Fact-checked by Grok 2 weeks ago

Pulsatile flow

Pulsatile flow is the rhythmic, intermittent propagation of a through a or piping system, in contrast to constant, smooth propagation that produces . This type of features periodic variations in and , often driven by an oscillatory , resulting in pressure pulses that propagate through the medium. In , pulsatile flow is distinguished by its unsteady nature, where the flow rate oscillates over time, typically modeled as a superposition of steady and oscillatory components. A key dimensionless parameter governing its behavior is the (α), defined as α = R √(ω/ν), where R is the tube radius, ω is the of , and ν is the kinematic ; it represents the ratio of transient inertial forces to viscous forces, influencing profiles from parabolic (low α) to nearly flat in the core with a thin near the wall (high α). For biological flows like in arteries, α typically ranges from 1 to 20, leading to complex phenomena such as wave propagation and wall shear stress variations that differ markedly from steady Poiseuille flow. Pulsatile flow plays a critical role in , particularly in the cardiovascular system, where it mimics the systolic ejection and diastolic relaxation phases of the , ensuring efficient nutrient delivery while subjecting vessel walls to oscillatory that can influence endothelial function and disease processes like . In engineering and biomedical applications, it is harnessed in microfluidic devices to replicate physiological conditions for , , and biosensing, offering advantages like enhanced particle manipulation and reduced compared to steady flows. Additionally, in circulatory systems such as left ventricular assist devices (LVADs), generating pulsatile profiles has been shown to improve end-organ and reduce complications like .

Introduction

Definition

Pulsatile flow refers to motion characterized by periodic variations in and over time, typically induced by a time-varying that is often sinusoidal, distinguishing it from steady flows where these quantities remain constant. This oscillatory nature arises from external periodic forcing, leading to time-dependent behavior in contrast to the uniform characteristics of steady laminar or turbulent flows. In its basic physical setup, pulsatile flow is modeled as occurring within rigid or elastic tubes, driven by an oscillating pressure difference along the tube length. A key real-world example is blood circulation in arteries, where the flow becomes pulsatile due to the rhythmic contraction and relaxation of the heart during each . Theoretical analyses commonly adopt assumptions of an incompressible , axisymmetric flow in cylindrical pipes, and no-slip boundary conditions at the vessel walls to simplify the governing dynamics. Pulsatile flow was first systematically studied in the 1950s by mathematician John R. Womersley, who developed foundational models specifically for arterial blood flow.

Significance

Pulsatile flow plays a pivotal role in the cardiovascular system, where the heart's rhythmic contractions generate periodic pressure and velocity variations essential for maintaining normal organ function, particularly in critical organs such as the brain, kidneys, and heart. This natural pulsatility enhances end-organ perfusion by promoting tissue metabolism, reducing edema formation, and lowering peripheral vascular resistance compared to steady flow conditions. In arterial mechanics, pulsatile blood flow interacts with vessel geometries to produce spatially and temporally varying biomechanical forces on the endothelial walls, influencing shear stress distribution and vascular remodeling. Disruptions in pulsatile patterns contribute to disease states like atherosclerosis, where abnormal flow induces endothelial dysfunction and plaque progression; for instance, low or oscillatory shear stress can accelerate lesion formation at arterial bifurcations. Additionally, pulsatile flow influences oxygen transport to tissues through variations in convective mixing and diffusion across the blood-vessel interface, as studied in models of stenosed arteries. In engineering applications, pulsatile flow is integral to the design of devices and systems involving periodic forcing, such as microfluidic pumps that replicate physiological conditions for and bioassays, where it improves mixing efficiency and particle manipulation over steady flows. In ocean engineering, wave-induced pulsatile flows govern the dynamics of structures like risers and pipelines, affecting and under oscillatory environmental loads. Biomedical devices, including artificial hearts and ventricular assist systems, incorporate pulsatile mechanisms to mimic natural , reducing complications like associated with non-pulsatile alternatives; for example, rotary total artificial hearts generate inherent pulsatility to support long-term cardiac replacement. Beyond and , pulsatile flow exhibits distinct advantages over steady flow in energy dissipation and transport processes, with oscillatory components increasing overall dissipation rates while enabling superior mixing in confined geometries, which is critical for applications in ventilators simulating respiratory cycles and pipelines handling variable pressures. In modern bioengineering, post-2020 advancements leverage pulsatile flow in organ-on-chip models to recapitulate vascular microenvironments, such as in tri-culture coronary chips that assess endothelial responses under physiological . Computational simulations of patient-specific pulsatile flows further enable by predicting hemodynamic risks in cardiovascular diseases, informing tailored interventions like placement or strategies.

