Fact-checked by Grok 2 weeks ago

Morison equation

The Morison equation is a semi-empirical used to calculate the hydrodynamic forces exerted by waves and currents on slender, cylindrical marine structures, such as offshore piles and legs of platforms, where the structure's is small compared to the (typically D/L < 0.2). Developed in 1950 by J.R. Morison, M.P. O'Brien, J.W. Johnson, and S.A. Schaaf based on laboratory experiments with model piles subjected to oscillatory wave motion, it addresses the limitations of inviscid potential flow theory by incorporating viscous effects. The equation assumes linear wave theory for fluid kinematics and neglects wave diffraction, making it suitable for non-breaking waves on vertical or near-vertical members. The total force per unit length f on a structural element is given by f = f_D + f_I, where the drag force f_D = \frac{1}{2} \rho C_D D |u| u accounts for velocity-dependent viscous drag, and the inertia force f_I = \rho C_M \left( \frac{\pi D^2}{4} \right) \dot{u} represents acceleration-dependent pressure and added mass effects. Here, \rho is the fluid density, D is the diameter, u and \dot{u} are the undisturbed horizontal fluid velocity and acceleration, C_D is the drag coefficient (typically 0.6–1.2 for rough cylinders), and C_M is the inertia coefficient (often 1.5–2.0 for circular sections). These empirical coefficients are determined from model tests or field data, varying with Reynolds number, Keulegan-Carpenter number, and surface roughness. Originally formulated for fixed vertical piles extending from the seabed through the water surface, the equation has been extended to moving structures by incorporating relative velocity and acceleration, as f = \frac{1}{2} \rho C_D D |u - \dot{x}| (u - \dot{x}) + \rho C_M V \dot{u} - \rho (C_M - 1) V \ddot{x}, where x is the structural motion and V is the displaced volume. It remains a cornerstone of offshore engineering design standards, such as , for predicting loads on jackets, tension-leg platforms, and subsea templates under combined wave-current conditions. Despite its simplicity, limitations include inaccuracy for large-diameter members (where diffraction dominates), steep waves, or multi-member interactions, prompting ongoing refinements like wake effects and vortex-induced vibrations.

Introduction

Historical Development

The Morison equation was introduced in 1950 by J. R. Morison, M. P. O'Brien, J. W. Johnson, and S. A. Schaaf in their seminal paper titled "The Force Exerted by Surface Waves on Piles," published in the Journal of Petroleum Technology. This work presented an empirical formulation to estimate the hydrodynamic forces exerted by unbroken surface waves on cylindrical piles, combining drag and inertia components based on experimental observations. The equation emerged from studies aimed at understanding wave-induced loads on offshore pilings during the early expansion of petroleum exploration in coastal waters. The original development stemmed from wave tank experiments conducted at the University of California, Berkeley, where researchers measured forces on vertical circular piles under controlled oscillatory flows simulating ocean waves. These laboratory tests involved varying pile diameters, wave heights, and periods to fit an empirical model that captured the observed force profiles, revealing the relative contributions of velocity-squared drag and acceleration-proportional inertia terms. Initially referred to as the MOJS equation after its four authors, the formulation provided a practical tool for engineers designing pile-supported structures in wave environments. In the years following its publication, the equation underwent initial validations through field measurements on offshore structures during the 1950s, confirming its applicability beyond controlled settings. Notable efforts included the 1951 Monterey field tests by Snodgrass, Rice, and Hall, which recorded wave forces on instrumented pilings to assess real-sea conditions. Further laboratory studies were presented by Morison, Johnson, and O'Brien in 1953. These validations demonstrated reasonable agreement for non-breaking waves, establishing the equation's role in early offshore engineering practices, though impact forces from breakers required additional considerations.

