours at the stern are
towing tank experiments. presented in Fig.7. Suction pressure exists over the parallel
section and most of the ship stern, with a peak contour value
The simulation numerical uncertainty, USN, for steady-state close to the keel at Station 18 ( x/L = 0.90). Between Stations
simulations consists of iterative uncertainty, UI, and grid 17 and 18 (0.85≤x/L≤0.90) and close to the free surface, a lower
uncertainty, UG. Iterative uncertainty, which is defined as half suction pressure region exists, which combined with the suction
the range of maximum and minimum values of the integral pressure peak produces a considerable girthwise pressure
variable over the last two periods of oscillation (before the gradient. The latter results in the formation of a streamwise
stopping criterion is reached), was monitored and found to be vortex in that area; a typical phenomenon for ships with full hull
negligible compared to grid uncertainty. The estimated order of shapes.
accuracy based on the solution change between grids and the
refinement ratio value is 0.86. This value is relatively far from
the theoretical value of 2.0, which is the formal order of
accuracy of the utilized spatial discretization formulation.
However, this result is not uncommon for non-orthogonal, non-
uniform grids (Wilson et al. 2001). The grid uncertainty of the
uncorrected solution is estimated using a generalized
Richardson Extrapolation. The estimated value is equal to about
7.6% of the computed value using the fine grid.

Analysis of Flow around the Hull of the Ballast-Free Bulk


Carrier Model and Investigation of Nominal Wake. The
pressure coefficient contours and the velocity vectors at the bow Figure 7. Pressure Coefficient Contours at the Stern of the
of the Ballast-Free bulk carrier model in the ballast condition are Ballast-Free Bulk Carrier Model – Initial Case
shown in Figs. 5 and 6, respectively. Large pressure gradients
exist in a small region around the stagnation point, with a Axial vorticity contours at x/L = 0.93 and x/L = 0.95 are
significant pressure relief occurring immediately after that displayed in Figs. 8 and 9, respectively. At x/L = 0.93, the
region. The positive pressure area at the bow extends up to vorticity contours reveal the existence of a streamwise bilge
approximately 7% of the ship length aft of the F.P.; thus, the vortex. This vortex crosses the propeller plane, as demonstrated
available locations for the inlet of the bow plenum are limited. in Fig. 9. In the same figure, the existence of a small counter-
In order to ensure an adequate pressure differential to drive the rotating vortex at the base of the bulbous stern is also displayed.
ballast trunk flow, it was decided to place the water inlet at the
center of the bulbous bow and close to the keel to take full The nominal wake at the propeller plane was investigated in
advantage of the stagnation pressure in that area and to avoid order to help understand and quantify the effect of the
potential problems due to bow emergence. The velocity vectors longitudinal bilge vortex. Axial velocity contours (non-
depicted in Fig. 6 show the downward fluid flow induced by the dimensionalized by the parallel flow velocity) and axial vorticity
bulbous bow, even though the ballast condition is not the contours in the propeller disk are shown in Figs. 10 and 11,
operating condition for which the bulbous bow has been respectively. In order to investigate the uniformity of the flow
optimized. This downward flow is expected to reduce the in the propeller disk, the simplistic method of computing the
height of the bow wave; a low bow wave height was observed, standard deviation of the axial velocity was applied. The
indeed, during the towing tank model tests. computed value is 0.205 m/s. The area-weighted average value
-1
of the axial vorticity is 10.7 s .

5
Figure 6. Velocity Vectors at the Bow of the Ballast-Free Bulk Carrier Model – Initial Case

Figure 8. Axial Vorticity Contours at x/L = 0.93 – Initial Figure 9. Axial Vorticity Contours at x/L = 0.95 – Initial
Case Case (the propeller disk is also shown in the figure)

6
Selection of Inlet and Outlet Plena Locations
Based on the aforementioned flow analysis, it was decided to
locate the water inlet right on the face of the bulbous bow in the
area around the stagnation point to take advantage of the high
positive pressure in this region. In this way, the water exchange
goal of 99% in less than two hours can be reached, or even
exceeded (Kotinis 2005). For the utilized ballast speed of 15.5
knots (assuming no voluntary speed reduction due to heavy
weather), the flow exchange will be achieved in a distance less
than 30 nautical miles. A similar approach regarding the water
inlet location selection has been followed successfully by
Teekay Shipping for a source of pressure to drive flow-through
ballast exchange, at a reduced ballast tank level, without the use
of pumps (BWN 2002).