Mathematical Modeling

Governing Equations

Pulsatile flow in rigid tubes is fundamentally governed by the incompressible Navier-Stokes equations coupled with the , expressed in cylindrical coordinates (r, \theta, z) for unsteady, axisymmetric flow where the \mathbf{u} = (u_r, u_\theta, u_z) has no azimuthal component and depends only on r and t. The ensures mass conservation: \frac{1}{r} \frac{\partial (r u_r)}{\partial r} + \frac{\partial u_z}{\partial z} = 0, while the momentum equations in the radial and axial directions are: \frac{\partial u_r}{\partial t} + u_r \frac{\partial u_r}{\partial r} + u_z \frac{\partial u_r}{\partial z} - \frac{u_\theta^2}{r} = -\frac{1}{\rho} \frac{\partial p}{\partial r} + \nu \left( \frac{\partial^2 u_r}{\partial r^2} + \frac{1}{r} \frac{\partial u_r}{\partial r} - \frac{u_r}{r^2} + \frac{\partial^2 u_r}{\partial z^2} - \frac{2}{r^2} \frac{\partial u_\theta}{\partial \theta} \right), \frac{\partial u_z}{\partial t} + u_r \frac{\partial u_z}{\partial r} + u_z \frac{\partial u_z}{\partial z} = -\frac{1}{\rho} \frac{\partial p}{\partial z} + \nu \left( \frac{\partial^2 u_z}{\partial r^2} + \frac{1}{r} \frac{\partial u_z}{\partial r} + \frac{\partial^2 u_z}{\partial z^2} \right). These equations describe the balance of inertial, , and viscous forces in the . For typical pulsatile flows, such as those in arterial systems, further simplifications are applied assuming fully developed flow along the tube axis, where \partial u_z / \partial z = 0 and u_r = 0, reducing the problem to a one-dimensional axial u_z(r, t). The is prescribed as oscillatory, often \partial p / \partial z = -A \cos(\omega t) or its complex form \partial p / \partial z = -\text{Re}[A e^{i \omega t}], with A the and \omega the . Under the small- assumption, where the oscillatory component is much smaller than the mean flow, the nonlinear convective terms \mathbf{u} \cdot \nabla \mathbf{u} are neglected, yielding the linearized momentum equation: \frac{\partial u_z}{\partial t} = -\frac{1}{\rho} \frac{\partial p}{\partial z} + \nu \left( \frac{\partial^2 u_z}{\partial r^2} + \frac{1}{r} \frac{\partial u_z}{\partial r} \right). This linearization facilitates analytical solutions and is central to classical analyses of pulsatile flow. Appropriate boundary conditions complete the formulation: no-slip at the tube wall requires u_z(R, t) = 0, where R is the tube radius; symmetry at the centerline imposes \partial u_z / \partial r = 0 at r = 0 to ensure finite velocity; and periodicity in time aligns with the oscillatory pressure, u_z(r, t + 2\pi / \omega) = u_z(r, t). These conditions enforce physical realism for confined, periodic flows. Extensions to more realistic biomedical scenarios incorporate rheology, replacing the constant viscosity \nu with shear-rate-dependent models like the Casson or Carreau-Yasuda equations, which account for blood's yield stress and shear-thinning behavior under pulsatile conditions. Similarly, for elastic walls, the governing equations couple the to a wall deformation model, such as the thin-walled approximation where radial h(t) satisfies a derived from wall elasticity and fluid pressure, enabling fluid-structure interaction analyses. These modifications are essential for accurate modeling of arterial pulsatile flow in contemporary .