Overview and Scope

The Morison equation is a semi-empirical relation for estimating the inline hydrodynamic forces acting on slender structures with a small diameter-to-wavelength ratio in oscillatory flows, such as those generated by ocean waves and currents. It combines two primary force components: an inertia force proportional to the acceleration of the surrounding fluid, and a drag force proportional to the square of the relative velocity between the fluid and the structure. This formulation allows for practical predictions of wave-induced loads on structural elements like piles and tubular members. The equation's scope is centered on ocean engineering applications where the diameter of the body is significantly smaller than the wavelength of the oscillatory flow—typically less than 20% of the wavelength—to ensure the slender body approximation holds and diffraction effects are negligible. It is widely applied in the design and analysis of offshore installations, including oil platforms and wind turbines, to assess hydrodynamic loading under combined wave and current conditions. For larger structures approaching the wavelength scale, diffraction effects become dominant, rendering the Morison approach less suitable. As a semi-empirical model, the Morison equation extends beyond inviscid potential flow theory by incorporating viscous effects through calibrated coefficients, enabling it to capture both added mass and frictional drag in real fluid environments. This distinction makes it a cornerstone tool for engineering approximations where full viscous simulations are computationally prohibitive.

Mathematical Formulation

Derivation

The derivation of the Morison equation begins from Newton's second law applied to the motion of a structure in a fluid, where the net hydrodynamic force equals the rate of change of momentum of the structure plus the surrounding fluid. This force arises from the integration of pressure and viscous shear stresses over the structure's surface, but for slender bodies where the diameter is much smaller than the wavelength, the interaction is simplified by decomposing the force into an inertial component proportional to fluid acceleration and a viscous drag component proportional to relative velocity squared. The inertial term originates from inviscid potential flow theory, capturing the unsteady pressure forces due to accelerating fluid. In undisturbed potential flow past a circular cylinder, the force includes the Froude-Krylov component from the pressure gradient in the incident flow, which is equivalent to the mass of displaced fluid times the fluid acceleration, plus an added mass component from the kinetic energy of the fluid induced by the body's motion. The added mass coefficient C_a quantifies the effective fluid mass entrained by the body; for a circular cylinder, C_a = 1 in two-dimensional potential flow. Thus, the total inertia coefficient is given by C_m = 1 + C_a, where the "1" represents the displaced fluid mass and C_a the added mass. This formulation is extended to oscillatory flows by assuming the local flow acceleration dominates and potential flow approximations hold over short timescales. The drag term adopts an empirical quadratic form inspired by steady-flow viscous drag correlations, such as those for cylinders in uniform currents, where force scales with the square of velocity to account for boundary layer separation and wake formation. Morison et al. hypothesized that this nonlinear viscous effect persists in unsteady oscillatory flows, combining it with the inertial term to model the total local force per unit length without deriving it from first principles of viscous flow. The total force on an extended structure is obtained by integrating the local force contributions along its length, assuming spatial uniformity of the flow over each cross-section and gradual variation along the axis relative to the structure's diameter. This allows evaluation using depth-dependent velocity and acceleration from at each point. Finally, the drag and inertia coefficients are calibrated empirically via least-squares regression against experimental measurements of total force, using known incident flow kinematics to isolate and fit each term's contribution. This method, applied in early wave tank tests on piles, confirms the superposition assumption and determines coefficient values for specific conditions.

Fixed Structures

The Morison equation for fixed structures, such as stationary offshore piles or legs, estimates the hydrodynamic force exerted by oscillatory fluid flows, like waves, on slender bodies where the structure diameter is small compared to the wavelength. For a fixed body with zero velocity relative to the fluid, the total force F is given by F = \rho C_m V \dot{u} + \frac{1}{2} \rho C_d A u |u|, where \rho is the fluid density, C_m is the inertia coefficient, V is the displaced volume of the body, \dot{u} is the undisturbed fluid acceleration, C_d is the drag coefficient, A is the projected area perpendicular to the flow, and u is the undisturbed fluid velocity. This equation decomposes the total force into an inertia component F_I = \rho C_m V \dot{u}, which arises from the pressure gradient in the accelerating fluid and the added mass effect around the body, and a drag component F_D = \frac{1}{2} \rho C_d A u |u|, which accounts for the nonlinear viscous drag due to the relative flow past the structure. The inertia term dominates in low-velocity, high-acceleration regimes, such as deep-water waves, while the drag term becomes significant in shallower waters or with higher velocities. For vertical cylinders, common in fixed offshore platforms, the force is computed by integrating the local force per unit length along the submerged depth, as velocity and acceleration vary with elevation due to wave kinematics. Using linear wave theory, the horizontal velocity u(z, t) and acceleration \dot{u}(z, t) at depth z (with z = 0 at the mean water level and negative downward) for a sinusoidal wave of height H, period T, wavenumber k, and water depth d are u(z, t) = \frac{\pi H}{T} \frac{\cosh[k(z + d)]}{\sinh(kd)} \cos(kx - \omega t), \dot{u}(z, t) = -\frac{2\pi^2 H}{T^2} \frac{\cosh[k(z + d)]}{\sinh(kd)} \sin(kx - \omega t), where \omega = 2\pi / T is the angular frequency. The force per unit length f(z, t) on a cylinder of diameter D is then f(z, t) = \rho C_m \left( \frac{\pi D^2}{4} \right) \dot{u}(z, t) + \frac{1}{2} \rho C_d D u(z, t) |u(z, t)|, and the total horizontal force is F(t) = \int_{-d}^{0} f(z, t) \, dz (or up to the wave crest if applicable). This integration captures the depth-dependent distribution, with maximum forces typically occurring near the surface where velocities peak. In practice, full depth integration often amplifies the forces by a factor related to the hyperbolic distribution, emphasizing the need for site-specific wave parameters.

