design studio
design studio
Submitted by
MASTER OF TECHNOLOGY
IN
CIVIL ENGINEERING
(OFFSHORE STRUCTURES)
DEPARTMENT OF CIVILENGINEERING
NATIONAL INSTITUTE OF TECHNOLOGY
1
CONCLUSION .................................................................................................................... 63
REFERENCES .................................................................................................................... 63
5. FLOW PAST A 2D CIRCULAR CYLINDER AT REYNOLDS NUMBER 250 USING
ANSYS FLUENT AND VISUALIZATION OF VORTEX SHEDDING. ..................... 65
5.1 INTRODUCTION ......................................................................................................... 65
5.2 RESEARCH SIGNIFICANCE ...................................................................................... 67
5.3 NUMERICAL METHODOLOGY ................................................................................ 68
5.4 RESULTS AND DISCUSSION..................................................................................... 75
CONCLUSION .................................................................................................................... 82
REFERENCES .................................................................................................................... 83
2
LIST OF FIGURES
Page
Figure No Figure Name
No
Figure 1.1 MATLAB Code 12
Figure 1.2 ODE Code. 13
Figure 1.3 RMS Displacement VS Reduced velocity 14
Figure 1.4 Code for FFT Transformation. 15
Figure 1.5 Time domain and Frequency domain 17
Figure 2.1 Frequency vs amplitude plot 24
Figure 2.2 Deflected Shape (Self Load). 25
Figure 2.3 Deflected Shape (Wave Load) 26
Figure 2.4 Mode shape 1 27
Figure 2.5 Mode shape 2 27
Figure 2.6 Mode shape 4 28
Figure 2.7 Mode shape 4 28
Figure 2.8 Mode shape 5 29
Figure 2.9 Mode shape 6 29
Figure 2.10 Mode shape 10 30
Figure 2.11 Deflection VS Modes. 32
Figure 3.1 Geometry of cylinder. 39
Figure 3.2 Wave direction. 40
Figure 3.3 Boundary Conditions. 41
Figure 3.4 Mesh Generation. 42
Figure 3.5 Analysis setting. 43
Figure 3.6 Hydrostatic result. 44
Figure 3.7 Stability of Floating Body 47
Figure 3.8. RAO vs Frequency 49
Figure 4.1 Ship motion and axis 54
Figure 4.2 Static drift 56
Figure 4.3 Excel result 58
Figure 4.4 forces and yaw moment vs time 62
3
Figure 5.1 Regimes of fluid flow over circular cylinder. 66
Figure 5.2 Problem model 71
Figure 5.3 Computational domain 72
Figure 5.4 Mesh generation 73
Figure 5.5 Solver setup 74
Figure 5.6 Setup. 75
Figure 5.7 Lift Coefficient 75
Figure 5.8 Drag Coefficient 76
Figure 5.9 Frequency vs amplitude plot 76
Figure 5.10 Strouhal Number vs Reynolds Number plot. 78
Figure 5.11 Velocity contour at time 0.75 s. 79
Figure 5.12 Velocity contour at time 1.37s. 79
Figure 5.13 Static Pressure contour. 81
Figure 5.14 Streamlines and formation of vortices. 82
4
Table NO. Table Name Page No.
Table 2.1 Modes and frequency 31
Table 2.2 Modes and Deflection 32
Table 2.3 Modes and Moments 35
Table 3.1 Frequency and Amplitude 48
Table 4.1 Sway Force, Surge Force and 61
Yaw Moment
5
EXERCISE 1
1.1 INTRODUCTION
In fluid dynamics, vortex-induced vibrations (VIVs) refer to the oscillatory motions
experienced by bodies interacting with external fluid flows, caused by periodic irregularities in
the flow. A classic example is the VIV of an underwater cylinder. When a cylinder moves
through water perpendicular to its axis, the flow around it slows near its surface due to viscosity,
forming a boundary layer. As the boundary layer separates from the cylinder due to excessive
curvature, vortices form, altering the pressure distribution along its surface. Asymmetrical
vortex formation induces lift forces on opposite sides of the cylinder, resulting in transverse
motion. This interaction limits the vibration amplitude but leads to significant dynamic effects
(El-Reedy, 2015, p. 179).
VIV is particularly critical in offshore engineering, as it can cause fatigue damage to slender
structures such as oil exploration risers. These risers are subject to both steady current flows
and top-end vessel motions, creating complex flow-structure interactions and unsteady flow
profiles. Understanding and mitigating VIV is essential to ensure the structural integrity and
longevity of these systems.
6
3. Development of Predictive Models: Investigating VIV responses contributes to the
creation of accurate predictive models and simulation tools. These models assist in
forecasting the behavior of columns under various flow conditions, facilitating
informed decision-making during the design and assessment phases.
In summary, studying the response of columns to VIV is vital for ensuring structural safety,
optimizing design, developing effective mitigation measures, and minimizing economic and
environmental risks associated with these oscillatory phenomena.
As fluid flows past the cylinder, the flow separates from the surface at a certain point, forming
vortices in the wake of the cylinder. These vortices shed alternately from opposite sides,
creating a pattern known as the von Kármán vortex street. When a vortex is shed from one side,
the pressure on that side decreases, while the opposite side has higher pressure. This
asymmetric pressure generates a lift force perpendicular to the flow, causing the cylinder to
oscillate in the cross-flow direction.
7
In-Line: Parallel to the flow direction, influenced by changes in drag forces.
As the flow velocity increases, the frequency of vortex shedding (𝑓𝑠 ) can approach the natural
frequency (𝑓𝑛 ) of the cylinder’s elastic system.
When 𝑓𝑠 ≈𝑓𝑛 , a resonance condition occurs, known as lock-in, where the cylinder's vibration
frequency matches the vortex shedding frequency. In this state, large-amplitude oscillations
occur, as energy from the fluid is efficiently transferred to the structure. As the oscillation
amplitude increases, the interaction between the moving cylinder and the vortex-shedding
process changes. The motion of the cylinder modifies the wake structure, leading to a self-
limiting effect where the vibration amplitude reaches a steady-state value. This is a result of
the balance between energy input from the fluid and energy dissipated due to damping and
structural resistance (Kawai, 1993).
Cross-flow oscillations typically occur at a frequency close to the vortex shedding frequency
and In-line oscillations occur at approximately twice this frequency (second harmonic. The
combined effect can produce complex motion paths, such as elliptical trajectories.
The flow transfers kinetic energy to the cylinder, sustaining the oscillations.
• The amplitude and frequency of the response depend on factors such as:
o Mass ratio (ratio of the cylinder’s mass to the displaced fluid mass).
o The Strouhal number relates the frequency of shedding to the velocity of the
flow and a characteristic dimension of the body (diameter in the case of a
cylinder).
2.Governing equations
8
• Cross-flow (transverse to the flow)
Cross-Flow Direction:
In-Line Direction:
where:
• 𝑥: In-line displacement.
• 𝑦: Cross-flow displacement.
• 𝐹𝑥(𝑡), 𝐹𝑦(𝑡): Fluid forces acting on the cylinder in the respective directions.
The fluid forces (F𝑥(t) and F𝑦(t)) are generated by vortex shedding and are typically expressed
as:
𝐹(𝑡) = 𝐹0 cos(𝛺𝑓𝑡 )
1
𝐹0 = 𝐶 𝜌𝐷𝑈 2
2 𝐿
where:
• 𝜌: Fluid density
• 𝑈: Flow velocity
9
• 𝐶𝐿 : Time-dependent in-line and cross-flow force coefficients, which depend on the
vortex shedding dynamics (𝑉𝑟 and Re) (Narendran et al., 2015, p.8).
The frequency of vortex shedding, 𝛺𝑓 , is related to the flow velocity and the cylinder diameter
via the Strouhal number (𝑆𝑡 ):
𝑈
𝛺𝑓 = 2𝜋𝑆𝑡
𝐷
𝑈
𝑓𝑠 = 𝑆𝑡
𝐷
where:
• 𝑆𝑡 : Strouhal number (typically between 0.18 and 0.21 for a circular cylinder).
Rough surfaced cylinders or vibrating cylinders (both smooth and rough sur faced) have
Strouhal numbers that are relatively insensitive to the Reynolds number. For cross sections
with sharp corners, the vortex shedding is well defined for all velocities, giving Strouhal
numbers that are independent of flow velocity (Re). Structural damping is due to internal
friction forces of the member material and depends on the strain level and associated
deflection. For wind-exposed steel members, the structural damping ratio may be taken as
0.0015, if no other information is available. For slender elements in water, the structural
damping ratio at moderate deflection typically ranges from 0.005 for pure steel pipes to
0.03 0.04 for flexible pipes (El-Reedy, 2015, p. 182).
d. Non-Dimensionalized Form
10
The governing equations are often non-dimensionalised for analysis, using parameters such as:
𝑈
• Reduced velocity: 𝑉𝑟 = 𝑓 𝐷, where 𝑓𝑛 is the natural frequency of the cylinder.