Dimensionless Parameters

In pulsatile flow, dimensionless parameters are essential for characterizing the relative importance of inertial, viscous, and unsteady forces, allowing for the scaling of solutions across different physical systems . These parameters emerge from the non-dimensionalization of the Navier-Stokes equations governing in channels or pipes subjected to oscillatory gradients. By normalizing variables such as by a characteristic R, time by the oscillation period $1/f (where f is ), velocity by a mean speed U, and by \rho U^2 (with \rho as fluid density), the equations reveal key non-dimensional groups that dictate flow behavior. The Womersley number, \alpha = R \sqrt{\omega / \nu}, where \omega = 2\pi f is the angular frequency and \nu is the kinematic viscosity, quantifies the ratio of oscillatory inertial forces to viscous forces in pulsatile flows. It arises naturally in the non-dimensional form of the axial momentum equation for laminar, oscillatory pipe flow, where the unsteady term \partial u / \partial t balances the viscous diffusion term \nu \nabla^2 u; scaling time with $1/\omega and length with R yields \alpha^2 as the coefficient multiplying the unsteady term relative to viscous effects. Physically, \alpha represents the ratio of the pipe radius R to the oscillatory boundary layer thickness \delta \approx \sqrt{2\nu / \omega}, indicating how far viscous effects penetrate into the flow during one oscillation cycle. This parameter was first introduced by Womersley in his analytical solution for arterial blood flow, enabling predictions of velocity profiles under pulsatile conditions. Complementing the Womersley number, the Reynolds number \mathrm{Re} = U D / \nu (with D = 2R as diameter) captures the ratio of mean inertial to viscous forces in the steady component of the flow, while the Strouhal number \mathrm{St} = \omega L / U (where L is a characteristic length, often D) measures the relative importance of unsteady (local acceleration) to convective inertial forces. In pulsatile contexts, these parameters interact such that \alpha^2 \approx (\mathrm{Re} \cdot \mathrm{St}) / 4 when L = D, linking unsteadiness across scales; low \mathrm{St} implies quasi-steady behavior dominated by \mathrm{Re}, whereas high \mathrm{St} emphasizes transient effects modulated by \alpha. Together, they classify flow regimes: for \alpha \ll 1, viscous forces dominate throughout the cross-section, yielding parabolic-like profiles similar to steady Poiseuille flow; as \alpha increases beyond approximately 3–5, inertial effects prevail in the core, leading to flatter velocity profiles and phase lags between pressure and flow. In modern (CFD) applications, particularly for high-\alpha flows in biomedical contexts like intracranial aneurysms, these parameters guide simulations of complex geometries and unsteady . For instance, studies in the have used \alpha > 10 alongside elevated \mathrm{[Re](/page/Re)} (up to 1000) to model turbulent transitions and wall distributions in aneurysmal sacs, revealing how oscillatory exacerbates vortex formation and rupture risk under physiological pulsation frequencies around 1 Hz. Such analyses, often employing patient-specific imaging for boundary conditions, underscore the parameters' role in validating CFD models against experiments and advancing predictive tools for cardiovascular interventions.

Flow Properties

Womersley Number

The , denoted as \alpha, is a dimensionless parameter that characterizes the relative importance of unsteady inertial forces to viscous forces in pulsatile flows. It is defined by the formula \alpha = R \sqrt{\frac{\omega}{\nu}}, where R is the radius of the conduit, \omega is the of the pulsatile , and \nu is the kinematic of the fluid. This parameter arises naturally from the nondimensionalization of the Navier-Stokes equations for oscillatory flow in cylindrical tubes, quantifying how pulsatile effects penetrate the viscous . Physically, the governs key aspects of pulsatile flow dynamics, including the phase lag between the driving and the resulting , the of oscillatory components, and the degree of flow flattening across the conduit cross-section. At low \alpha, viscous dominates, leading to minimal phase lag and parabolic-like profiles similar to steady Poiseuille flow; as \alpha increases, inertial effects cause the flow to lag the pressure by up to approximately 90 degrees, reduce viscous of higher-frequency harmonics, and produce a more uniform (flattened) velocity distribution in the core region. In flow, typical Womersley numbers range from approximately 1 to 20, with values around 15–20 in the and 5–6 in the , reflecting variations in and the cardiac pulsation of about 1 Hz. These ranges highlight \alpha's role in hemodynamic contexts, where it influences wave propagation and energy dissipation. The foundational theoretical framework for \alpha was established through analytical solutions to pulsatile flow in rigid tubes, as derived in Womersley's seminal work using to solve the linearized . Experimental validations in controlled pulsatile setups have confirmed these predictions, with modern studies using techniques like and Doppler ultrasound demonstrating accurate reproduction of phase lags and profile shapes for \alpha up to 10. In non-rigid tubes, such as compliant arteries, variable \alpha interacts with wall distensibility to further modulate flow behavior, introducing additional phase shifts and altering through coupled fluid-structure interactions that are critical in contemporary hemodynamic modeling. This extension beyond rigid assumptions underscores \alpha's interpretive value in physiological simulations.