A full-scale diameter of approximately 1 m was chosen for the


plena inlet and outlet to ensure a smooth inflow and outflow
without imposing severe constraints on the structural
arrangements. The corresponding inlet/outlet diameter value in
model scale is approximately 2.6 cm. The flow rate in the
longitudinal trunks was calculated assuming a full-scale volume
of ballast water equal to 18,500 m3. This value, which was
Figure 10. Axial Velocity Contours in the Propeller Disk – obtained from similar ships, is based upon flooding both the
Initial Case normal ballast tanks and a central cargo hold for a heavy
weather ballast condition. Assuming an exchange time of 90
min and utilizing Froude scaling, the internal flow rate in model
scale is Qm = Qs λ-5/2 = 3.9·10-4 m3/s. Using the continuity
equation and assuming a symmetrical plenum about the
centerplane, the discharge fluid speed is 0.382 m/s.

The centroid of the water inlet was placed a distance


approximately 25% DWL above the keel as shown in Fig. 12.
The relatively low forward draft in ballast condition (40%
DWL) imposes the controlling constraint regarding this
selection. The selected position was a trade-off between
structural arrangement feasibility and avoidance of water inlet
emergence.

The location of the water discharge can only be determined after


a careful consideration of the pressure coefficient contours at the
stern and the axial velocity contours in the propeller disk. If
trunk flow rate maximization were the only criterion, the water
outlet should be located in an area with high suction pressure to
maximize the pressure differential. On the other hand, when the
propeller operation is taken into account, the objective is to
minimize the power requirements subject to achieving adequate
flow. The discharge location then can only be determined
through a propeller flow analysis and propeller performance
optimization procedure. It is generally accepted that a more
Figure 11. Axial Vorticity Contours in the Propeller Disk – uniform wake and, thus, a more uniform loading on the
Initial Case propeller blades results in less required power for a given
delivered thrust. This criterion, though simplistic, can be
evaluated initially by merely computing the standard deviation
of the axial velocity in the propeller disk.

7
exact positioning of the discharge locations could only be
determined after all the model construction and arrangement
constraints in the Ballast-Free bulk carrier model were taken
into account. The major constraints are the available interior
space of the model for the testing equipment (i.e. propeller
motor, dynamometer, heave staff) and the feasibility of drilling
holes and installing the pipes required for the trunk flow
modeling at certain locations inside the model.

Figure 12. Location of Forward Ballast Trunk Inlet

The primary difficulty in the discharge location selection


process stems from the fact that required power minimization
and trunk flow rate maximization constitute conflicting criteria
in the system optimization. This fact was demonstrated in the
initial investigation and development of the Ballast-Free Ship Figure 13. Location of Ballast Trunk Discharges
concept (Kotinis 2005). Observing the axial velocity contours
in Fig. 10, the upper half-plane is dominated by the primary Modeling and Numerical Investigation of the
longitudinal vortex, which provides a significant level of
homogenization, but also low fluid velocity, which reduces the
Water Inlet at the Bow and the Water Discharge
propeller loading in this region. On the other hand, a strong at the Stern
velocity gradient dominates the lower half-plane. Even though The effect of the water suction at the bow and the water
the secondary vortex covers a smaller region in the propeller discharge close to Station 17 on the flow around the model was
disk, it has considerable vorticity and opposes the action of the investigated numerically with the aid of FLUENT®. The
primary vortex. discharge flow direction was set to 10° with respect to the
surface tangent. Apparently the lower the angle value, the less
If the water outlet is placed in a position where it can supply the obstructive the water discharge would be to the boundary layer
fluid particles in the inner boundary layer with additional flow. This was suspected to be the reason behind the significant
momentum to delay the separation point and also provide increase in power requirements (+ 7.4% for a faster hull form)
additional momentum to the fluid that flows through the upper observed during the initial investigation of the Ballast-Free Ship
half-plane in the propeller disk, then a more uniform wake concept (Kotinis et al. 2004). A smaller angle, even though
might occur. On the other hand, given that the water suction is feasible in the numerical computations, would probably be an
located at a high positive pressure region and the longitudinal unattainable goal during the towing tank experiments or a full-
pressure gradient at the stern is not steep, the trunk flow rate is scale implementation.
not expected to be highly sensitive to the longitudinal position
of the water outlet. The major modeling requirement was to provide adequate grid
resolution to account for the interaction between the boundary
In order to investigate the effect of the water discharge location layer flow and the trunk inflow and outflow. This was
on the flow at the stern, it was decided to test two different accomplished by creating a mesh fine enough to resolve the
discharge locations during the towing tank experiments; one flow in the viscous sublayer, which is the portion of the
close to Station 17 and one close to Station 19 as shown in Fig. boundary layer adjacent to the hull solid boundary, in the region
13. Station 17 is approximately at the location of the forward around the inlet and discharge locations. For this purpose, a
engineroom bulkhead in the full-scale ship; Station 19 is slightly modified version of the fine grid created for the initial
approximately at the aft engineroom bulkhead. The discharge at case was utilized. The modifications were limited to the
Station 17 was located about 45% DWL and the discharge at modeling of the region close to the inlet and discharge locations.
Station 19 about 30% DWL. In this way, the effect on the flow In this case, the total number of cells is 1,074,580.
in the upper half-plane of the propeller disk could be
investigated, in addition to the effect of discharging water before
and after the formation of the primary longitudinal vortex. The