𝑛
Solving the equation of motion of cylinder mentioned above will give us the response of
cylinder under vortex induced vibration.
1.Toolboxes: ODE Solvers (for solving differential equations, e.g., ode45, ode23).
2.Scripts:
11
Figure 1.1 MATLAB Code
• Ode_code.m to numerically solve the equations of motion using the ode45 solver.
12
Figure 1.2 ODE Code.
3.Parameters: The analysis used an EMRC subjected to flow velocity varying from
0.1m/s to 1m/s. The properties of EMRC are given below in the table
13
1.4 POST PROCESSING OF RESULTS
Using the MATLAB code shown in fig 1. and fig 2. VIV graph with Reduced velocity along
abscissa and Normalized RMS displacement along the ordinate as shown in fig 3.
The code from fig 1 is used to plot the time domain response of EMRC as displacement y(t)
along y-axis and time (seconds) along x-axis and these results are used to plot the
corresponding frequency domain responses using the Fast Fourier Transform (FFT) method.
The FFT decomposes a complex signal into its individual frequency components using a
mathematical process called the Fourier Transform. The MATLAB code used for the
transformation is shown in fig 4.
14
Figure 1.4 Code for FFT Transformation.
.
Following observations were seen when the time domain and frequency domain analysis was
done for an EMRC of diameter 0.1m is induced with waves of velocity ranging from 0.1m/s
to 1m/s.
15
16
Fig 5. Time domain and Frequency domain
17
In fig 5 left section shows the time domain analysis and right section displays respective
frequency domain analysis. Frequency domain is derived from time domain using FFT
method mentioned as earlier.
𝑈
From equation 𝑉𝑟 = 𝑓 , 𝑈 for natural frequency (𝑓𝑛 ) 1.2314 hz will be 0.6142m/s.
𝑛𝐷
𝑈
The shedding frequency 𝑓𝑠 = 𝑆𝑡 × 𝐷 , implying 𝑈 = 0.6142m/s, 𝐷 = 0.1𝑚 and 𝑆𝑡 = 0.2 will
18
From the above table it is shown that max amplitude 0.05 m is observed for 𝑈 = 0.61𝑚/𝑠
when frequency is 1.203 hz. For the above considering EMRC natural frequency 𝑓𝑛 = 1.23.
Hence from the all above discussions it is proven that here max response is observed when
𝒇 ≈ 𝒇𝒏 which is nothing but the Resonant condition. From the above 2 approaches it is
shown that when Velocity 𝑈 = 0.6142m/s at a frequency of 𝑓 = 1.2 hz maximum response
is observed.
CONCLUSION
The normalized RMS displacement increases rapidly as the reduced velocity (𝑉𝑟) approaches
a critical value of around 5, where it reaches a peak. This indicates that the structure
experiences the maximum amplitude of VIV at this critical velocity of 0.614m/s, likely due to
resonance when the vortex-shedding frequency matches the natural frequency of the
structure. Beyond this point, the displacement decreases, suggesting a detuning effect where
the excitation frequency moves away from the natural frequency. This behaviour highlights
the importance of avoiding certain reduced velocities during the design of structures prone to
VIV, such as offshore risers, pipelines, or cables, to prevent excessive vibrations and potential
structural fatigue.
REFERENCES
2. Narendran, K., Murali, K., & Sundar, V. (2015). Vortex-induced vibrations of elastically mounted
circular cylinder at Re of the O(10⁵). Journal of Fluids and Structures.
https://doi.org/10.1016/j.jfluidstructs.2014.12.006.
3. Kawai, H. (1993). Vortex induced vibration of circular cylinder. Journal of Wind Engineering and
Industrial Aerodynamics, 46-47(1), 605–610. https://doi.org/10.1016/0167-6105(93)90111-M
19
EXERCISE 2
2.1 INTRODUCTION
Offshore jacket platforms are widely used in the oil and gas industry to support drilling and
production facilities in shallow to medium water depths. These structures are subjected to
complex environmental loads, including wind, waves, currents, and seismic forces, making
their structural integrity and stability a critical aspect of design and operation. Ensuring the
reliability of offshore jacket platforms requires a comprehensive understanding of their static
and dynamic behaviour under various loading conditions.
The static analysis aims to find out the structural response under steady-state loads such
as self-weight, buoyancy, and environmental forces. This accounts with strength and
serviceability criteria. Whereas the dynamic analysis considers time-dependent effects such as
wave-induced vibrations, wind gusts, and seismic excitations, which play a significant role in
the fatigue and long-term performance of the platform.
• By analyzing the platform under different loading scenarios, this research provides
insights into optimizing structural design, leading to efficient material utilization, cost
savings, and improved sustainability without compromising safety standards.
• Offshore structures must adhere to stringent international design codes and standards
such as API RP 2A, ISO 19902, and DNVGL guidelines. This research aids in verifying
compliance with these standards, ensuring that the design meets legal and operational
requirements.
1.Environmental loads: Water force can be classified as forces due to waves and forces
due to current. Wind blowing over the ocean’s surface drags water along with it, thus
forming current and generating waves. The forces induced by ocean waves on platform are
dynamic in nature. However, it is the accepted practice to design shallow water platforms
by static approach. As a water depth increases and platforms become flexible, dynamic
effect becomes significant [9].
2.Waves: Regular wave theories used for calculation of wave forces on fixed offshore
structures are based on the three parameters water depth (d), wave height (h) and wave
period (T) as obtained from wave measurements adapted to different statistical models.
Wave plus current kinematics (velocity and acceleration fields) are generated using 5th
order Stokes wave theory, the forces on individual structural elements are calculated using
Morison equation, based on hydrodynamic drag and mass coefficients (Cd , Cm) and
particle velocity and acceleration obtained by the 5th order Stokes wave theory. Stokes 5th
order wave is defined by providing wave height and period in the input data with the wave
type specified as Stokes in the SACS options [7].
21
3.Wind loads: When a structure is placed in the path of the moving air so that wind is
stopped or is deflected from its path, then all or part of the kinetic energy is transformed
into the potential energy pressure. Wind forces on any structure therefore result from the
differential pressure caused by the obstruction to the free flow of the wind. These forces
are functions of the wind velocity, orientation, area, and shape of the structural elements.
Wind forces on a structure are a dynamic problem, but for design purposes, it is sufficient
to consider these forces as an equivalent static pressure [1].
Analysis software
SACS (structural analysis computer systems), a design and analysis software for offshore
structures and vessels, is used for the modelling and analysis of the jacket. SACS is an
integrated software of finite element-based software that supports the analysis, design and
fabrication of offshore structures, including oil, gas, and wind farm platforms and topsides
[7]. Its ability to dynamically iterate designs allows users to perform advanced analysis,
comply with offshore design criteria, and visualize the response and results of the structure.
SACS provides reliable beam member code checking and tubular joint code checking
capacity; therefore, it is very suitable for topsides structures and jackets consisting of
plate girders and tubular columns/ braces [4].
Modelling Data
Drawing-Details:
The inputs given are as follows:
• Water depth: 79.5 m
• Working point elevation: 4.0 m
• Pile connecting elevation: 4.0 m
• Mud line elevation, pile stub elevation, and leg extension elevation: -79.5 m
• Other intermediate elevations: -50.0, -21.0, 15.3 (cellar deck), 23.0 m (main deck)
Number of legs: 4
• Leg type: Ungrouted
• Leg spacing at working point: X1 = 15 m, Y1 = 10 m
• Leg spacing at Mud line: X1 = 20 m, Y1 = 15 m
Connectivity:
• Diagonal braced with horizontals
22
• Only horizontals at cellar and main deck
Inputs given in the structural definition gives to member groups with undefined properties.
The properties of these groups are LG1, LG2, LG3 are defined in segments as follows:
• Segment 1: D = 48.5 in, T = 1.75 in, Fy = 34.50 kN/cm²
• Segment 2: D = 47.0 in, T = 1.0 in, Fy = 24.80 kN/cm²
• Segment 3: D = 48.5 in, T = 1.75 in, Fy = 34.50 kN/cm²
• DL4 = 42"x1.5"
• DL5 = 42"x1.5"
Connections (both horizontal and diagonals):
• 30"x1" (flooded)
• PL* = 42"x1.5" (pile)
• W.B. = 30"x1" flooded
Where D is outer diameter and T is thickness.
Density of all members is 7.849 tonne/m³.
23
Figure 2.1. Jacket model
Static analysis refers to a structural analysis method where the forces acting on a structure are
considered to be constant and unchanging over time, allowing the software to calculate the
stresses, strains, and displacements within the structure without considering dynamic effects
like vibrations or wave loading.