Velocity Profile

In pulsatile flow through a rigid cylindrical tube, the axial velocity profile u(r, t) is derived analytically under the assumption of a fully developed, laminar flow driven by a periodic pressure gradient of the form -\frac{\partial p}{\partial z} = A e^{i \omega t}, where A is the amplitude, \omega is the angular frequency, \rho is the fluid density, R is the tube radius, and r is the radial coordinate. The solution takes the form u(r, t) = \Re \left[ \frac{A}{i \omega \rho} \left( 1 - \frac{J_0 \left( \alpha \sqrt{i} \, \frac{r}{R} \right)}{J_0 \left( \alpha \sqrt{i} \right)} \right) e^{i \omega t} \right], where \Re[\cdot] denotes the real part, J_0 is the zeroth-order Bessel function of the first kind, and \alpha is the Womersley number. This velocity profile exhibits distinct characteristics depending on the value of \alpha. For low \alpha (quasi-steady conditions), the profile approximates a parabolic shape similar to steady Poiseuille flow, with velocity peaking at the centerline and decreasing symmetrically to zero at the wall. In contrast, for high \alpha, the profile becomes flatter across most of the tube cross-section, resembling plug-like flow, where the velocity is nearly uniform except in a thin boundary layer near the wall. Additionally, phase differences arise radially, with near-wall fluid elements lagging the centerline motion due to viscous effects, leading to a non-uniform temporal response across the radius. Over a pulsatile cycle, the profiles evolve dynamically: during phases, the centerline leads the wall , creating forward-skewed profiles; in deceleration, reverse flows can occur near the wall while the core remains forward, resulting in oscillatory layers. These near-wall oscillations contribute to elevated gradients, which are critical in physiological contexts like arterial blood flow. While the analytical solution applies to straight, rigid tubes, numerical methods such as finite element analysis extend these insights to complex , such as bifurcations or aneurysms, where velocity profiles deviate from axisymmetry and exhibit secondary flows. For instance, in idealized carotid bifurcations, finite element simulations reveal asymmetric velocity distributions during peak , with jet-like flows into branches and recirculation zones influenced by plaque , achieving good agreement with experimental data. These approaches incorporate realistic pulsatile inlet conditions and wall compliance, highlighting distortions not captured by one-dimensional models.

Limiting Cases

Low Womersley Number

In pulsatile flow through a cylindrical tube, the low regime occurs when the dimensionless parameter α = R √(ω/ν) is small, typically α < 1–3, where viscous forces dominate over inertial effects due to low oscillation frequencies relative to viscous diffusion timescales. In this limit, the flow approximates quasi-steady conditions, meaning the velocity responds instantaneously to changes in the driving pressure gradient without significant phase differences or wave propagation. The asymptotic behavior of the axial velocity profile in this regime closely resembles the steady Poiseuille parabolic distribution, modulated temporally by the pressure gradient. For a pressure gradient of the form -∂p/∂z = A cos(ωt), the velocity is given by u(r, t) \approx \frac{A}{4\mu} (R^2 - r^2) \cos(\omega t), with minimal phase lag between velocity and pressure, as inertial acceleration terms become negligible compared to viscous diffusion. This approximation holds with low error (typically <5–10% deviation from exact Womersley solutions) for α ≲ 1, but accuracy decreases as α approaches 3, where transitional inertial effects introduce small phase shifts and profile flattening near the wall. The implications of low α include negligible unsteady inertial contributions, resulting in flow rates that remain nearly in phase with the pressure gradient and wall shear stresses that follow the quasi-steady parabolic scaling. This regime is particularly relevant for low-frequency pulsatile systems, such as slow-pulsing peristaltic pumps in chemical processing or hydraulic actuators, where the pulsation period is long enough for viscous equilibrium at each instant. In microscale applications, low Womersley numbers are common due to small tube radii (R ~ 10–100 μm), enabling pulsatile microfluidics in lab-on-a-chip devices for biomimetic simulations, enhanced cell perfusion in organ-on-chip models, and automated bioassays without complex inertial dynamics. For instance, these flows facilitate controlled particle focusing or mixing in physiological mimicry, where the quasi-steady parabolic profile ensures predictable transport with minimal energy dissipation from unsteadiness.

High Womersley Number

In pulsatile flows where the Womersley number α exceeds 10–20, inertial forces dominate over viscous effects, resulting in a velocity profile that approximates a uniform plug flow across most of the pipe cross-section. Viscous influences are confined to a thin Stokes boundary layer near the wall, with thickness δ ≈ √(2ν/ω), where ν is the kinematic viscosity and ω is the angular frequency; this layer thickness becomes negligible relative to the pipe radius R as α increases, since δ/R ≈ √(2/α²). The exact velocity profile, derived from the solution involving Bessel functions of the first kind, can be approximated for high α using asymptotic expansions that highlight the near-uniform core flow and the oscillatory boundary layer behavior. A key characteristic of high-α flows is the substantial phase lag between the pressure gradient and the bulk velocity, approaching 90° in the inertial limit, which arises from the dominance of fluid acceleration over diffusion. This lag reflects the flow's response being primarily governed by unsteady inertia, with the velocity waveform shifting relative to the driving pressure oscillation. These conditions lead to enhanced pressure wave propagation along the flow domain, as viscous damping is minimized outside the thin boundary layer, reducing overall energy dissipation compared to low-α regimes. Such behavior is physiologically relevant in large arteries like the aorta, where α values of 10–15 promote wave-like pulse transmission essential for efficient blood distribution. In engineering contexts, high-α pulsatile flows occur in high-frequency applications, such as acoustic streaming or oscillatory pumps, where the plug-like motion simplifies design predictions for momentum transfer. Recent studies highlight the role of high in influencing turbulent transitions within , where the thin can trigger instability mechanisms distinct from steady flows, delaying or altering the onset of turbulence. This is particularly pertinent to 2025 aerospace developments, including pulsating flows in advanced propulsion systems and hypersonic inlets, where understanding these transitions optimizes efficiency and reduces drag in unsteady operating conditions.