8
Figure 14. Velocity Vectors at the Bow of the Ballast-Free Bulk Carrier Model – Modified Case

The pressure coefficient contours at the bow are shown in Fig. 19 corroborate the aforementioned conclusion. The area-
15. A comparison with the pressure coefficient contours in Fig. weighted average value of axial vorticity is 10.5 s-1.
5 reveals that the positive pressure levels increase in the vicinity
of the water inlet, even though this effect vanishes downstream,
of x/L = 0.02. This observation is corroborated by the velocity
vectors displayed in Fig. 14. In the same figure, a slightly
accelerated fluid flow on the edge of the water inlet is displayed,
which can be attributed to the suction effect.

The pressure coefficient contours at the stern are displayed in


Fig. 16. The effect of the water discharge on the pressure
distribution seems to be limited to the vicinity of the discharge
location. A small reduction in suction pressure, relative to the
pressure distribution of the initial case shown in Fig. 7, is
observed slightly upstream of the discharge location. The Figure 15. Pressure Coefficient Contours at the Bow of the
opposite effect is observed slightly downstream. Therefore, the Ballast-Free Bulk Carrier Model – Modified Case
net effect on the pressure force is expected to be minimal.

The computed values of the friction and pressure drag


coefficients are 2.995e-3 and 1.195e-3, respectively. The total
drag coefficient is equal to 4.190e-3, an increased of 0.9%
compared to the initial case. The corresponding form factor
value is 0.277, approximately 4.1% higher, than the initial case.
This increase can be attributed mainly to the existence of the
water inlet at the bow .

Axial vorticity contours at x/L = 0.93 and x/L = 0.95 in the


initial case are demonstrated in Figs. 17 and 18, respectively. At Figure 16. Pressure Coefficient Contours at the Stern of the
x/L = 0.93, the vorticity contours reveal that the flow discharge Ballast-Free Bulk Carrier Model – Modified Case
close to Station 17 causes a slight stretching of the longitudinal
vortex in the transverse direction. Further downstream, at x/L = The axial velocity contours in the propeller disk for the modified
0.95, the interaction between the two counter-rotating vortices is case are plotted in Fig. 20. Compared to the initial case, the
significantly affected by the stretching of the primary vortex, as standard deviation of the axial velocity is slightly reduced. Its
the secondary vortex generated at the bulbous stern, now affects value is 0.201 m/s, which corresponds to a reduction of
a larger region of the lower half-plane in the propeller disk. approximately 2%. Even though this suggests a more
The axial vorticity contours in the propeller disk shown in Fig. homogeneous wake field, with a potential benefit for the

9
propeller operation, the changes in the wake distribution
necessitate the direct investigation of their effect on the
operating propeller.

Figure 19. Axial Vorticity Contours in the Propeller Disk –


Modified Case
Figure 17. Axial Vorticity Contours at x/L = 0.93 – Modified
Case

Figure 18. Axial Vorticity Contours at x/L = 0.95 – Modified Figure 20. Axial Velocity Contours in the Propeller Disk –
Case (the propeller disk is also shown in the figure) Modified Case

10
EXPERIMENTAL HYDRODYNAMIC
INVESTIGATION
The insight gained through the numerical hydrodynamic
investigation of the Ballast-Free Ship concept described above
was utilized as the starting point for the experimental towing
tank model test investigation. The experimental investigation
also provided the opportunity for a qualitative and quantitative
validation of the numerical results.