Moving Structures

For moving structures, such as floating offshore platforms or risers, the Morison equation is extended to include the effects of body motion by replacing absolute fluid kinematics with relative values between the fluid and structure, enabling accurate prediction of hydrodynamic forces in dynamic conditions. This adaptation is essential for time-domain simulations where structural responses interact with wave-induced flows, distinguishing it from the fixed structure formulation by incorporating terms that account for the body's velocity and acceleration. The force F on a moving body is expressed as: F = \rho V \dot{u} + \rho C_a V (\dot{u} - \dot{v}) + \frac{1}{2} \rho C_d A (u - v) |u - v| where \rho is the fluid density, V is the volume of the body element, A is the projected area normal to the flow, u and \dot{u} are the undisturbed fluid velocity and acceleration, v and \dot{v} are the corresponding body values, C_a is the , and C_d is the . The first term, \rho V \dot{u}, represents the unsteady pressure gradient force () due to the fluid acceleration alone. The second term, \rho C_a V (\dot{u} - \dot{v}), captures the relative inertia effect, arising from the added mass of fluid accelerated with the body. The third term is the nonlinear drag force based on the square of the relative velocity, modeling viscous effects. This form is integrated along the structure's length in time-domain simulations to compute dynamic responses, such as those of tension-leg platforms under wave loading. For instance, in studies of oscillating cylinders in regular waves, the relative terms highlight resonance effects when the body's natural frequency aligns with the wave frequency, amplifying forces by up to several times the static values and emphasizing the need for tuned damping.

Coefficients and Parameters

Drag Coefficient

The drag coefficient C_d in the Morison equation is a non-dimensional parameter that quantifies the magnitude of the viscous drag force acting on a slender structure, such as a circular cylinder, in a fluid flow. It appears in the drag force term, which is proportional to the square of the relative fluid velocity, and is defined such that the drag force per unit length is f_d = \frac{1}{2} \rho C_d D |u| u, where \rho is the fluid density, D is the cylinder diameter, and u is the fluid velocity normal to the cylinder axis. For smooth circular cylinders, typical values of C_d range from 0.6 to 1.2, depending on flow conditions; recommended design values from industry standards include 0.65 for smooth surfaces and up to 1.05 for roughened or marine-growth-covered cylinders. The value of C_d depends primarily on the \mathrm{KC} = \frac{u_{\max} T}{D} (where u_{\max} is the maximum velocity amplitude and T is the flow period), the \mathrm{Re} = \frac{u_{\max} D}{\nu} (where \nu is the kinematic viscosity), and surface roughness, with roughness increasing C_d by up to 50% in oscillatory flows. Empirical determination of C_d relies on experimental measurements from both steady uniform flow tests and oscillatory flow tests. In steady flow experiments, C_d is obtained by correlating drag forces with \mathrm{Re}, using approximations like for low-\mathrm{Re} regimes where viscous effects dominate without flow separation. For oscillatory flows relevant to waves, values are derived from specialized facilities like U-tube oscillators, with seminal curves developed by in the 1970s and 1980s showing C_d as a function of \mathrm{KC} and \mathrm{Re} for both smooth and rough cylinders. Variations in C_d are pronounced at low \mathrm{KC} (< 10), where values are higher due to enhanced flow separation and initial vortex formation around the cylinder, leading to increased drag relative to inertial effects. In high-\mathrm{Re} flows (> $10^5), typically post-critical, C_d stabilizes around 0.6–0.7 for smooth cylinders as suppresses separation, making the drag term dominant in regimes where velocity-squared effects prevail over acceleration-based forces.