The following responses were observed under static analysis of the structure. The static analysis
includes wave load interaction and self weight responses for the structure.
24
Figure 2.2 Deflected Shape (Self Load).
Figure two shows the response of structure under its own self weight. The self weight of
structure observed was -9205.33KN. The red representation in fig 2 is the deflected shape
from its original configuration. Under this loading maximum deflection observed was for
joint 304L in Y direction which is highlighted in fig 2. Max deflection occurs in joint at
almost at mid bay of structure.
25
Figure 2.3 Deflected Shape ( Wave Load)
Figure 3 represents the response of the jacket under a wave action. The wave induced was an
airy wave of height 10m and time period 12 seconds. The wave attack is at an angle of 120
degree. The red representation in the fig 3 shows the deflected shape of structure under the
action of wave. For wave induced loading the maximum deflection observed for joint 604L
which is highlighted in fig 3. Max deflection observed in joint at top bay of structure. Even
though maximum defelction due to wave load and self weight occurs in Y direction they are in
opposite directions. The one is in direction of interaction of wave and deflection due to self
weight is opposite to thatdirection.
26
Dynamic analysis
The following six mode shapes where observed from the dynamic analysis for the above
structure (fig 1).
27
Figure 2.5 (mode shape 2).
28
Figure 2.7 (mode shape 4).
29
Figure 2.9 (mode shape 6).
30
The deflected shapes or mode shapes and their respective frequency of vibration are show from
figure 4 to 9
Table 1 shows the results obtained from dynamic analysis of structure. It shows the mode
shapes and respective frequency of vibration for that mode shape. From the above figures and
Table 1 one can easily visualise possible vibrating or deflected shapes of the structure under
given loading conditions.
31
9 2.2 603L 2.54 601L 0.13 204L
10 2.54 601L 2.67 601L 0.2 302L
TABLE 2.2. Modes and Deflectiom
The table 2 gives the maximum deflection observed at different joints for different mode shapes
in X, Y and Z directions. From table 2 it is observed that the deflection along Z direction is our
least concern. Whereas a maximum deflection of 2.54 cm is observed for different modes at
different joints in X and Y directions. From the designers point of view this much of deflection
for this number of joints is not good while considering strength as well as serviceability of the
structure. The maximum deflection permitted at a joint in an offshore jacket platform depends
on several factors, including the design codes, environmental conditions, and operational
requirements. Typically, offshore structures are designed to limit deflections to ensure
structural integrity, maintain operational performance, and avoid excessive fatigue damage [1].
From a designers point of view smaller modes are of more concern as it will excite in smaller
excitation frequency.
Figure 11 gives the maximum values of the deflection foe different mode shapes from 1-10,
in X, Y and Z directions.
32
From API RP 2A the below mentioned are the provisions to be followed while considering the
deflections.
General Guidelines for Maximum Deflection in Offshore Structures (as per API RP 2A):
API does not specify a fixed universal deflection limit for joints but provides guidelines based
on serviceability, functionality, and operational criteria. The allowable deflections are generally
controlled by[1]:
1. Operational Requirements:
o For topside structures (e.g., decks, living quarters):
▪ 1/400 to 1/600 of the span length to avoid excessive vibration and
serviceability issues.
o For slender members like risers and conductors:
▪ Limited deflections to avoid misalignment and fatigue.
2. Environmental Load Considerations:
o Jacket structures exposed to wave and current forces should limit deflections to
prevent fatigue failure and misalignment of critical equipment.
o Displacements should generally not exceed H/400 to H/200, where H is the
height of the structure.
3. Fatigue and Structural Integrity:
o Excessive deflections can induce fatigue damage, so API recommends
performing fatigue analysis to ensure deflections do not compromise long-term
integrity.
4. Specific Guidelines in API RP 2A:
o Legs and Bracing: Typically, the overall platform deflection should not exceed
1/200 of the total height.
o Deck Deflections: Generally limited to 1/240 to 1/360 of the span length,
depending on functional requirements.
o Local Joint Displacements: Should be limited based on equipment tolerances
and alignment requirements.
Practical Values for Jacket Platforms:
• Vertical Deflection (Deck):
o Acceptable range: Span/400 to Span/600
• Horizontal Deflection (Under Environmental Loads):
33
o Acceptable range: H/400 to H/200
• Serviceability Considerations:
o Deflections should not impair equipment operation or human comfort.
From table 2 the joints 601L,603L and 304P are having maximum deflections which has to be
considered for further analysis or during construction.
From all figures from figure 4 to 10 the lower part of the structure (legs and braces) experiences
relatively less movement compared to the upper deck. This is typical for offshore platforms
where the jacket structure is anchored to the seabed. The upper portion of the platform seems
to show larger deflections, possibly indicating flexibility in the topside area due to higher
exposure to environmental forces. Joints 601L, 602L, and 603L frequently appear in both X
and Y directions, meaning these joints experience the highest sway movement. In the Z-
direction, joints 304L and 204L show the highest vertical deflection, potentially indicating
areas affected by dynamic forces such as wave impact or deck loads.
Mode-Specific Observations:
• Mode 1:
o Minimal deflection in the X and Z directions but significant lateral sway (2.54
cm) at joint 602L.
• Mode 2:
o Major deflection (2.54 cm) in the X-direction at joint 603L, indicating strong
longitudinal movement.
• Mode 5:
o Maximum lateral displacement at joint 601L, which could indicate local
flexibility or bending effects.
• Mode 7:
o Balanced large deflections in both X and Y directions, highlighting complex
structural movement.
• Mode 10:
o Notable deflections in both X and Y axes, showing coupled motion effects.
35
o MOMENT-Y exhibits extreme values, such as 68612.5 FT-KIP in MODE 10,
indicating a dominant bending moment likely due to wave or wind loads along
one axis.
o MOMENT-Z, while smaller in magnitude compared to other moments, can still
influence stability due to out-of-plane forces, especially in Mode 8.
CONCLUSION
Typical four-legged jacket is modeled in SACS. It is analysed for environmental and gravity
load conditions. The deflection for the structure and overturning moments are found. Different
mode shapes are generated and analysed for dynamic analysis of the structure. The identified
joint locations with high deflection should be reviewed for potential fatigue or stress
concentration. If specific modes have high deflections, they might contribute significantly to
the platform's dynamic response and should be included in fatigue analysis. Excessive lateral
36
deflection (Y-axis) in some modes might indicate the need for additional bracing or stiffness to
resist wave and wind forces. Vertical deflections (Z-axis) are relatively low, meaning the
platform has sufficient stiffness against gravity and buoyancy effects. Higher modes exhibit
higher frequency but lower time period for oscillation. This means even though the deflection
is more in these higher modes their oscillation or vibration will be executed for a very small
time period which means the specific excitation could die out faster than lower modes.
REFERENCES
1.American Petroleum Institute. (2005). Recommended practice for planning, designing and
constructing fixed offshore platforms: Working stress design (21st ed.). USA: American
Petroleum Institute. Retrieved from
https://books.google.com/books/about/Recommended_Practice_for_Planning_Design.html?i
d=_2DCGAAACAAJ
2. Faseela, A., & Jayalekshmi, R. (2015). In-place strength evaluation of jacket platforms and
optimization of bracing configurations. International Conference on Technological
Advancements in Structures and Construction, 15, 121–125. Retrieved from
http://www.ijrat.org/downloads/tasc15/TASC%2015-304.pdf
3. Kabir, S. (2007). An overview of design, analysis, construction and installation of offshore
petroleum platforms suitable for Cyprus oil/gas fields. GAU Journal of Social & Applied
Sciences, 2(4), 1–16. Retrieved from https://cemtelecoms.iqpc.co.uk/media/6514/786.pdf
4. Shehata, E. A., Elsayed, M. A. A., Aly, G. A. A., & Fayez, K. A. S. (2012). Nonlinear analysis
of offshore structures under wave loadings. Proceedings of the 15th World Conference on
Earthquake Engineering (15WCEE), Paper No. 3270, 1–10. Retrieved from
https://www.iitk.ac.in/nicee/wcee/article/WCEE2012_3270.pdf
37
EXERCISE 3
3.1 INTRODUCTION
ANSYS AQWA is an engineering toolset for the investigation of the effects of wave, wind and
current on floating and fixed offshore and marine structures, including spars, floating
production, storage, and offloading (FPSO) systems, semi-submersibles, tension leg platforms
(TLPs), ships, renewable energy systems, and breakwater design. AQWA Hydrodynamic
Diffraction provides an integrated environment for developing the primary hydro dynamic
parameters required for undertaking complex motions and response analyses. Three-
dimensional linear radiation and diffraction analysis may be undertaken with multiple bodies,
taking full account of hydrodynamic interaction effects that occur between bodies. While
primarily designed for floating structures, fixed bodies such as breakwaters or gravity-based
structures may be included in the models. AQWA Hydrodynamic Time Response provides
dynamic analysis capabilities for undertaking global performance assessment of floating
structures in the time domain.