The experiments were performed at the University of Michigan


Marine Hydrodynamics Laboratory (MHL) towing tank using
the Ballast-Free bulk carrier model. The total resistance and the
propulsion requirements of the model were determined for three
different speeds in the ballast condition where the ballast trunks
would be used. All the tests were carried out at the ballast
drafts. The experimental test plan is shown in Table 6. Figure 21. Internal Flow Details in the Bow Region

Table 6. Ballast Condition Experimental Test Plan

Ship speed Model speed Froude


(knots) (m/s) number
1st speed 14.50 1.210 0.173
2nd speed 15.50 1.295 0.185
3rd speed 16.50 1.378 0.197

Four different measurements were obtained at each speed in an


attempt to minimize the precision error. In addition, a
randomized testing sequence was utilized to allow for
minimization of the effect of extraneous variables. Prior to the
model testing, a static calibration of the load cell was performed
in order to estimate the instrument error. The unmodified hull,
shown in Figs. 2 and 3, was tested first. Subsequently, the hull
was modified, as shown in Figs. 12 and 13, and the tests were
repeated, first, for the water discharge high close to Station 17
and then for the low discharge close to Station 19.
Figure 22. Internal Flow Details in the Stern Region –
Because the modeling of the internal flow trunks could not be Looking Forward
reliably scaled at the small model scale used, the scaled total
trunk flow was pumped rather than using natural flow. The
internal flow was modeled primarily by a 1-inch internal
diameter flexible pipe, which was connected to the water suction
at the bow and the water discharge at the stern. The internal
flow was aided by a flexible-impeller pump, which was used to
achieve the scaled flow rate. The flow rate was controlled by a
high-precision needle valve and monitored by a flow meter.
The flow was diverted to the selected discharge location and
subsequently split to provide a symmetric water discharge at the
stern of the model. Details of the internal pipe connections and
system are displayed in Figs. 21, 22 and 23. As already
mentioned, the model-scale internal flow rate was set to 3.9·10-4
m3/s or 6.1 gallons per minute.

Figure 23. Internal Flow Details – Pump and Flow Meter

11
Resistance Tests
The resistance of the Ballast-Free bulk carrier model was Table 8. Form Factor and Uncertainty
determined for the three speeds listed in Table 6. The measured
total resistance was corrected using the calibration results and Form factor k Uncertainty (%D)
then extrapolated to full scale by utilizing the ITTC- Initial case 0.291 6.0
recommended method (ITTC 1978). For all testing conditions, Discharge at St.17 0.259 9.4
the results are reported at a standard temperature of 15°C. The Discharge at St.19 0.282 6.3
uncertainty in the full-scale resistance and effective power was
estimated by taking into account both bias and precision errors. The experimental form factor value (D) of the baseline (initial)
Errors related to the static calibration and the form factor case listed in Table 8 can be utilized to validate the CFD results
derivation were considered as sources of bias error. During the using the methodology derived in Stern et al. (2001). The
uncertainty analysis, a 95% level of confidence was assumed. experimental data uncertainty, UD, combined with the computed
In general, the propagation of the bias limit from a single error simulation numerical uncertainty USN = 7.6%D, provides the
source to the result (i.e. full-scale resistance) was calculated by validation uncertainty, UV:
utilizing sensitivity indices and assuming that the error sources
are statistically independent. The sensitivity index relates how 2
U V = U SN + U 2D = 9.2%D (2)
changes in each of the error sources affect the result. The total
bias limit includes the contribution of all the bias error sources. This value is compared to the absolute value of the comparison
The total precision limit is a measure of the random variation in error, |E|, which is equal to the difference between the numerical
the results. It is equal to the standard deviation of the average result at the fine grid and the experimental result. In this case,
result value at each speed multiplied by the appropriate Student- the absolute value of the comparison error is equal to 8.4%D;
t distribution value based on the confidence interval and the thus, the form factor value is validated at the uncertainty level of
number of measurements. The total uncertainty is calculated as UV = 9.2%D. Further reducing the levels of validation
the root sum square of the total bias limit and the total precision uncertainty would require reduction in both experimental and
limit: simulation numerical uncertainty.