Inertia Coefficient

The inertia coefficient C_m in the Morison equation quantifies the inertial loading on a due to the of the surrounding fluid. It is defined as C_m = 1 + C_a, where the value of 1 corresponds to the mass of fluid displaced by the , and C_a is the coefficient representing the additional fluid mass accelerated in the vicinity of the . In inviscid potential flow theory, C_a = 1 for a two-dimensional circular cylinder, yielding an ideal C_m = 2. This theoretical value assumes irrotational flow without viscous effects or flow separation. The coefficient's magnitude primarily depends on the Keulegan-Carpenter number (KC), which characterizes the oscillatory flow regime, and shows relatively low sensitivity to the Reynolds number compared to the drag coefficient. At high KC values, C_m decreases owing to vorticity generation and the formation of circulating flows around the structure, which reduce the effective added mass. Experimental determination of C_m integrates predictions with measurements from wave tank tests, as pioneered in the seminal work of Morison et al. (1950), which reported values near 2 for small-amplitude waves on piled structures. Subsequent validations, including large-scale experiments, have refined these estimates while confirming the theoretical baseline for low-KC conditions. For practical applications in oscillatory flows, C_m \approx 2 holds for KC < 6, with empirical relations depicting a progressive decline—such as to approximately 1.5 at KC ≈ 20—derived from regression analyses of force measurements. These variations underscore the need for -based selection in engineering design.

Applications

Offshore Engineering

The Morison equation is widely applied in offshore engineering to estimate wave and current loads on fixed platforms, jackets, and piles, particularly for slender members where the structure diameter is small relative to the wavelength. In regions like the North Sea, it facilitates the calculation of hydrodynamic forces on steel jacket platforms exposed to severe wave environments, enabling designers to assess structural integrity under combined environmental loading. This approach is essential for pile foundations in fixed offshore oil rigs, where currents and oscillatory waves contribute to both drag and inertia components of the force. Integration of the Morison equation with linear wave theory, such as , provides velocity profiles for input into force computations, allowing for the determination of total base shear and overturning moments on offshore structures. approximates water particle kinematics under small-amplitude waves, which are then used in the fixed-structure form of the equation to predict distributed loads along vertical and horizontal members. These calculations are critical for evaluating global stability, such as the shear at the mudline and moments that could lead to foundation failure in deepwater installations. A notable early application occurred in the analysis of wave force data from Gulf of Mexico offshore platforms during the late 1960s and 1970s, where the Morison equation was calibrated against measured forces from hurricane events and used to predict responses to individual large waves, showing good agreement between observed and calculated values. This validation demonstrated the equation's reliability for irregular seas, with predicted forces aligning closely with instrumentation data after adjusting coefficients based on site-specific measurements. In modern practice, the Morison equation is implemented in software tools like for time-domain simulations of offshore structures, incorporating wave-current interactions to model dynamic responses of fixed platforms. Similarly, employs the equation for hydrodynamic load prediction in coupled analyses, supporting detailed assessments of base shear and moments under transient conditions.