1. Geometry and Meshing – Defining the cylindrical structure and discretizing it into
elements suitable for hydrodynamic analysis.
2. Wave Environment Setup – Specifying wave parameters such as wave height, period,
and direction to simulate realistic ocean conditions.
38
3.2 RESEARCH SIGNIFICANCE
Understanding the hydrodynamic behaviour of floating structures is crucial for ensuring their
safety, stability, and performance in marine environments. This analysis provides details about
how the structure responds to wave forces, which is essential for optimizing design parameters
such as stability, buoyancy, and motion characteristics. The results can be used to improve the
structural integrity and operational efficiency of offshore platforms, ships, and other marine
structures.
Model setup
The geometry of the structure was defined, including its mass properties, centre of gravity, and
moments of inertia. The model was set up in a metric unit system (kg, m, N). The structure
consists of a single part with a fully defined cylinder of 0.8 m diameter and 0.6 m depth. The
part is represented as a solid body with specific dimensions and properties. The total mass of
the structure is 180 kg, with the centre of gravity (CoG) located at (0.0 m, 0.0 m, -0.15
m) relative to the global coordinate system. This mass of 180kg is defined as a point mass at a
depth of -0.15m from the surface. The moments of inertia were defined to capture the structure's
rotational behaviour. For example: Ixx: 21 kg·m² , Iyy: 21 kg·m², Izz: 19 kg·m²
39
Environmental conditions.
The environmental conditions define the water properties and wave characteristics that the
structure will be exposed to. These conditions are essential for simulating realistic sea states.
The water depth was set to 100 meters and the density of water was set to 1025 kg/m³,
representing seawater. The acceleration due to gravity was set to 9.80665 m/s². The wave
directions ranged from -180° to 180° with intervals of 45°, covering all possible wave approach
angles. The wave frequencies ranged from 0.01562 Hz to 1.48231 Hz, capturing both low-
frequency (long-wavelength) and high-frequency (short-wavelength) waves. Sea mud
mudlayer depth of 2m is also defined.
Boundary Conditions.
Boundary conditions define how the structure interacts with its environment. In this analysis a
fixed point was defined at a depth of 100 meters to represent the seabed. This fixed point was
used to anchor the structure during the analysis. A connection point was defined at (0.0 m, 0.0
m, -0.45 m) to represent the attachment point for cables or mooring lines to the structure. A
linear cable with a stiffness of 60 N/m was defined to connect the structure to the fixed point
at the seabed. The unstretched length of the cable was set to 0.0 meters.
40
Fig 3.3. Boundary Conditions.
Mesh Generation
The structure was discretized into a finite element mesh to enable numerical calculations. The
default element size of 0.09878 meters was used, ensuring a balance between accuracy and
computational efficiency. The mesh was checked for quality, with a total of 322 nodes and 321
elements generated. Automatic waterline nodes were created to capture the interaction between
the structure and the water surface.
41
Fig 3.4. Mesh Generation.
The hydrostatic and stability analysis was performed to evaluate the structure's buoyancy and
stability in calm water. Here, the actual volumetric displacement is the fluid displacement of
the structure calculated by Aqwa, based on the diffracting surface mesh and any line bodies.
The equivalent volumetric displacement is the total mass of the structure divided by the water
density. The hydrostatic details observed were as follows, the actual displaced volume
was 0.2229632 m³, the metacentric height in the X-direction was 1.25396e-2 meters, indicating
the structure's initial stability. The Centre of Buoyancy was located at (7.3493e-5 m, -5.4027e-
9 m, -0.2249993 m) with respect to the global reference.
The hydrodynamic analysis settings define the type of analysis to be performed and the
parameters to be calculate. The analysis type was set to Hydrodynamic Diffraction/Radiation,
which calculates the wave forces and motion response of the structure. The wave grid
resolution was set to Standard (81 x 51), ensuring accurate representation of wave forces. The
42
analysis generated output files for hydrodynamic forces, RAOs, and hydrostatic properties and
also parallel processing was enabled to speed up the computation.
The heave stiffness represents the structure's resistance to vertical motion (up and down). A
high heave stiffness value indicates that the structure will experience significant restoring
43
forces when displaced vertically. In this case, the heave stiffness of 4979.5938 N/m is relatively
high, suggesting that the structure will resist heave motion effectively.
This is important for maintaining stability in waves, as excessive heave motion can lead to
operational issues or structural damage.
44
The roll stiffness represents the structure's resistance to rolling motion (side-to-side tilting ie,
rotation about X-axis). A low roll stiffness value indicates that the structure is less resistant to
rolling motions. The roll stiffness of 1.2304e-2 N/° is relatively low, suggesting that the
structure may experience significant rolling motions in waves. This could be a concern in rough
sea conditions, where large rolling motions can affect stability and safety.
The pitch stiffness represents the structure's resistance to pitching motion (front-to-back tilting
ie, rotation about y axis). A moderate pitch stiffness value indicates that the structure will
experience some resistance to pitching motions. The pitch stiffness of 0.4946387 N·m/° is
moderate, suggesting that the structure will resist pitching motions to some extent. However,
like roll stiffness, the pitch stiffness may not be sufficient to prevent large pitching motions in
rough seas.
The high heave stiffness (4979.5938 N/m) is a positive finding, as it indicates that the structure
will resist vertical motions effectively. This is important for maintaining stability and
preventing excessive heave in waves. However, the high heave stiffness also means that the
structure may experience significant vertical forces in waves, which could lead to fatigue or
structural damage over time.
The low roll stiffness (1.2304e-2 N/°) and moderate pitch stiffness (0.4946387 N·m/°) suggest
that the structure is less resistant to rolling and pitching motions. This could be a concern in
rough sea conditions, where large rolling and pitching motions can affect stability, safety, and
operational performance. The low roll stiffness, in particular, indicates that the structure may
be prone to rolling motions.
The hydrostatic stiffness values highlights the need for design improvements to enhance the
structure's resistance to rolling and pitching motions. For example, increasing the water plane
area or adding stabilizing fins could improve roll and pitch stiffness.
Hydrostatic Displacements
The hydrostatic displacement properties are critical for understanding the buoyancy, stability,
and equilibrium of the floating structure. These properties were calculated as part of the
hydrostatic analysis, and the results provide valuable insights into the structure's behavior when
subjected to hydrostatic forces. From fig 6 the following observations and results can be
discussed,
45
The actual displaced volume of the structure is 0.2229632 m³. The displaced volume represents
the volume of water displaced by the structure when it is floating. This value is directly related
to the buoyant force acting on the structure, which balances the structure's weight. A displaced
volume of 0.2229632 m³ indicates that the structure is relatively small, which is consistent with
the total mass of 180 kg. This value is crucial for ensuring that the structure remains afloat and
maintains its equilibrium. Also the equivalent volumetric displacement is 0.1756098 m³. The
equivalent volumetric displacement is a theoretical value that represents the volume of water
displaced if the structure were perfectly shaped to minimize drag and maximize buoyancy. The
difference between the actual displaced volume (0.2229632 m³) and the equivalent volumetric
displacement (0.1756098 m³) suggests that the structure's shape is not perfectly optimized for
buoyancy. This could be due to the presence of appendages, irregular geometry, or other design
features that increase the displaced volume.
The Centre of Buoyancy is the point where the buoyant force acts on the structure. Its location
is critical for determining the structure's stability. The centre of buoyancy (CoB) is located at (
7.3493e-5, -5.4027e-9, -0.2249993) with respect to the reference. In this case, the CoB is
located slightly below the center of gravity (CoG) at Z = -0.15 m. This vertical separation
between the CoB and CoG is essential for ensuring positive stability. The structure will tend to
return to its equilibrium position if tilted, as the buoyant force creates a restoring moment.
From fig 6 other parameters observed was out of balance forces and moments. The out-of-
balance force and moment indicate any resulting or residual forces or moments acting on the
structure after considering for buoyancy and weight. The small values of the out-of-balance
force (0.2696517 N) and moment (2.8702e-9 N·m) suggest that the structure is in near-perfect
hydrostatic equilibrium. This is a positive finding, as it indicates that the structure is well-
balanced and stable in calm water.
The other important parameter to be discussed from the hydrostatic result is the metacentric
height GM, which has great influence on stability of a floating structure. The metacentric height
in the X-direction (GMX) is 1.25396e-2 meters. The metacentric height (GM) is a measure of
the structure's initial stability. It represents the distance between the centre of gravity (CoG)
and the metacentre (the point where the buoyant force acts when the structure is tilted) (fig 7).