U T = B2 + P 2 (1) Ship Resistance and Effective Power Results. The full-scale


ship resistance and the corresponding effective power were
calculated for each of the three cases. The results are shown in
Static Calibration. A static calibration of the load cell was Figs. 24 and 25. The water discharge close to Station 17
performed by utilizing a sequential test. The results are shown appears to have a negative effect on ship resistance and, to a
in Table 7. It needs to be mentioned that the hysteresis error lesser extent, the same is true for the discharge close to Station
observed during the calibration procedure was negligible. The 19. This conclusion, regarding the effect of the water discharge
bias limit reported in Table 7 corresponds to the standard error close to Station 17 on the hull resistance, is supported,by the
due to the scatter in the load cell calibration dataset relative to a numerical results. The computed form factor value for the
linear least-squares regression curve fit assuming a 95% modified case is greater than the computed value for the initial
confidence interval. Other contributions to the bias limit of the case. The corresponding values based on the experimental
load cell measurement can be considered negligible (Longo and results showed a different trend; however, the experimental
Stern 2005). uncertainty was rather high, especially when discharging close
to Station 17. The analysis of the numerical results provided a
Table 7. Static Calibration Results reasonable explanation for the relative increase of the form
factor value. Despite the fact that the average resistance values
Calibration curve slope 1.0003 plotted in Fig. 24 reveal an increase in resistance, the difference
Zero drift -0.012 N with respect to the baseline case is not statistically significant as
Bias limit 0.036 N seen by the overlapping error bands.

Form Factor Calculation and Validation Procedure. The form Propulsion Tests
factor k was obtained by measuring the model resistance at low The resistance tests were followed by a series of propulsion tests
Froude numbers and then utilizing a linear least-squares using the MHL stock model propeller No. 23. The thrust and
regression curve fit. The intercept of the regression model torque measurements at the self-propulsion condition at each
corresponds to the form factor value plus 1. The form factor speed were analyzed using the ITTC-recommended method
uncertainty is also reported in Table 8. (ITTC 1978). The calculated delivered power is shown in Fig.
26. An uncertainty analysis was also performed for the
propulsion test results.

12
900

850

800

750
baseline
RTs (kN)

St.19
700
St.17

650

600

550

500
14.0 15.0 16.0 17.0
Ship Speed (knots)

Figure 24. Calculated Ship Total Resistance

8,000

7,000

baseline
Effective Power (kW)

St.19
6,000 St.17

5,000

4,000

3,000
14.0 15.0 16.0 17.0
Ship Speed (knots)

Figure 25. Calculated Effective Power

13
11,000

10,000

9,000
Delivered Power (kW)

baseline
St.19
St.17
8,000

7,000

6,000

5,000
14.0 15.0 16.0 17.0
Ship Speed (knots)

Figure 26. Calculated Delivered Power

The plotted results reveal a significant reduction in the powering unchanged. Thus, an appropriate measure of merit is the change
requirements caused by the water discharge at the stern. It in the Required Freight Rate (∆RFR) using the equation
needs to be emphasized that these results were obtained with a (Mackey et al. 2000):
non wake-adapted stock propeller. This reduction can be
attributed to increased homogenization of the wake field in the ∆RFR = (CRF(i, n)· ∆P + ∆A)/ACC (3)
propeller disk, as demonstrated in the numerical investigation,
but also to changes in the inflow to the propeller, as shown in where ∆P is the change in the capital cost, ∆A is the change in
Figs. 10 and 20. The physics behind the significant reduction in the annual operating cost, ACC is the constant annual cargo
powering requirements due to the water discharge close to
capacity, and CRF(i, n) is the Capital Recovery Factor for an i
Station 17 cannot be explained further without a detailed
return on investment over a ship life of n years.
analysis of the effective wake. However, judging from the
numerical results obtained for the discharge close to Station 17,
A realistic scenario was adopted for the economic analysis: a
it can be speculated that the interaction between the water
discharge and the longitudinal vortices contributes to the Handymax bulk carrier transporting grain from the upper Great
considerable change in delivered power. Lakes (e.g. Duluth, Thunder Bay) to ports in Northern Europe
and occasionally transporting steel into the Great Lakes. A
At the ballast speed of 15.5 knots, the reduction in delivered North Atlantic voyage route between Rotterdam and Montreal,
power is 7.3% for the discharge close to Station 17 and 2.1% for entering the Great Lakes through the St. Lawrence Seaway
the discharge close to Station 19. The location near Station 17 while in a ballast condition, is assumed.
would also be preferred from an engineroom arrangements
viewpoint since the ballast trunks would not have to be carried A major, conservative assumption is that the 2.1% and 7.3 %
through the engineroom. reductions in the required power of the Ballast Free bulk carrier
will not be enough to permit a change in the main engine; thus,
ECONOMIC ANALYSIS no propulsion machinery capital cost reduction is included. For
The economic impact of the Ballast-Free Ship concept on the the specific vessel type investigated in this study, the reduction
capital and operating cost of a typical Seaway-sized bulk carrier in the delivered power indicated by the propulsion results would
was estimated in a manner similar to that used in the initial lead to decreased fuel use, assuming constant ship speed, but
investigation of the concept (Kotinis et al. 2004). Because the would probably not actually permit a change in number of
Ballast-Free bulk carrier was designed to maintain the same cylinders or a move to a smaller alternative engine model. The
grain capacity, its annual cargo capacity (ACC) would be