Other Engineering Contexts

The Morison equation has been adapted to estimate hydrodynamic forces on tidal turbine blades operating in oscillatory tidal flows, where slender rotor elements experience combined steady current and wave-induced velocities. In experimental studies, an extended form of the equation separates drag into mean and oscillatory components to model blade root bending moments, assuming attached flow conditions and neglecting Froude-Krylov forces for simplicity. This approach reveals that inertia contributions are minor relative to drag, particularly at low reduced frequencies, allowing quasi-steady models to approximate loads effectively on stiff blades with constant rotational speed. For coastal structures such as breakwaters and piers, the Morison equation calculates wave and current forces on vertical cylinder arrays, which form essential components like pile-supported platforms. Experimental investigations demonstrate that inter-cylinder spacing significantly influences total forces, with reductions in gap-to-diameter ratios below 0.25 leading to rapid increases in array loading due to flow blockage, while currents of 0.15 m/s amplify individual cylinder forces by 10-20%. The equation applies to non-breaking waves on small-diameter piles (diameter/wavelength < 0.2), incorporating drag coefficients of 0.7-2.5 and inertia coefficients of 1.05-2.5 to predict horizontal forces per unit length. In cases of breaking waves on pier piles, it estimates impact forces based on local crest elevations, aligning with guidelines for piled jetties exposed to oscillatory flows. The drag component of the Morison equation finds analogy in wind engineering for estimating gust loads on slender towers, where quasi-steady aerodynamic forces are modeled by adapting the quadratic drag term to account for air density and wind velocity profiles, though the inertia term is typically negligible due to lower fluid mass. This adaptation supports load spectra analysis for offshore tower structures under combined wind and wave conditions, highlighting similarities in force superposition for dynamic response prediction. Post-2000 applications extend the Morison equation to floating wind turbines, where it computes viscous drag and inertia on platform members like barges in deepwater sites (200 m depth), integrated into tools such as for coupled aero-hydro-elastic simulations of the . For such systems, the equation handles nonlinear wave kinematics on slender cylinders but requires supplementation with potential flow methods for radiation and diffraction effects, revealing excessive pitch motions under extreme conditions that necessitate damping enhancements. Recent 2025 studies have incorporated frequency-dependent coefficients into the Morison equation for improved hydrodynamic modeling of offshore wind turbines. In subsea equipment, recent 2020s studies apply the equation within finite element models like to assess hydrodynamic loads on deepwater steel catenary risers (2000 m depth) under 100-year return period waves and currents, optimizing designs via bio-inspired algorithms to reduce weight by 35.76% while evaluating stresses at hang-off and touchdown points, with drag coefficients of 1.6 and inertia coefficients of 2.0.

Limitations and Extensions

Key Limitations

The Morison equation relies on the fundamental assumption that the structure is slender, meaning its characteristic diameter D must be significantly smaller than the incident wavelength \lambda, typically D/\lambda < 0.2, for the equation to accurately predict forces without diffraction effects dominating. When this ratio exceeds 0.2, wave diffraction around the body becomes significant, rendering the inertia and drag components unreliable as they fail to capture the altered pressure field. This limitation restricts the equation's use to small-diameter elements like piles or risers, excluding larger structures such as caissons or breakwaters where full diffraction theory is required instead. Another key constraint arises from the equation's neglect of non-uniform flow fields, assuming spatially uniform velocity and acceleration across the body's cross-section, which introduces errors in steep waves or near the free surface where vertical variations in kinematics are pronounced. In such conditions, the underpinning the velocity profile overpredicts or underpredicts local forces, particularly in the splash zone or under breaking waves, leading to inaccuracies in total load estimates. This assumption holds reasonably only for deep submergence or gentle wave slopes, but falters in realistic sea states with strong gradients. The equation performs poorly in the intermediate Keulegan-Carpenter (KC) number regime, specifically $6 < \mathrm{KC} < 20, where oscillatory flow induces complex vortex shedding that the simple drag term cannot adequately model, often underpredicting peak forces. At these KC values, the flow transitions from inertia-dominated to a regime influenced by boundary layer separation and wake dynamics, which the empirical formulation overlooks, resulting in scattered experimental validation. This gap is particularly evident for rough or marine-growth-covered cylinders, amplifying discrepancies. Furthermore, the Morison equation exclusively computes inline (horizontal) forces, ignoring transverse components such as lift forces or those from vortex-induced vibrations (VIV), which can be substantial for KC > 6 and contribute to or dynamic in flexible structures. Without accounting for these perpendicular loads, the model underestimates total hydrodynamic loading and risks, necessitating supplementary analyses for comprehensive . The reliance on empirical drag (C_D) and inertia (C_M) coefficients introduces inherent uncertainty, as their values exhibit significant variability—often 20-40% scatter—for non-standard geometries, such as inclined, roughened, or non-circular members, due to limited experimental data and sensitivity to or surface conditions. Calibration from canonical tests (e.g., smooth vertical cylinders) does not generalize well, leading to conservative or optimistic force predictions in unconventional applications like jacket legs with marine fouling.