46
Figure 3.7. Stability of Floating Body
A positive GM value (1.25396e-2 meters) indicates that the structure is stable. If the structure
is tilted, the buoyant force will create a restoring moment that returns the structure to its upright
position. However, the relatively small GM value suggests that the structure has limited
stability. This could be a concern in rough sea conditions, where larger waves might induce
significant rolling or pitching motions.
Fig
The cut water plane area represents the area of the structure that is in contact with the water
surface. A larger water plane area generally contributes to greater stability for floating objects
whereas for submerging the structure need to have deep draft with lesser waterplane area. The
47
centre of flotation is the centroid of the water plane area. Its location is important for
understanding how the structure will respond to tilting or heeling. The principal second
moments of area (Ixx and Iyy) indicate the distribution of the water plane area about the
principal axes. These values are used to calculate the structure's resistance to rolling and
pitching motions.
Frome fig 6 the small angle stability parameters conveys that a BG of 7.4999e-2 meter and
restoring moments (MX/MY) of 0.4904677 N·m/° (MX) and 0.494702 N·m/° (MY). The
distance between the center of gravity (CoG) and the center of buoyancy (CoB) is 7.4999e-2
meters. This distance known as BG, is an important parameter for assessing stability and the
restoring moments (MX and MY) represent the structure's ability to resist rolling and pitching
motions. The values of 0.4904677 N·m/° and 0.494702 N·m/° indicate that the structure might
have moderate resistance to tilting.
Parameters vs Frequency
The Response Amplitude Operators (RAOs) were analysed to understand the structure's motion
response in the global Z-direction (heave) under various wave frequencies. The RAOs
represent the ratio of the structure's motion amplitude to the wave amplitude, providing details
into how the structure responds to wave forces.
The RAOs show a clear dependence on wave frequency. The structure's heave response
increases as the wave frequency increases, peaking at 0.63334 Hz, and then decreases as the
frequency increases further. The maximum RAO value at 0.63334 Hz indicates resonance,
where the structure's natural frequency could match the wave frequency, leading to amplified
motion. This is a critical finding for operational safety, as resonance can lead to excessive
motion and potential structural damage. Whereas considering this structure response in view
49
of wave energy converters peak RAO at 0.63334 Hz that it is most efficient at capturing energy
from waves near this frequency and performance is limited at higher frequencies, where the
RAO values are minimum. To maximize energy capture, the structure should be optimized for
operation in wave conditions near its resonant frequency.
From fig 6 the heave displacement analysis focuses on the vertical motion of the structure's
centre of gravity (CoG) relative to the origin of the fixed reference axes (FRA). The heave
displacement is a critical parameter for assessing the structure's stability and comfort in induced
conditions. At the resonant frequency of 0.63334 Hz, the heave displacement reaches its
maximum value of 3.15003 m/m (Table 1). This indicates that the structure experiences
significant vertical motion at this frequency, which could impact stability and performance. At
higher frequencies, such as 1.48231 Hz, the heave displacement is minimal (8.19248e-4 m/m).
This suggests that the structure is relatively stable and less affected by high-frequency waves.
At low frequencies, such as 0.01562 Hz, the heave displacement is also minimal. This indicates
that the structure is less responsive to very low-frequency waves, which is typical for floating
structures (Table 1).
CONCLUSION
The hydrostatic displacement properties provide a combined understanding of the structure's
buoyancy, stability, and equilibrium in calm water. Key findings include:
1. The structure has a displaced volume of 0.2229632 m³, which ensures sufficient
buoyancy to support its weight.
50
2. The centre of buoyancy (CoB) is located slightly below the centre of gravity (CoG),
contributing to positive stability.
3. The metacentric height (GMX) of 1.25396e-2 meters indicates that the structure is
stable but has limited resistance to tilting.
4. The out-of-balance forces and moments are minimum, confirming that the structure is
in near a stable hydrostatic equilibrium.
5. The structure has a high heave stiffness (4979.5938 N/m), indicating strong resistance
to vertical motions. This is beneficial for maintaining stability in waves but may lead
to significant vertical forces in rough seas.
6. The low roll stiffness (1.2304e-2 N/°) suggests that the structure is prone to rolling
motions, which could affect stability and safety in rough sea conditions.
The analysis of RAOs and heave displacement provides critical insights into the structure's
hydrodynamic behaviour. The structure exhibits significant heave motion at its resonant
frequency (0.63334 Hz), which must be considered in operational planning to avoid excessive
motion. At higher frequencies, the structure remains stable, with minimal heave displacement.
These observations are essential for optimizing the design and ensuring the safe operation of
the structure in marine environments. Whereas considering the structure for a wave energy
converter, a good heave response is essential for better energy conversion processes. Hence
maintaining this frequency level for higher amplitude also considering structural failure for
these responses could effectively use these data for energy conversion models.
REFERENCES
1. Ghadimi, P., Paselar Bandari, H., & Bankhshandeh Rostami, A. (2012). Determination
of the heave and pitch motions of a floating cylinder by analytical solution of its
diffraction problem and examination of the effects of geometric parameters on its
dynamics in regular waves. International Journal of Applied Mathematical Research,
1(4). https://doi.org/10.14419/ijamr.v1i4.396
2. Riyanto, R. D., & Rahmawati, S. (Year). Hydrostatic stiffness as displacement
boundary condition of floating cylindrical structural analysis in waves.https://doi.org/
10.5220/0010058501310137
51
3. Zheng, S., Zhang, J., Lei, Z., Yan, X., Gao, S., Hu, W., & Cheng, K. (2024). *Structure
design and static analysis of a cylindrical wave buoy*. *Proceedings of SPIE, 12981,
Ninth International Symposium on Sensors, Mechatronics, and Automation System
(ISSMAS 2023)*, 1298136. https://doi.org/10.1117/12.3014824
52
EXERCISE 4
1. Surge Motion (forward and backward motion): This is the linear motion of the ship
along its longitudinal axis (along global X-axis). Resistance in surge is primarily due
to the hydrodynamic drag acting on the hull as it moves forward or backward through
the water.
2. Sway Motion (sideways motion): This is the linear motion of the ship along its
transverse axis (along global Y-axis). Resistance in sway is influenced by the lateral
forces acting on the hull, which can be significant during manoeuvrs or in crosswinds
and currents.
3. Yaw Motion (Rotational Motion about Z-Axis): Yaw motion is the rotational
movement of a ship around its vertical (z) axis. It determines the ship’s ability to turn,
manoeuvring, and maintain course. Yawing occurs due to external forces like waves,
wind, rudder deflection, or asymmetric hull resistance. A yaw moment (M_z) is
generated, causing the bow and stern to rotate in opposite directions.
By using CFD simulations in OpenFOAM, engineers can accurately predict these resistance
forces and develop better, more efficient ship designs.
53
Figure 4.1. Ship motion and axis
1. Design Optimization: Accurate prediction of ship resistance in surge, sway, and rolling
motions allows for the optimization of hull forms and appendages to minimize
resistance, leading to more fuel-efficient ships.
2. Stability and Safety: Understanding the resistance in these motions is important for
assessing the stability and safety of ships, especially in rough seas where large
amplitudes of motion can occur.
3. Operational Efficiency: Ships with lower resistance in various motions can operate
more efficiently, leading to cost savings and improved performance in commercial and
military applications.
54
4. Manoeuvrability: Resistance in sway and rolling motions directly impacts the ship's
manoeuvrability. Accurate simulations can help in designing ships with better handling
characteristics.
Model setup
A ship model of length (𝐿𝑝𝑝 ) 5.72 m is considered within a domain. Ship is considered as under
a static drift of β= 100 . Static drift refers to a condition where a ship moves forward while
maintaining a constant side-slip angle (drift angle, β\betaβ) due to external forces like wind,
waves, or manoeuvring. This leads to lateral (sway) resistance and affects the ship’s
hydrodynamic performance. It helps in evaluating how well a ship maintains course without
excessive rudder corrections. Here a constant inlet velocity of 0.819 m/s is given for analysis
in the direction of ship motion (Global X direction). As the ship has a drift angle with respect
to this direction, forces along y direction and concerns about sway motion is coming into
picture. From figure 2 we can depict the direction of motion of ship with a velocity U and drift
angle of β degrees. But in this analysis an equivalent different approach is employed, that is
instead of considering the ship moving, the ship is considered to be at rest and the water id
flowing against the ship at rest with a velocity of U. So, the direction of U will be opposite to
the U represented un figure 2. So whatever U we are processing for it would be entered as -U
in analysis domain. Also as mentioned earlier due to the drift angle from the X-axis or direction
of motion, motion and forces along Y-axis (sway) is also considered. Which is velocity V
should be also occupied for the analysis. The inputs for Velocity along X-axis (U) and velocity
along Y-axis can be represented as
From equation 2 it is to mention that V will be negative when β is negative and will be positive
if β is positive.