14
results for the water discharge close to Station 17 and close to useful insight regarding the physics of the flow and, to a certain
Station 19 are presented in Fig. 27. extent, provided quantitative results to compare with the
experimental outcomes.
Based on the results of the initial investigation (Kotinis 2005),
an increase in the block coefficient value from 0.835 to 0.841 The selection of the trunk flow inlet and outlet locations utilized
was required in order to permit the transportation of equal in the towing tank experiments was guided by these numerical
annual cargo capacity and offset the lost buoyancy of the results. The trunk flow inlet was located in the center of the
flooded plena at constant Seaway draft while fully loaded. The bulbous bow. Two different locations were tested for the water
increase in ship powering requirements due to the higher block discharge: one high close to Station 17 and one lower close to
coefficient value was factored in the analysis input data. The Station 19.
increase in the hull steel weight results in a moderate increase in
the hull steel cost. Foreign new construction, typical of Korea, The experimental hydrodynamic investigation consisted of
was assumed for the calculation of the hull steel and other resistance and propulsion experiments in the towing tank. The
construction costs. The eliminated ballast water treatment analysis of the model test data revealed that the experimental
system was assumed to consist of automatic backflush filtration results were in good agreement with the numerical results.
as a primary treatment combined with UV irradiation for a
secondary treatment. The estimated cost of this treatment In general, the water discharge at the stern was shown to
equipment was based upon a study commissioned by the Great increase ship resistance. On the other hand, it has a very
Lakes Ballast Technology Demonstration Project (Hurley et al. favorable effect on the propulsion characteristics for this
2001). Several other costs relative to changes in arrangements Seaway-sized bulk carrier design. The reduction in powering
and outfitting are also taken into account. A detailed description requirements, compared to the baseline case, was 7.3% for water
of these can be found in (Kotinis 2005). discharge close to Station 17 and 2.1% for water discharge close
to Station 19 at the assumed ballast speed of 15.5 knots.
It needs to be emphasized that no account was taken for the
power consumption, crew workload, and maintenance savings It must be cautioned that since some of the effect of the water
associated with the elimination of the ballast water treatment discharge may be related to the location of the separation near
system. This would further reduce the annual operating cost and the stern, and this is known to not scale effectively for models of
favor the use of the Ballast-Free Ship concept. The lifecycle the size used in this testing, the exact magnitude of the full-scale
cost of the aforementioned ballast treatment system would power improvement cannot be stated absolutely. The 5-m
probably exceed $800,000 (Hurley et al. 2001). model used in this experimental work was the largest that could
be tested in the University of Michigan MHL so further
The net savings in terms of the ∆RFR with the ballast trunk confirmation would have to be made in a larger facility that
water discharge close to Station 17 is estimated to be about would permit the use of a larger model.
$0.93 per ton of cargo. The corresponding savings with the
water discharge close to Station 19 is estimated to be about A realistic operating scenario was adopted in order to
$0.44 per ton of cargo. These savings are relative to the use of investigate the economic impact of the Ballast-Free Ship
filtration primary and UV secondary ballast water treatment concept. Based on the aforementioned reduction in powering
when ballast water exchange is no longer permitted in the requirements, the net savings with respect to a ballast treatment
future. system would be $0.93 per ton of cargo for the water discharge
close to Station 17 and $0.44 per ton of cargo for the water
discharge close to Station 19.
SUMMARY AND CONCLUSIONS
The initial development and investigation of the Ballast-Free
The Ballast-Free Ship concept essentially eliminates the
Ship concept demonstrated its feasibility and efficiency, but
transport of foreign ballast water from ships operating in the
failed to show its full cost-effectiveness. The main reason was
ballast condition. The elimination of the fuel penalty found
the significant fuel penalty, resulting from increased power
earlier makes the Ballast-Free Ship concept a very attractive
requirements found in the initial preliminary testing of an
alternative to more costly ballast water treatment systems.
existing, higher-speed vessel with a non-optimum propeller
(Kotinis et al. 2004).
Further numerical hydrodynamic investigation that includes the
propeller effect is expected to provide an optimum location for
The current paper focused on a more detailed hydrodynamic
the water discharge position. This will be confirmed by
investigation of the concept for a Seaway-sized bulk carrier in
additional MHL testing planned in 2008.
order to obtain a more realistic assessment of these effects. For
this purpose, a new Seaway-sized bulk carrier model was
designed and built. ACKNOWLEDGEMENTS
This paper is a result of work sponsored by the Great Lakes
The hydrodynamic investigation was performed in two parts. In Maritime Research Institute (GLMRI), a Consortium of the
the first part, the numerical investigation using CFD provided University of Wisconsin, Superior, and University of Minnesota,