Modern Extensions

Modern extensions of the Morison equation have addressed key limitations in handling complex flow regimes, particularly for high numbers where wake effects become prominent. Sarpkaya's 1986 experimental on smooth and rough cylinders in oscillatory flows at high Reynolds numbers introduced modifications to incorporate wake-induced forces for KC > 10, demonstrating through U-tube experiments that and in the wake significantly alter drag and inertia coefficients, leading to more accurate force predictions in regimes where the original equation underestimates nonlinear effects. To extend applicability to three-dimensional geometries, refinements have been developed for non-circular cross-sections and inclined structural members. For non-circular sections, such as rectangular or square profiles common in platforms, the and form factors are adjusted in the term to account for shape-dependent , improving load estimates in validation studies against experimental data. For inclined members, the velocity components are resolved normal and tangential to the axis, with the normal component feeding into the standard Morison terms while tangential effects are often neglected or approximated via empirical reductions in coefficients, enabling reliable predictions for structures with sloped legs. Additionally, models combining theory with terms, such as the MacCamy-Fuchs approach adapted for intermediate body sizes, blend solutions for large-diameter cylinders (where KC < 2 and dominates) with Morison's component, providing a smooth transition for bodies where neither regime fully applies and reducing errors in wave force calculations by integrating analytical kernels with empirical . Numerical enhancements in the 2010s leveraged (CFD), particularly Reynolds-averaged Navier-Stokes (RANS) models, to validate and refine coefficient predictions. These simulations, often using k-ε or k-ω turbulence closures, have shown that traditional constant coefficients overlook flow unsteadiness and separation, allowing for dynamic tuning based on local KC and Reynolds numbers; for instance, RANS validations against towing tank experiments improved inertia coefficient accuracy for oscillatory flows around vertical piles, establishing CFD as a for calibrating Morison parameters in software. Recent advancements up to 2025 incorporate for coefficient tuning and coupled hydro-elastic formulations for flexible structures. models, trained on large datasets from CFD and experiments, predict drag and inertia coefficients as functions of KC, , and , achieving improved prediction accuracy across diverse scenarios and enabling adjustments in simulations for irregular waves. Coupled hydro-elastic models integrate the Morison equation with finite element , accounting for deformation-induced changes in hydrodynamic loading on flexible elements like lines or VLFS connectors; for example, time-domain analyses of platforms reveal that flexibility amplifies local stresses under combined waves and currents, with the model iteratively updating geometry and velocities for enhanced fidelity.