Governing equations
In OpenFoam, the Navier-Stokes equations are solved numerically to model fluid flow, which
then needed for the analysis of ship resistance and motion simulations. The equation governs
the motion of incompressible and compressible fluids and are fundamental to Computational
Fluid Dynamics (CFD). Basically Navier-Stokes equation is developed from the conservation
of momentum which is one among the three basic conservation laws of CFD [1]. The equation
is as follows
𝜕𝑈 1
+ (𝑈. ∇)𝑈 = − 𝜌 ∇𝑝 + 𝜗∇2 𝑈 + 𝐹 eq (3)
𝜕𝑡
Where:
P=Pressure
𝜌= density of fluid
56
𝜗= Kinematic viscosity
F= External forces
Along with this conservation of mass is applied and OpenFoam discretizes these equations
using Finite Volume Methods then solves them iteratively. Here for ship resistance problem
OpenFoam uses Volume of Fluid (VOF) method (interFoam), which captures the interaction
between the ship hull and the free surface.
In OpenFOAM, the interFoam solver is commonly used for VOF simulations. The VOF
method captures the interaction between the ship hull and the free surface. The process
involves:
1. Mesh Generation: Creating a high-quality mesh around the ship hull using
snappyHexMesh. The shape, size and aspect ratio of elements formed under mesh
have great importance in accuracy of the results generated. The more the elements
the more will be the accuracy and computational time. The mesh generation is
controlled by a file named blockMeshDict, which is located in the system directory.
Here a total of 88*54*28, which is equals to 59136 cells are generated.
2. Boundary Conditions: The boundary conditions involve specifying domain
boundary and size. Implementing inlet velocity and wave profile. It also has to
imply with turbulence flow prediction and no slip conditions for the ship hull for
this specific problem.
3. Post-Processing: Post-Processing includes extracting pressure distribution and
shear stress thus calculating drag forces and hull resistance.
The significance of each of the forces and moments are discussed below.
1. Pressure forces: Pressure forces arise due to the difference in pressure distribution on
the ship’s hull as it moves through water. These forces influence resistance,
manoeuvrability and stability. Pressure forces are generated by fluid flow around the
hull and are influenced by wave formation at the bow and stern which affects drag,
Flow separation and vortices which affects side forces and yaw moments and buoyancy
and hydrostatic pressure differences which affects vertical stability. The pressure forces
can be further break own into force along X and Y direction (which is our only concern).
Longitudinal Pressure Force (Fp_X) is the pressure-induced component of total
resistance along X direction. These are generated by high pressure at the bow and low
pressure at the stern which contributes to wave-making resistance. They are higher at
58
higher speeds or drift angles due to increased wave height and flow separation. Second
one is Transverse Pressure Force (Fp_y) or Side Force. This force is responsible for
lateral drift when the ship is at an angle which are caused by asymmetric pressure
distribution on the hull due to drift. It is important for course stability and
manoeuvrability. They increase significantly with higher drift angles.
2. Viscous Force: Viscous forces arise due to the shear stress between the ship’s hull and
the surrounding water. These forces are caused by fluid viscosity and contribute
significantly to total resistance. Viscous forces are generated by friction between the
water and the hull surface which creates shear stress. Here also we are concerned only
about its X and Y components. Longitudinal Viscous Force (Fv_x) or Frictional
Resistance (Drag) is the main contributor to total ship resistance which is caused by
skin friction along the hull due to viscosity. They dominate at low speeds but remains
significant at high speeds. Transverse Viscous Force (Fv_y) or Side Force which acts
against lateral motion of the ship. Thes are affected by eddy formation and shear stress
asymmetry at drift angles. Significant for manoeuvring studies during turns.
3. Yaw Moment (M_Z) or Rotation About Z-Axis: Yaw motion is the left-right turning
tendency of the ship and they have a great important for manoeuvrability and course
stability. These moments are caused by lateral force asymmetry due to drift angle.
𝑅𝑇 = 𝑅𝑝 + 𝑅𝑉 + 𝑅𝑝𝑜 eq(4)
59
and 𝑅𝑝𝑜 is the porous resistance which is zero in this case.
Therefor the resistance against the X, Y and Z direction of motion of ship can be found out by
adding the respective forces along that direction. Here porous forces and moments are zero
because no porous zone is defined in the simulation. If the porous region is not included in the
OpenFOAM setup (e.g., missing fvOptions for porosity modelling), then 𝐹𝑝𝑜𝑟 and 𝑀𝑝𝑜𝑟 will
be zero. Here a representation of ship model is simulated then the hydrodynamic forces and
moments are depicted in terms of similarity rules and dimensionless parameter for the
prototype.
𝑋
𝑋′ = eq(5)
0.5𝜌𝑈𝛼2 𝑇𝑚 𝐿𝑝𝑝
𝑌
𝑌′ = eq(6)
0.5𝜌𝑈𝛼2 𝑇𝑚 𝐿𝑝𝑝
𝑁
𝑁′ = eq(7)
0.5𝜌𝑈𝛼2 𝑇𝑚 𝐿2𝑝𝑝
Where,
𝑋 ′ and 𝑌 ′ are the non-dimensional longitudinal (surge) and transverse (sway) force
coefficients, respectively.
In this post-processing, the data from figure 3 is analysed and forces along X direction
(pressure, viscous and porous) is sum up in order to get the X terms in equation 5. Similarly
60
forces along Y direction and moments along z direction is sum up for substituting Y and Z in
equation 6 and 7 respectively. After plotting graph for forces (Surge and Sway) and Yaw vs
time, a great variation was observed at the start of time steps as shown in figure 4.
61
Figure 4.4. forces and yaw moment vs time (plotted using python).
These variations are neglected and averaging for the same ranged values imply
𝑋 ′ = -0.09312
𝑌 ′ = 0.06754
𝑁 ′ = -0.004
Ship resistance is a combination of viscous drag, pressure forces, and wave-making resistance.
The non-dimensional coefficients you provided (X′, Y′, N′) help in understanding the
hydrodynamic forces acting on the ship, particularly under static drift conditions (where the
ship is at an angle to the incoming flow). X ′ is directly related to the total resistance of the
62
ship. Since 𝑋′ is negative (−0.09312), it confirms that the resistance force is acting opposite to
the ship’s forward motion. From this non dimensional force the actual resistance for prototype
can be found by considering representative similarity rules. Y′ represents the lateral force on
the ship due to drift angle (β). A nonzero Y′ indicates side forces, which contribute to increased
total resistance due to the additional cross-flow drag and unstable motion, requiring correction
via rudders or thrusters. Based on this X ′ and Y′ we could determine how much thrust should
be implied through propellers to overcome this resistance thus for the smooth motion of ship.
The N′ represents the moment that causes the ship to turn. A nonzero yaw moment means the
ship is experiencing asymmetric flow, leading to additional resistance as more force is required
to correct its path and increased rudder forces to maintain direction, adding to drag.
CONCLUSION
1. X′=−0.09312 indicates resistance is acting against forward motion. A more negative
X′ suggests significant drag, which can impact fuel efficiency and ship speed.
2. Y′=0.06754 imply a positive sway force suggests side forces due to drift angle. This
could lead to increased cross-flow drag and affect manoeuvrability.
3. N′=−0.0044 imply a small negative yaw moment suggests a slight tendency to turn,
meaning some asymmetry in the flow but not a major stability concern.
The ship experiences noticeable resistance, indicating potential drag sources (viscous, wave-
making). Sway forces suggest possible course deviation, important for manoeuvring in currents
or side winds. Yawing tendency is low, meaning the ship is relatively stable but may require
rudder corrections [2].
For Fuel Efficiency reducing X′ (resistance) through hull shape optimization can improve
energy efficiency. For manoeuvrability controlling sway forces (Y′) and yaw moments (N′) is
important for course stability, especially in crosswinds or currents. This analysis provides a
foundation for improving ship design, optimizing performance, and reducing energy
consumption in real-world operations.
REFERENCES
1. Islam, H. A CFD study of a ship moving with constant drift angle in calm water and
waves.
2. Zhang, Y., Díaz-Ojeda, H. R., Windén, B., Hudson, D., & Turnock, S. (2024). A
numerical study of drift angle effect on hydrodynamic performance of a fully appended
63
container ship in head waves. Ocean Engineering, 313, 119343.
https://doi.org/10.1016/j.oceaneng.2024.11934.