15
Duluth. The GLMRI is funded by the U. S. Department of KIM, S.E., and RHEE, S.H. 2002 “Assessment of Eight
Transportation, Maritime Administration, under contract Turbulence Models for a Three-Dimensional Boundary Layer
DTOS59-05-G-00019 and agreement DTMA1G0605. This involving Crossflow and Streamwise Vortices,” AIAA 2002-
financial support is gratefully acknowledged. 0852, 40th AIAA Aerospace Sciences Meeting and Exhibit,
Reno, NV, Jan.
REFERENCES
BWN 2002 “Alternative Exchange Design Proposed,” Ballast KOTINIS, M. 2005 “Development and Investigation of the
Water News, Global Ballast Water Management Program, Ballast-Free Ship Concept,” Ph.D. Dissertation, University of
IMO, 8, January-March. Michigan, Department of Naval Architecture and Marine
Engineering.
DUVIGNEAU, R., VISONNEAU, M. and DENG, G.B. 2002
“On the Role Played by Turbulence Closures in Hull Shape KOTINIS, M., PARSONS, M. G., LAMB, T. and SIRVIENTE,
Optimization at Model and Full Scale,” 24th Symposium on A. 2004 “Development and Investigation of the Ballast-Free
Naval Hydrodynamics, Fukuoka, Japan, August 8-13. Ship Concept,” Transactions SNAME, 112, pp. 206-240.

FARLEY, R. 1996 “Analysis of Overseas Vessel Transits Into LONGO, J. and STERN, F. 2005 “Uncertainty Assessment for
the Great Lakes and Resultant Distribution of Ballast Water,” Towing Tank Tests with Example for Surface Combatant
Report No. 331, University of Michigan, Department of Naval DTMB Model 5415,” Journal of Ship Research, 49(1), pp. 55-
Architecture and Marine Engineering. 68.

FLUENT 2006 “FLUENT 6.3 User’s Manual,” Fluent Inc., MACKEY, T. P., TAGG, R. D. and PARSONS, M. G. 2000
Lebanon, NH. “Technologies for Ballast Water Management,” 8th
ICMES2000/SNAME New York Metropolitan Section
Symposium, May 22-23.
FORMATION DESIGN SYSTEMS Pty 2006 “MAXSURF
Windows Version 12 User’s Manual.”
MATHUR, S.R. and MURTHY, J.Y. 1997 “A Pressure-Based
Method for Unstructured Meshes,” Numerical Heat Transfer,
FORSBERG, R., BAIER, R., MEYER, A., DOBLIN, M., and
31, pp. 195-215.
STROM, M. 2005 “Fine Particle Persistence in Ballast Water
Sediments and Ballast Tank Biofilms,” 28th Annual Meeting
MENTER, F.R. 1994 Two-Equation Eddy-Viscosity Turbulence
of The Adhesive Society, Mobile, Alabama, February 13-16.
Models for Engineering Applications,”
AIAA Journal, 32(8):1598-1605, August.
GRIDGEN 2007 “Gridgen 15.1 User Manual,” Pointwise Inc.,
Fort Worth, Texas.
STERN, F., WILSON, R.V., COLEMAN, H.W., and
PATERSON, E.G. 2001 “Comprehensive Approach to
HURLEY, W. L. JR., SCHILLING, S. S. JR. and MACKEY, T.
Verification and Validation of CFD Simulations – Part 1:
P. 2001 “Contract Designs for Ballast Water Treatment
Methodology and Procedures,” ASME Journal of Fluids
Systems on Containership R. J. Pfeiffer and Tanker Polar
Engineering, 123, pp. 793-802.
Endeavor,” SNAME/ ASNE Marine Environmental
Engineering Technical Symposium, Arlington, VA, May 31-
TECPLOT 2006 “Tecplot 360 User’s Manual,” Tecplot Inc.,
June 1 (CD).
Bellevue, WA.
IMO 2004 “International Convention for the Control and
U. S. PATENT #6,694,908 2004 “Ballast Free Ship System,” U.
Management of Ships’ Ballast Water & Sediments,”
S. Patent and Trademark Office, Washington, DC.
Diplomatic Conference, February, London.
WILSON, R.V., STERN, F., COLEMAN, H.W., and
ITTC 1978 “15th International Towing Tank Conference, 3-10
PATERSON E.G. 2001 “Comprehensive Approach to
September 1978, The Hague, The Netherlands”, Netherlands
Verification and Validation of CFD Simulations – Part 2:
Ship Model Basin, Wageningen.
Applications of RANS Simulation of a Cargo/Container Ship,”
ASME Journal of Fluids Engineering, 123, pp. 803-810.
KENT, C.P. and PARSONS M.G. 2004 “Computational Fluid
Dynamics Study of the Effectiveness of Flow-Though Ballast
Exchange,” Transactions SNAME, 112.