References

  1. [1]
    [PDF] Assessment of the Morison Equation. - DTIC
    The Morison Equation, derived 30 years ago, is an engineering model used to determine wave forces on structures. Given the geometry of the structure, the water ...
  2. [2]
    [PDF] 13.012 READING 7: FLUID FORCES ON BODIES 1. Steady Flow In ...
    Morrison's equation can be used to formulate the force on a cylinder in waves with a wavelength much larger than the body diameter. Recall from basic wave ...
  3. [3]
    [PDF] 13.42 04/01/04: Morrison's Equation
    Apr 1, 2004 · The approach in practice is to evaluate, using Morrison's formula, the force acting on a small section of the cylinder (eq. 1.2) at each depth ...Missing: original | Show results with:original
  4. [4]
    The Force Exerted by Surface Waves on Piles - OnePetro
    Journal Paper| May 01 1950. The Force Exerted by Surface Waves on Piles Available. J.R. Morison ... Original copyright American Institute of Mining ...
  5. [5]
    The Force Exerted by Surface Waves on Piles - OnePetro
    1. Vol. 189, 1950. PETROlEUM TRANSACTIONS, AIME. 149. Page 2. T.P.2846. THE FORCE EXERTED BY SURFACE WAVES ON PILES as functions of z/ d a\ld d/L. Values off"/C ...
  6. [6]
    [PDF] FORCES INDUCED BY BREAKERS ON PILES
    Some field measurements of forces induced on a circular pile by breaking waves in the surf zone are given in papers by Snodgrass, Rice, and Hall,. 1951; Morison ...
  7. [7]
    [PDF] Resistance in Unsteady Flow - Search for a Model - DTIC
    The new model is expected to replace the Morison equation (hereafter called the. MOJS equation) and offer greater universality, particularly in the inertia/drag ...
  8. [8]
    Morison's equation - Orcina
    This equation was originally formulated by Morison et al (1950), for calculating the wave loads on fixed vertical cylinders, as having two force components.Missing: ocean | Show results with:ocean<|control11|><|separator|>
  9. [9]
    Morison Equation - QBlade Documentation
    The original MCFC is formulated in terms of the normal force per unit length on the cylinder. It can be recast so that the inertia part of the Morison equation ...Missing: ocean | Show results with:ocean
  10. [10]
    Wave Calculation : Morison Equation - MOTIVE
    Jan 31, 2024 · The Morison equation consists of inertia force and drag force, excluding hydrostatic force. In this context, inertia force is related to the ...Missing: original | Show results with:original
  11. [11]
    Morison Equation - an overview | ScienceDirect Topics
    The Morison equation is a semi-empirical equation calculating hydrodynamic forces on submerged objects, considering pressure, added mass, and viscous drag.Missing: original | Show results with:original
  12. [12]
    [PDF] Potential Time Domain Model with Viscous Correction and CFD ...
    Jun 8, 2019 · Morison equation provides a semi-empirical for- mulation to model the unsteady force on rigid structures in oscillatory flow.
  13. [13]
    [PDF] Introduction to Added Mass
    Thus the added mass of a cylinder in potential flow is equal to the mass of fluid displaced by the body whereas the added mass of a sphere in potential flow ...
  14. [14]
    [PDF] Chapter 12 WAVE FORCES ON SLENDER CYLINDERS
    Morison formulated his equation simply by hypothesizing that the superposition of two separate and well know phenomena (drag in a current and hydrodynamic ...
  15. [15]
    [PDF] ssc-347 strategies for nonlinear analysis of marine structures - DTIC
    The test showed that the relative velocity form of the Morison equation is appropriate, even though the observations were limited. Considerable work is needed ...
  16. [16]
    Wave Forces Acting on a Moving Cylinder - ASME Digital Collection
    The Morison equation extended to the case of relative motion is employed to calculate the forces on the cylinder. The inertia, added mass and drag coefficients ...Missing: resonance | Show results with:resonance
  17. [17]
    [PDF] API RP 2A Recommended Practice for Planning, Designing and ...
    Nov 15, 2009 · For typical design situations, global platform wave forces can be calculated using the following values for unshielded circular cylinders:.
  18. [18]
    [PDF] Forces on Roughened Cylinders in Harmonic Flow at High ... - DTIC
    The in—line forces acting on sand—roughened circular cylinders im- mersed in an harmonically oscillating flow have been measured using a. U-shaped water tunnel.
  19. [19]
    [PDF] 2.6 Oseen's improvement for slow flow past a cylinder
    We shall demonstrate the Oseen approximation for the two-dimensional case of a circular ... so that the drag coefficient is. CD = D. 1. 2. ρW2 2a. = 2π. 1. 1. 2 − ...
  20. [20]
    [PDF] Drag and inertia coefficients for horizontally submerged rectangular ...
    The high drag coefficients at low KC numbers are associated with the flow separation and the first appearances of vortices, which take place for sharp-edged ...
  21. [21]
    [PDF] Values of the Added Mass
    Added mass values are tabulated for shapes like circular cylinders (M11=ρπR), elliptical cylinders, and spheres (M11=M22=M33=2/3ρπR).
  22. [22]