64
EXERCISE 5
5.1 INTRODUCTION
Fluid flow around a circular cylinder is a problem of extreme relevance to both research and
engineering practice. It is essential for the design of structures like heat exchangers, cooling
towers, nuclear reactors, bridge piers, and subsea pipelines, all of which are cylindrical in
nature. Flow conditions in most engineering applications are found in the subcritical
Reynolds number range.
As the fluid interacts with the cylinder, vortex shedding and boundary layer separation
happen, especially at high and moderate Reynolds numbers. These flow patterns can cause
unsteady forces, resulting in unwanted vibrations that can lead to fatigue failure or structural
damage. Thus, the study of the impact of vortex shedding on cylindrical structures is an
important research field in fluid dynamics.
The nature of flow around a cylinder is a function of the Reynolds number (Re), which is the
ratio of inertial forces to viscous forces. With an increase in the Reynolds number, the
behaviour of the flow goes through various regimes, each with unique characteristics figure
[1].
• Re < 1: At low Reynolds numbers, the flow is fully laminar with no separation.
Viscous forces prevail, leading to smooth, symmetrical flow around the cylinder, with
no wake formation.
• 4 < Re < 40: When the Reynolds number increases somewhat, the flow starts to
separate in the rear of the cylinder and creates a symmetric, stable pair of vortices in
the near wake. These vortices are attached to the cylinder surface.
• 40 < Re < 100: The wake region gets destabilized, and periodic vortex shedding
occurs. This results in the development of the Kármán vortex street, in which
65
alternating vortices are shed from the cylinder owing to oscillatory flow and pressure
gradients.
• 100 < Re < 200: Vortex shedding becomes more significant, with intense periodic
oscillations in the wake area.
• 200 < Re < 400: The wake flow becomes more unstable, with random disturbances
appearing, indicating the onset of laminar to turbulent flow.
• 400 < Re < 2.5 × 10⁵: The flow becomes subcritical, and boundary layer separation is
at about 80° from the cylinder front. This results in a region of low-pressure wake
behind the cylinder, considerably affecting drag forces.
• 2.5 × 10⁵ < Re < 3.5 × 10⁵: In this critical flow regime, the turbulent and unsteady
boundary layer separation occurs, which results in a significant drag reduction of
almost 70%.
• 3.5 × 10⁵ < Re < 3 × 10⁶: The separated flow region is turbulent and reattaches at the
back of the cylinder, which results in a rise in drag.
• Re > 3 × 10⁶: The flow is completely turbulent, with separation delayed along the
surface of the cylinder (around 140° from the front), further enhancing drag.
67
• Through the study of shedding frequency and aerodynamic forces, this research
assists in the design of flow control devices, including vortex generators and passive
flow control devices, to reduce flow-induced oscillations.
Governing Equations
The flow is governed by the 2-D incompressible Navier-Stokes equations, which consist of:
𝜕𝑣 𝜕𝑣 𝜕𝑣 1 𝜕𝑃 𝜕2 𝑣 𝜕2 𝑣
+𝑢 +𝑣 =− + 𝜗( 2
+ )…………………….eq (4)
𝜕𝑡 𝜕𝑥 𝜕𝑦 𝜌 𝜕𝑦 𝜕𝑥 𝜕𝑦 2
Where,
68
• P = pressure
• 𝜗 = kinematic viscosity (1.7894 × 10−5 𝑘𝑔/𝑚 ⋅ 𝑠)
• 𝑥, 𝑦 = spatial coordinates
Explanation:
𝜕𝑢 𝜕𝑣
• and is the Unsteady term which represents changes in velocity with time
𝜕𝑡 𝜕𝑡
(transient effects).
𝜕𝑢 𝜕𝑢 𝜕𝑣 𝜕𝑣
• 𝑢 +𝑣 and 𝑢 +𝑣 is the convective term which represents the transport
𝜕𝑥 𝜕𝑦 𝜕𝑥 𝜕𝑦
𝜕2 𝑢 𝜕2 𝑢 𝜕2 𝑣 𝜕2 𝑣
• 𝜗( 2
+ ) and 𝜗(𝜕𝑥 2 + 𝜕𝑦2 ) is the viscous diffusion term which represents
𝜕𝑥 𝜕𝑦 2
For Re = 250, the flow is in the laminar unsteady regime, where vortex shedding occurs,
forming a periodic wake.
4. The drag coefficient (Cd) and lift coefficient (Cl) are defined as:
𝐹𝑑
𝐶𝑑 =
1 2
2 𝜌𝑈 𝐴
𝐹𝑙
𝐶𝑙 =
1 2
2 𝜌𝑈 𝐴
69
Where,
• 𝐴 = 𝐷 ⋅ 𝐿 projected area of the cylinder
• 𝑈 =freestream velocity
𝑓𝐷
𝑆𝑡 =
𝑈
where:
• 𝐷 = cylinder diameter.
• 𝑈 = freestream velocity.
Model setup
This section provides a detailed model setup for simulating flow past a 2D circular cylinder at
Reynolds number 250 using ANSYS Fluent. A rectangular domain is created to ensure that the
boundaries do not affect the flow around the cylinder. The dimensions are chosen based on
standard CFD best practices for flow past bluff bodies.
70
Figure 5.2 Problem model
Domain Size:
• Length: 300 mm
• Height: 150 mm
Cylinder Specifications:
• Diameter: 10 mm
• Position: Placed at 1/3rd of the domain length from the inlet to allow sufficient wake
development.
Boundary Conditions.
Proper boundary conditions are set to simulate external flow over a circular cylinder
accurately.
71
Pressure Outlet (0 Pa gauge pressure)
Cylinder Surface No-Slip Wall (Zero velocity)
Top & Bottom Symmetric (Slip condition)
𝜌𝑈𝐷 1.225∗𝑈∗0.01
𝑅𝑒 = = ; 𝑈 = 0.365 𝑚/𝑠
µ 1.7894e−05
Mesh Generation
A structured quadrilateral mesh with inflation layers is used to capture boundary layer effects
and wake formation accurately.
• Number of Layers: 20
72
• Maximum Thickness: 0.003 m
The cylinder is positioned at 1/3rd of the domain length from the inlet to provide sufficient
space for both flow acceleration upstream and wake development downstream.
73
Solver Setup
The pressure-based solver is used since the flow is incompressible. The transient solver
captures the unsteady vortex shedding behind the cylinder.
𝑘𝑔 𝑘𝑔
• Material: Air (𝐷𝑒𝑛𝑠𝑖𝑡𝑦 = 1.225 𝑚3 , 𝑉𝑖𝑠𝑐𝑜𝑠𝑖𝑡𝑦 = 1.7894 × 10−5 𝑚−𝑠)
74
Figure 5.6. Setup.
75
Figure 5.8. Drag Coefficient
Even though lift (Cl) and drag (Cd) are coefficients, they still change over time in this
simulation because of the unsteady nature of vortex shedding at Re = 250. The fluctuations in
these coefficients are a direct result of the periodic flow behaviour around the cylinder.
At Re = 250, the flow around the circular cylinder is in the laminar unsteady regime, meaning
that the flow is not steady but continuously changing. This happens because of vortex shedding,
where alternating vortices are released from the top and bottom of the cylinder, forming a
Kármán vortex street. The Kármán Vortex Street is a repeating pattern of swirling vortices
formed behind a bluff body (such as a circular cylinder) when fluid flows past it at a moderate
Reynolds number.
76
Considering the bluff body, that induced with more drag than lift. The figure 7 represents the
drag coefficient (Cd) variation over time for a bluff body (circular cylinder) flow simulation.
The general behaviour of Cd in the plot can be divided into three key phases:
1. Initial Transient Phase (0 - 0.2 s): A sharp peak in Cd followed by a rapid drop. This is
because at the start, the flow is developing, and the fluid suddenly interacts with the
cylinder, creating a high pressure drag at the front. As the boundary layer starts forming,
separation occurs, leading to a reduction in Cd. This is a common behaviour in unsteady
flow simulations where the flow takes time to stabilize.
2. Stabilization & Mean Cd Plateau (0.2 - 0.8 s) : Cd settles into a more stable region (~1.1 -
1.3). This is due to the flow fully develops and aperiodic vortex shedding cycle starts. Cd
stabilizes around its expected average value, which is consistent with experimental and
numerical studies for Re ≈ 250. The drag force is now primarily due to the combination of
pressure drag and shear drag.
3. Periodic Oscillations in Cd (0.8 - 1.6 s): Cd starts oscillating regularly in a sinusoidal
pattern. The main reason for this can be considered as vortex shedding from the cylinder
alternates between the top and bottom, causing fluctuations in the wake pressure. These
alternating low-pressure zones cause periodic variations in the drag force. The Strouhal
frequency (vortex shedding frequency) is responsible for this periodic behaviour and this
oscillation is characteristic of unsteady wake flow (Kármán vortex street), which dominates
at Re = 250.