Derivations

Flow Rate

In pulsatile flow through a rigid cylindrical tube, the volumetric flow rate Q(t) is derived by integrating the axial velocity profile u(r, t) over the tube cross-section, yielding Q(t) = \int_0^R u(r, t) \, 2\pi r \, dr. For a sinusoidal pressure gradient -\frac{\partial p}{\partial z} = A e^{i \omega t}, where A is the complex amplitude, \omega is the angular frequency, and the real part is taken for physical quantities, the resulting expression is Q(t) = \operatorname{Re} \left[ \frac{\pi R^2 A}{i \omega \rho} \left( 1 - \frac{2 J_1(\alpha \sqrt{i})}{\alpha \sqrt{i} \, J_0(\alpha \sqrt{i})} \right) e^{i \omega t} \right], where R is the tube radius, \rho is the fluid density, \alpha = R \sqrt{\omega \rho / \mu} is the Womersley number (\mu is dynamic viscosity), J_0 and J_1 are Bessel functions of the first kind of orders zero and one, respectively, and i = \sqrt{-1}. This derivation assumes incompressible, Newtonian fluid, axisymmetric laminar flow, and no-slip boundary conditions at the wall. The formula can be interpreted through the concept of complex admittance Y(\omega), defined as the ratio of the complex flow rate amplitude to the pressure gradient amplitude, Y(\omega) = \hat{Q} / A, where |\hat{Q}| / |\hat{Y}| gives the magnitude and the phase difference reflects the lag between pressure and flow. The admittance magnitude and phase depend critically on \alpha: for low \alpha (viscous-dominated), it approaches the steady Poiseuille value \pi R^4 / (8 \mu); for high \alpha (inertia-dominated), the magnitude scales as $2 / (\alpha^2 \omega \rho / \mu), with a phase lag approaching 90 degrees. In applications, such as biomedical flow measurement devices, the flow rate formula enables non-invasive estimation of pulsatile blood volume transport using ultrasound or magnetic resonance techniques calibrated to . In elastic tubes, vessel compliance introduces wave propagation effects that modify the admittance, reducing peak flow rates and altering phase by up to 20-30% compared to rigid cases, as seen in arterial models. The mean flow rate over one cycle, \overline{Q} = \frac{1}{T} \int_0^T Q(t) \, dt, is zero for purely oscillatory pressure without a steady component but plays a key role in net transport in physiological systems like arteries, where a superimposed steady provides the average cardiac output (e.g., 5 L/min in humans). This separation ensures forward net propulsion despite oscillatory reversals near the wall.