    [PDF] DETERMINATION OF COEFFICIENTS IN MORISON FORMULA BY ...
    A few procedures are established for the determination of the drag and inertia coefficients from the measurements. Some authors claim that the coefficients ...
  23. [23]
    The Force Exerted by Surface Waves on Piles - Semantic Scholar
    The Force Exerted by Surface Waves on Piles · J. Morison, J. Johnson, S. A. Schaaf · Published 1 May 1950 · Engineering, Geology · Journal of Petroleum Technology.
  24. [24]
    Wave-in-deck loading on fixed steel jacket decks - ScienceDirect.com
    In this paper focus has been on finding realistic load time histories for wave-in-deck loading on jacket platforms in the North Sea.
  25. [25]
    [PDF] Hydrodynamics - USFOS
    Airy waves are based on linear wave theory. Consider wave propagating along ... theory while the pressure integration is based in stretched Airy (Wheeler) theory.
  26. [26]
    Wave Load Analysis in Structural Engineering - SDC Verifier
    Jun 19, 2024 · It is typically used in conjunction with Morison's equation to estimate wave forces on slender members. SDC Verifier allows users to define ...
  27. [27]
    Analysis of Wave Force Data | Journal of Petroleum Technology
    Mar 1, 1970 · The relative agreement of the predicted and observed forces is used as a measure of the degree of correctness of the Morison equation and as a ...<|control11|><|separator|>
  28. [28]
    3XL monopile lowering analysis: A comprehensive study
    Nov 1, 2024 · OrcaFlex was used to simulate the time domain response of the MP during lowering operation using Morison's approach. ANSYS AQWA model.Missing: implementation | Show results with:implementation
  29. [29]
    [PDF] Tidal Turbine Blade Load Experiments for Oscillatory Motion
    This paper presents results of an experimental based study which investigated the sensitivity of the blade loads of a scale model tidal turbine to unsteady ...
  30. [30]
    [PDF] 1 AN EXPERIMENTAL STUDY OF WAVE AND CURRENT ...
    Arrays of vertical cylinders are essential components of many coastal structures, such as bridge piers, platform legs, and pile breakwaters.
  31. [31]
    [PDF] Piers, jetties and related structures exposed to waves:
    The total wave force (Ftw) 'imposed' on a platform deck can be thought of as a slightly extended modification of Morison's equation, given by the following:.
  32. [32]
    Impact of focused wave groups on pier structures: a case study of ...
    ... Morison equation for breaking waves, as described in Goda et al. (1966). The calculation is based on the local wave conditions at x = 53.55 m, assuming that ...<|separator|>
  33. [33]
    Load spectra for slender offshore structures in waves and currents
    The analysis permits the implications of the modified Morison equation to be ... 4 Kurtis Gurley, Ahsan Kareem, Gust loading factors for tension leg ...
  34. [34]
    [PDF] Dynamics Modeling and Loads Analysis of an Offshore Floating ...
    in Morison's equation, other than those loads transmitted from the wind turbine and the weight of the support platform. Viscous i. dF. = ith component of the ...Missing: analogies | Show results with:analogies
  35. [35]
    [PDF] OPTIMISATION OF DEEPWATER RISER USING BIO-INSPIRED ...
    This research aims to optimise a deepwater riser using a bioinspired design while ensuring its structural integrity is maintained. The study notes the ...<|control11|><|separator|>
  36. [36]
    [PDF] Applications of Morison's equation to circular cylinders of varying ...
    Morison's equation is used for formulating forces and moments due to waves acting on circular cylinders of varying cross-sections and truncated forms.
  37. [37]
    3. Loads on Slender Bodies – Ocean Hydrodynamics for Engineers
    Limitations of Morison's Equation · Lift forces: The equation does not account for fluctuating lift forces that may arise, such as those generated by vortex ...
  38. [38]
    [PDF] In-line and transverse forces on smooth and rough cylinders ... - CORE
    Jul 4, 1986 · segment was mounted concentrically on a supporting cylinder whose ends were in turn connected to two force transducers.
  39. [39]
    A semi-analytical hydrodynamic model for floating offshore wind ...
    Nov 1, 2023 · This paper proposes a semi-analytical MacCamy–Fuchs–Morison (MFM) hybrid hydrodynamic model for very efficient simulations of FOWT responses without resorting ...
  40. [40]
    Benchmarking of a Computational Fluid Dynamics-Based Numerical ...
    ... Morison equation results are in agreement with CFD and experimental results. However, for wavelength to diameter less than π 2 π Morison equation will ...Case Studies And Results · Free Floating Barge · Wave Loads On A...
  41. [41]
    Machine Learning Prediction of the Morison Equation Coefficients
    Aug 5, 2025 · The Morison equation approximates the forces on a cylinder due to wave particle motion. The formula is semi-empirical and distinguishes ...
  42. [42]
    Hydroelastic effects on hydrodynamic loads on an LNG pump tower ...
    Sep 1, 2024 · We also compared these loads with approximated loads based on Morison's equation. 2. Test case description. Simulations were performed for a ...