Considering the frequency, the figure 8 is a frequency spectrum obtained from the flow analysis
around a bluff body. The x-axis represents frequency (in Hz), and the y-axis represents
magnitude. The sharp peak observed at around 7-10 Hz indicates the dominant vortex shedding
frequency from the bluff body. This peak corresponds to the vortex shedding phenomenon
caused by alternating vortices being released from the object. Using these values Strouhal
number can be found out by
𝑓𝑠 𝐷 7.31∗0.01
𝑆𝑡 = = ; 𝑆𝑡 = 0.2
𝑈 0.365
77
Figure 5.10. Strouhal Number vs Reynolds Number plot.
From figure 9, for a Reynolds number around Re = 250, the Strouhal number is approximately
0.2 (highlighted in the figure). This confirms that the vortex shedding frequency follows a well-
established trend seen in experiments. The peak in the frequency spectrum aligns with the
expected vortex shedding frequency based on St ≈ 0.2 thus implies the validation for this
problem. For low Reynolds numbers (Re ≈ 250), the shedding process is stable and periodic.
As Reynolds number increases, the Strouhal number remains nearly constant at 0.2 before
showing deviations at very high Re due to turbulence effects.
78
As the floe stabilizes at after 0.8 seconds indicated from figure 6 and 7. It is better to consider
velocity and pressure profiles after this time. Considering these results at 0.75 and 1.37 seconds
implies the following results.
The colour scale indicates velocity magnitude, with red representing the highest velocity and
blue representing the lowest. At 0.75 seconds, the flow has started forming a wake region
behind the cylinder, but the vortex shedding is still in its early stages. The velocity behind the
cylinder is lower, as shown by the blue regions, indicating a recirculation zone. The vortex
79
structures are developing asymmetrically, showing the onset of vortex shedding and the wake
is shorter, meaning the flow is still transitioning towards a fully periodic state.
At 1.37 seconds, the wake has fully developed into a periodic Kármán vortex street. The
alternating vortices shed downstream, forming a repeating pattern. Another observation found
was the wake has elongated, showing that the shedding process is fully established and
periodic. The velocity field shows alternating high- and low-velocity regions corresponding to
vortices detaching from the cylinder. The vortex street is characterized by oscillatory lift and
drag forces on the cylinder. The wake transitions from an initial unstable state to a periodic
vortex shedding pattern. It is observed that the highest velocities are around the edges of the
cylinder and in the detached vortices, while the lowest velocities appear in the recirculation
region behind the cylinder.
To infer about Vortex Induced Vibration, Vortex-Induced Vibrations (VIV) occur when
alternating vortex shedding exerts periodic forces on a bluff body, causing it to oscillate. The
second velocity contour (1.37s) shows a well-developed Kármán vortex street, indicating a
stable shedding pattern. Alternating vortices generate fluctuating lift and drag forces on the
cylinder, which can lead to structural oscillations. The previously shown drag coefficient graph
(figure7) displays oscillations, which suggest a periodic force acting on the body. The lift
coefficient plotted was (figure 6), likely shows sinusoidal variation which is the characteristic
of VIV. Frequency spectrum graph (FFT analysis) shows a strong peak around 8-10 Hz,
corresponding to the dominant vortex shedding frequency. This frequency is critical in VIV
analysis because if the shedding frequency (Strouhal frequency) matches the natural frequency
of the body, resonance can occur, amplifying the oscillations. A single dominant peak suggests
a lock-in phenomenon, where the vortex shedding synchronizes with the body’s oscillations,
leading to large-amplitude vibrations. The Strouhal number (St = 0.2) is consistent with vortex-
induced forces, meaning VIV risk is significant in your case if the structure is free to move.
Figure 12 is the pressure contour visualization around a circular bluff body provides key
insights into the flow dynamics and vortex shedding behaviour. The main observations are
mentioned below
1. High-Pressure Region at the Front (Stagnation Point): The red region at the front of the
cylinder indicates a high-pressure stagnation point, where the incoming flow directly
80
impacts the surface. This is expected because the flow velocity at the stagnation point is
zero (due to no-slip condition), leading to maximum static pressure.
2. Low-Pressure Regions in the Wake (Vortex Shedding): The alternating blue regions behind
the cylinder represent low-pressure vortices, formed due to the periodic vortex shedding.
These low-pressure zones cause fluctuating lift forces, which are responsible for Vortex-
Induced Vibrations (VIV). The main observation of Kármán vortex street is clearly visible,
showing a regular pattern of alternating pressure zones.
The pressure distribution confirms vortex shedding and wake formation, aligning with previous
drag and frequency results and the alternating pressure zones indicate periodic shedding, which
is crucial for VIV analysis.
This streamline contour plot (figure 13) provides a detailed visualization of flow separation,
vortex shedding, and wake dynamics around a 2-D circular cylinder. It is clear that
streamlines curve sharply around the cylinder, signifying flow separation from the surface. A
clearly defined recirculation region occurs behind the cylinder, with streamlines creating closed
loops. This recirculating flow is typical of vortex formation from the adverse pressure gradient
in the wake. The alternating whirling zones in the wake indicate vortex shedding, a
phenomenon of Vortex-Induced Vibrations (VIV). This indicates the existence of a Kármán
vortex street, in which vortices develop on opposite sides, imposing oscillatory lift forces on
the cylinder. This indicates the existence of a Kármán vortex street, in which vortices develop
81
on opposite sides, imposing oscillatory lift and drag forces on the cylinder. From the
calculations, the frequency of shedding can be attributed to the Strouhal number (St = 0.2 in
fig 9).
The color gradient in the streamlines shows velocity variation, with high-velocity areas in green
or yellow and low-velocity areas in blue. When the vortex shedding is periodic the chance for
VIV is very high at normal Reynolds number range. But at low Reynolds number (<300) and
high Reynolds number (>105 ) even at non-periodic shedding of vortices it can lead to irregular
vibration response of the structure.
The close-spaced thin streamlines about the cylinder demonstrate that there is a high-
velocity shear layer in which fluid speed rapidly changes. As it the streamlines detaches from
cylinder showing flow separation downstream. The small secondary vortex near the cylinder's
surface suggests a localized pressure variation, which can contribute to unsteady forces.
CONCLUSION
The simulation outcomes show a good visualization of vortex shedding and vortex-induced
vibrations (VIV) about a cylindrical shape, which is of utmost importance in offshore structural
design. The frequency spectrum analysis of Ansys shows the presence of a strong peak at 6-10
Hz, signifying a periodic vortex shedding phenomenon. The Strouhal number (St) is
approximately 0.2, as expected for this Reynolds number range. The velocity contours at 0.75s
and 1.37s illustrate the development of the wake from early formation of vortices to a fully
82
developed Kármán vortex street. At 0.75s, small vortices are starting to detach from the surface
of the cylinder, and by 1.37s, a periodic and stable regime of vortex shedding is established.
This periodic wake pattern leads to varying lift forces that can cause cross-flow oscillations,
resulting in VIV excitation if the frequency of shedding is in resonance with the natural
frequency of the cylinder. This oscillatory phenomenon is a significant issue in offshore
engineering since it can lead to fatigue failure in structures such as risers, mooring cables, and
floating wind turbine platforms.
The pressure contour plot emphasizes a high-pressure stagnation region at the upstream and
low-pressure regions in the wake, inducing drag and lift forces that affect structural oscillations.
The streamline contour also illustrates shear layer separation and recirculating vortices, further
affecting wake instability and vortex development. The alternating vortices cause oscillatory
lift forces, which may become a cause of resonance if not controlled. This highlights the need
for VIV suppression methods, like helical strakes or dampers, to reduce structural fatigue and
increase operational stability in offshore conditions. The findings confirm classical vortex
shedding behavior and indicate the importance of fluid-structure interaction analysis in
offshore structural design.
REFERENCES
1. Ali, B. M. S. (2023). Numerical modeling of the flow around a cylinder using FEATool
Multiphysics. Engineering, Technology & Applied Science Research, 13(4), 11290-11297.
https://doi.org/10.48084/etasr.6053
2. Sowoud, K. M., AL-Filfily, A. A., & Abed, B. H. (Year). Numerical investigation of 2D
turbulent flow past a circular cylinder at lower subcritical Reynolds number. [Conference
Name or Journal Name, Volume(Issue), Page Range]. Middle Technical University,
Baghdad, Iraq. doi:10.1088/1757-899X/881/1/01216
3. Rajani, B. N., Kandasamy, A., & Majumdar, S. (Year). Numerical simulation of laminar
flow past a circular cylinder. [Journal Name, Volume(Issue), Page Range].
https://doi.org/10.1016/j.apm.2008.01.017
83