Wall Shear Stress

In pulsatile flow through a rigid cylindrical pipe, the wall shear stress \tau_w(t) represents the tangential frictional force per unit area exerted by the fluid on the pipe wall, which is crucial for quantifying energy dissipation and biomechanical influences in applications such as arterial blood flow. It is defined as \tau_w(t) = \mu \left. \frac{\partial u}{\partial r} \right|_{r=R}, where \mu is the dynamic viscosity, u(r,t) is the axial velocity, r is the radial coordinate, and R is the pipe radius. For the Womersley solution assuming a sinusoidal pressure gradient - \frac{\partial p}{\partial z} = \mathrm{Re} \left[ A e^{i \omega t} \right] with amplitude A and angular frequency \omega, the velocity profile yields the wall shear stress through evaluation of the radial gradient at the wall. The explicit expression for the wall shear stress is derived as \tau_w(t) = \mathrm{Re} \left[ -\frac{A \mu}{i \omega \rho} \cdot \frac{\alpha \sqrt{i} \, J_1(\alpha \sqrt{i})}{R \, J_0(\alpha \sqrt{i})} \, e^{i \omega t} \right], where \rho is the fluid density, \alpha = R \sqrt{\omega \rho / \mu} is the , i = \sqrt{-1}, and J_0 and J_1 are of orders zero and one, respectively. This formula arises from differentiating the complex velocity profile u(r,t) = \mathrm{Re} \left[ \frac{A}{i \omega \rho} \left( 1 - \frac{J_0(\alpha \sqrt{i} \, r/R)}{J_0(\alpha \sqrt{i})} \right) e^{i \omega t} \right] and applying the \frac{d}{ds} J_0(s) = -J_1(s), evaluated at r = R. The negative sign accounts for the conventional direction of shear in pipe flow, with the real part extracting the physical oscillatory component. This wall shear stress exhibits distinct oscillatory characteristics that depend strongly on the Womersley number \alpha. For low \alpha (quasi-steady flow), \tau_w(t) closely follows the pressure gradient waveform, resembling steady Poiseuille flow with \tau_w \approx - \frac{A R}{2} \cos(\omega t). At high \alpha > 10, typical of large arteries, a thin viscous forms near , leading to of the shear stress by a factor approaching \alpha / \sqrt{2} compared to the quasi-steady case, due to the concentrating within the Stokes layer thickness \delta \approx \sqrt{2 \mu / (\rho \omega)}. Additionally, \tau_w(t) leads the by a phase up to $90^\circ at high \alpha, reflecting the inertial dominance in the core flow while viscous effects lag near . These features highlight the transition from viscous- to inertia-dominated s, influencing frictional losses in and hemodynamic applications. In biological contexts, particularly flow, wall modulates endothelial cell responses and atherogenesis. Physiological levels of laminar (10–20 dyn/cm² or 1–2 ) promote endothelial production and , maintaining vascular health. Conversely, low time-averaged wall (TAWSS < 4 dyn/cm²) or highly oscillatory (oscillatory shear index > 0.15) disrupts endothelial alignment and , upregulating molecules and promoting infiltration, which initiates plaque formation in regions like arterial bifurcations. The time-averaged wall in pulsatile arterial flow equals the steady value for the mean , \overline{\tau_w} = 4 \mu \overline{Q} / (\pi R^3), but local variations due to pulsatility correlate with cardiovascular risk; recent patient-specific computational models from the link regions of low TAWSS (< 0.4 ) and prolonged particle to accelerated plaque progression and rupture vulnerability, informing risk stratification beyond traditional factors like LDL .

Centerline Velocity

The centerline velocity in pulsatile pipe flow, denoted as u_\text{cl}(t) = u(0, t), represents the peak axial speed along the pipe's central axis and exhibits distinct temporal oscillations driven by the periodic . This velocity is derived from the analytical solution to the Navier-Stokes equations for in a rigid cylindrical , assuming a harmonically varying -\frac{\partial p}{\partial x} = \Re \left[ A e^{i \omega t} \right], where A is the complex , \omega is the , and t is time. The full expression for the centerline is u_\text{cl}(t) = \Re \left[ \frac{A}{i \omega \rho} \left( 1 - \frac{1}{J_0(\alpha \sqrt{i})} \right) e^{i \omega t} \right], where \Re denotes the real part, \rho is the fluid density, i = \sqrt{-1}, \alpha = R \sqrt{\omega / \nu} is the Womersley number with pipe radius R and kinematic viscosity \nu, and J_0 is the zeroth-order Bessel function of the first kind with complex argument. For high Womersley numbers (\alpha \gg 1), corresponding to high frequencies or large diameters where inertial effects dominate, the magnitude of J_0(\alpha \sqrt{i}) becomes large, rendering the Bessel ratio term negligible. The centerline then approximates to u_\text{cl}(t) \approx \Re \left[ \frac{A}{i \omega \rho} e^{i \omega t} \right], indicating a plug-like profile near the centerline with essentially from viscous effects. This approximation underscores that the amplitude of u_\text{cl} is independent of at high frequencies, scaling inversely with \omega and proportional to the amplitude, as the oscillatory layer confines to a thin near . A key characteristic of the centerline velocity is its relationship with the driving , which shifts from nearly in-phase at low \alpha (viscous-dominated, quasi-steady Poiseuille-like behavior) to leading by up to 90° at high \alpha, reflecting the of the acceleration term in the momentum where inertial response precedes viscous . This phase lead arises because $1/(i \omega) = -i / \omega, introducing a \pi/2 shift in the inviscid limit. In cardiovascular diagnostics, measurements of centerline via Doppler ultrasound provide critical insights into pulsatile blood flow, enabling non-invasive assessment of arterial ; the Womersley model facilitates reconstruction of full profiles and flow rates from these single-point centerline data. Modern numerical simulations of transient startup scenarios, where the pulsatile is abruptly imposed from rest, reveal an overshoot phenomenon in the centerline : during the initial cycles, u_\text{cl} temporarily exceeds its eventual periodic amplitude before decaying to the steady oscillatory state, driven by the interplay of inertial buildup and incomplete viscous adjustment absent in the idealized periodic Womersley assumption.