KIM, S.E., MATHUR, S.R., MURTHY, J.Y., and


CHOUDHURY, D. 1998 “A Reynolds-Averaged Navier-
Stokes Solver using Unstructured Mesh-Based Finite Volume
Scheme,” AIAA Paper 98-0231.

16
Vessel data and trip scenario Typical bulk carrier Ballast-free bulk carrier Comments
Discharge at St.17 Discharge at St.19
Round-trip distance (nautical miles) 6,280 Montreal (CAN) to Rotterdam (NL) through the Seaway
Service speed (kts) 14.5 Typical data for an ocean-going Handymax bulk carrier transporting grain cargo from the
Speed in ballast condition (kts) 15.5 Great Lakes (Duluth, Thunder Bay) to ports in Northern Europe and occasionally
Proportion of miles in ballast (%) 35 transporting steel into the Great Lakes.
Average loaded cargo / maximum cargo (%) 90
Load factor (%) 58.5
Days of navigation through the Great Lakes 8 Passage up through the Great Lakes towards the western end
Port days per round trip 14 Includes loading/unloading time, bunkering time, and time waiting for berth
Round trips per annum 7
Maximum payload (metric tons) 32,000
Cargo carried per annum (metric tons) 131,000
Engine nominal MCR (kW) 8,580 Data for the MAN B&W 6S50MC two-stroke engine
Block coefficient 0.835 0.841 Compensate for increased hull steel weight and lost buoyancy at plena
Hull steel weight (metric tons) 5,550 5,770
Hull steel cost ($) 2,220,000 2,308,000 Assuming a steel price of $400/metric ton
Continuous service rating in ballast condition (kW) 7,700 7,140 7,540 Includes 15% sea margin and effect of change in CB value
Continuous service rating in full load condition (kW) 7,700 7,155 7,555 Includes 15% sea margin and effect of inlet/outlet hull openings and change in CB value
Specific fuel consumption (g/(kW*hr)) 168.7 166.4 168.0 Data for the MAN B&W 6S50MC engine, ISO ambient conditions
Annual heavy fuel cost ($) 1,039,000 951,000 1,014,000 Fuel price (IFO380) of $270/metric ton, transatlantic part of trip only
Changes in capital cost
Additional hull steel cost ($) 88,000
Sluice gates cost ($) 260,000 Acquisition cost plus labor for 52 450x600 mm sluice gates (@ $5,000 each)
Elimination of ballast tank valves ($) -14,000 14 @ 1,000 each
Reduction in ballast piping cost ($) -314,000 Removal of main ballast headers (material plus labor)
Reduction in welding cost ($) -9,500 Reduced welding at the bottom of solid floors (material plus labor)
Additional ballast piping cost ($) 79,000 Addition of ballast piping for F.P. tank (material plus labor)
Additional welding cost ($) 2,600 Additional welding due to raise of inner bottom (material plus labor)
Elimination of ballast water treatment system ($) -375,000 Assuming automatic backflush filtration combined with UV irradiation
Changes in operating cost
Discharge at St.17 Discharge at St.19
Change in heavy fuel oil cost ($) -88,000 -25,000
Net capital cost change ($) -282,900
Net operating cost change per annum ($) -88,000 -25,000
Capital recovery factor 0.1175 i = 10%, n = 20 years
Change in required freight rate ($/metric ton) -0.93 -0.44 savings

Figure 27. Economics Summary Comparing a Typical Bulk Carrier with Filtration and UV Treatment with Ballast-Free Bulk Carrier with Two Discharge Locations

17

You might also like