References

  1. [1]
    Pulsatile Flow - an overview | ScienceDirect Topics
    In pulsatile flow, the fluid flow is periodically varied in magnitude or direction, creating a series of pressure pulses that propagate through the fluid.
  2. [2]
    [PDF] Oscillatory and pulsatile flows in environmental, biological ... - HAL
    In the circulatory system, the study of the pulsatile flow of blood is indispensable for the better comprehension of many cardiovascular diseases.
  3. [3]
    The role of the Womersley number in pulsatile blood flow
    The purpose of this Note is to clarify the meaning of the Womersley number α in pulsatile blood flow in small vessels.
  4. [4]
    Womersley's solution for the measurement of volume flow rates in ...
    Nov 14, 2022 · The characterization of transient flows within the Reynolds number range of Re = 10–100 and the Womersley number range α = 0.8–10 is ...
  5. [5]
    [PDF] 1 Pulsatile Flow in Microfluidic Systems - Sauret Lab
    Pulsatile flows can be used for mimicking physiological systems, to alter or enhance cell cultures, and for bioassay automation. Pulsatile flows offer unique.
  6. [6]
    Pulsatile flow in elastic tubes: A literature review of experimental ...
    Oct 27, 2025 · Pulsatile flow, described by oscillating velocity and pressure fields, is a key fluid dynamic phenomenon observed in both biological and ...Missing: definition | Show results with:definition
  7. [7]
    Method for the calculation of velocity, rate of flow and viscous drag in ...
    Method for the calculation of velocity, rate of flow and viscous drag in arteries when the pressure gradient is known. J. R. Womersley,. J. R. Womersley.Missing: pulsatile | Show results with:pulsatile
  8. [8]
    [PDF] navier-stokes solutions, cylindrical coordinates
    Do an example problem in cylindrical coordinates – fully developed laminar pipe flow. Continuity and Navier-Stokes Equations, Vector Forms.
  9. [9]
    Exact Solutions of the Navier–Stokes Equations - SpringerLink
    Sep 7, 2022 · We integrate the Navier–Stokes equations expressed in a cylindrical coordinate system. The vector velocity has components v_r, v_{\theta }, and ...3.2 Axisymmetric Stationary... · 3.4 Axisymmetric Transient... · 3.6 Pipe Flow<|control11|><|separator|>
  10. [10]
    [PDF] method for the calculation of velocity, rate of flow and viscous drag in ...
    The method calculates velocity, flow rate, and viscous drag in arteries when the pressure gradient is known, using a formula similar to Lambossy's.
  11. [11]
    [PDF] An exact solution of Navier-Stokes equations for the flow through a ...
    The Navier-Stokes equations and the equation of continuity that govern the motion of blood, in the cylindrical polar coordinates system can be written as: 0.
  12. [12]
    Chaotic flows in pulsating cylindrical tubes: a class of exact Navier ...
    Apr 28, 2003 · The Navier–Stokes equations admit solutions which are governed by nonlinear partial differential equations depending on the radial ...
  13. [13]
    A review on non-Newtonian fluid models for multi-layered blood ...
    Jan 31, 2023 · The Casson fluid model, in particular, more properly describes blood flow characteristics at minimal shear rates, during which it moves via ...
  14. [14]
    Non-Newtonian perspectives on pulsatile blood-analog flows in a ...
    Jul 8, 2015 · It is demonstrated that non-Newtonian, blood-analog fluid rheology results in shear layer instabilities that alter the formation of vortical ...
  15. [15]
    Modeling Viscoelastic Behavior of Arterial Walls and Their ...
    We study the flow of an incompressible viscous fluid through a long tube with compliant walls. The flow is governed by a given time-dependent pressure head ...
  16. [16]
    [PDF] Bio Fluid Mechanics: Lecture 14 - Imperial College London
    Navier-Stokes equations ρ. ∂u. ∂t= G(t) + µ∇. 2 u with no slip ... In a physiological context, α is known as the Womersley number, but similar parameters.
  17. [17]
    [PDF] Analysis of Reynolds, Strouhal and Womersley numbers in ... - ISMRM
    Table 1: Dimensionless numbers for pulsatile flow. Reynolds number. Womersley number. Strouhal number inertial forces / viscous forces unsteady forces ...Missing: parameters | Show results with:parameters
  18. [18]
    [PDF] The Use of the Dimensionless Womersley Number to Characterize ...
    Because of its appearance in the Navier-. Stokes equations, Wo is such a qualitative indicator for unsteady flow behavior. Before associations may be made ...
  19. [19]
    The effect of Dean, Reynolds and Womersley numbers on the flow in ...
    ... Womersley number, Wo, are investigated: 2.6 and 3.6. The Womersley number is defined as W o = 0.5 D P V 2 π / T v , where T is the length of the cardiac ...
  20. [20]
    The effect of Dean, Reynolds and Womersley numbers on the flow in ...
    Mar 31, 2021 · Pulsatile flows in a straight pipe have also been widely studied, and theoretical analysis showed that they depend on a single non-dimensional ...3.1. Steady Flow · 3.1. 1. Flow Field Topology · 4. Unsteady FlowMissing: dimensionless | Show results with:dimensionless
  21. [21]
    Literature Survey for In-Vivo Reynolds and Womersley Numbers of ...
    In pulsatile flow, velocity distribution and wall shear stress depend on compliance, and the Reynolds and Womersley numbers. However, matching such values may ...
  22. [22]
    Method for estimating pulsatile wall shear stress from one ...
    Apr 17, 2023 · 3.1 Validation of pulsatile wall shear stress predictions. Here the Poiseuille and Womersley flow models are first compared using experimentally ...
  23. [23]
    Finite Element Modelling of Pulsatile Blood Flow in Idealized ... - NIH
    Oct 17, 2011 · A three-dimensional computer model of human aortic arch with three branches is reproduced to study the pulsatile blood flow with Finite Element ...
  24. [24]
    The assignment of velocity profiles in finite element simulations of ...
    In this paper we present a new method for the assignment of pulsatile velocity profiles as input boundary conditions in finite element models of arteries.
  25. [25]
    [PDF] AN OVERVIEW ON PULSATILE FLOW DYNAMICS
    The critical magnitudes of. Womersley number, √ which is a non-dimensional frequency parameter are such that below √ =1.32 quasi-steady region and above √ =28 ...
  26. [26]
    Method for estimating pulsatile wall shear stress from one ... - NIH
    Apr 17, 2023 · This “quasi‐steady” approximation assumes frequency content is negligible and refers to reconstruction of transient flow profiles with ...
  27. [27]
    Pulsatile pipe flow experiment to study particle–fluid interactions ...
    Jul 11, 2025 · In this study, we present comprehensive three-dimensional measurements of particle-laden flow in a novel pipe flow experiment.
  28. [28]
    Pulsatile Flow in Microfluidic Systems - Dincau - Wiley Online Library
    Oct 28, 2019 · Pulsatile flows can be used for mimicking physiological systems, to alter or enhance cell cultures, and for bioassay automation.
  29. [29]
    Pulsatile blood flow in the entire coronary arterial tree: theory and ...
    The two major determinants of coronary flow pulsatility are 1) the pulsatile aortic pressure, and 2) the myocardial/vascular interaction in systole. To ...
  30. [30]
  31. [31]
    5.6 Pulsatile flow in straight tubes - Cardiovascular biomechanics
    Velocity profiles for various Womersley numbers (α=2,4,8,16). 5.6.2 Wall shear stress for pulsatile flow in straight tubes. By using the property of the Bessel ...
  32. [32]
    Wall shear stress and its role in atherosclerosis - PMC - NIH
    Apr 3, 2023 · The initiation of plaque formation has been strongly linked to a hemodynamic parameter: wall shear stress (WSS) (4). WSS is the frictional force ...
  33. [33]
    Expert recommendations on the assessment of wall shear stress in ...
    Wall shear stress has a major impact on endothelial function and therefore plays a key role in atherosclerotic disease development and in long-term evolution ...The biological relevance of... · Wall shear stress in native... · Clinical applications
  34. [34]
    The Story of Wall Shear Stress in Coronary Artery Atherosclerosis
    Dec 14, 2020 · Wall shear stress (WSS) affects coronary artery atherosclerosis by inducing endothelial cell mechanotransduction and by controlling the near-wall transport ...
  35. [35]
    Wall shear stress and relative residence time as potential risk factors ...
    Mar 18, 2022 · Wall shear stress and relative residence time as potential risk factors for abdominal aortic aneurysms in males: a 4D flow cardiovascular ...Data Acquisition And... · Wss Analysis · Aaa Patients Vs Elderly...
  36. [36]
    [PDF] Womersley flow Description of the problem - BioMedMath
    The equation (2)1 is nonstationary (i.e., evolution) Stokes equation; in this special case of the flow through a pipe (pressure at the pipe's ends is constant ...
  37. [37]
    Application of Womersley Model to Reconstruct Pulsatile Flow From ...
    The use of the Womersley's rigid-tube approximation in pulsatile tube flow to refine and reconstruct raw US measurements has been relatively successful.
  38. [38]
    Effects of transients on pulsatile flow in arteries - PubMed
    In this paper, we compare the Womersley and initial-boundary value solutions for model transients that stop then restart, explain these previously unreported ...Missing: overshoot startup