Hemodynamic Simulations Workbench: An Open-Source JavaScript Framework for Pulmonary Artery Pressure Modeling
This workbench delivers an intuitive, browser-based environment for simulating pulmonary artery pressure (PAP) models, turning intricate hemodynamic equations into interactive visualizations. Built entirely in JavaScript and released as open source, it lets clinicians and researchers adjust model coefficients in real time and immediately see the impact on predicted PAP, thereby bridging the gap between theoretical hemodynamics and bedside decision-making. By prioritizing clarity, ease of use, and community extensibility, the platform embodies George Box’s reminder that “all models are wrong, but some are useful,” providing a practical tool that makes complex vascular dynamics both accessible and clinically meaningful.
Graph vs Simulation: Why Simulations Are Essential for Hemodynamic Research
Traditional graph drawing, while effective in providing a clear visual representation of data, often falls short in addressing the dynamic and complex nature inherent in systems like hemodynamics. This is precisely where the power of simulations, especially hemodynamic models, becomes evident. Unlike static graphs, which display a fixed dataset, simulations of hemodynamic models offer a dynamic platform that enables researchers to interact with the model, adjust parameters, and observe how these changes influence the system in real-time. This level of interactivity is pivotal for understanding the impact of various physiological conditions on blood flow and heart function, offering insights that are both detailed and predictive.
Simulations excel in their predictive capabilities, allowing researchers to forecast cardiovascular behaviors under conditions that have not been previously observed. This is particularly invaluable in medical fields, where accurately predicting outcomes can guide the planning of surgical interventions or the design of therapeutic devices. Moreover, hemodynamic simulations can integrate complex fluid dynamics and biological interactions through advanced computational methods, providing a depth of realism and detail far surpassing what can be achieved through simple graph drawing. This includes modeling responses to physical activities, drug interactions, or therapeutic measures, thereby not only describing but also predicting physiological responses.
The interactivity offered by simulations supports experimental design and testing within a controlled virtual environment, which can dramatically reduce the costs, time, and ethical concerns associated with direct physical experimentation. In contrast, traditional graphs, once created, do not allow for modifications to test different scenarios or predict the outcomes of variable changes, limiting their utility in dynamic and complex areas such as cardiovascular health.
Adding to the clinical relevance, hemodynamic models help bridge the gap between clinical issues and theoretical models through effective visualization. For instance, the development of logistic-based equations for estimating Pulmonary Artery Pressure (PAP), based on hemodynamic models (Frank, 2018), illustrates how simulations can turn complex equations into accessible insights for clinicians. Such tools are essential not only for their ability to simulate but also for their user-friendliness in allowing clinicians to derive meaningful interpretations from complex data. Therefore, while traditional graph drawing remains useful for straightforward presentations, simulations are indispensable for deeper understanding, innovation, and advancement in dynamic and complex fields like hemodynamics, enhancing both the exploration and explanation of intricate medical phenomena.
Browser Cache Refresh & DevTools Shortcuts
-
Chrome (Windows):
- Ctrl + Shift + I – Open Developer Tools
- F12 – Alternate shortcut to open Developer Tools
- Ctrl + F5 – Hard reload (bypass cache)
- Ctrl + Shift + R – Hard reload (bypass cache)
-
Chrome (macOS):
- ⌘ + ⌥ + I – Open Developer Tools
- ⇧ + ⌘ + R – Hard reload (bypass cache)
-
Firefox (Windows):
- Ctrl + Shift + I – Open Developer Tools
- F12 – Alternate shortcut to open Developer Tools
- Ctrl + F5 – Hard reload (bypass cache)
- Ctrl + Shift + R – Hard reload (bypass cache)
-
Firefox (macOS):
- ⌘ + ⌥ + I – Open Developer Tools
- ⇧ + ⌘ + R – Hard reload (bypass cache)
-
Safari (macOS):
- First, enable the Develop menu:
- Safari → Preferences → Advanced → Check “Show Develop menu in menu bar.”
- ⌘ + ⌥ + I – Open Web Inspector (Developer Tools)
- Option + ⌘ + E – Empty Caches
- ⌘ + R – Reload page (standard)
- ⇧ + Click Reload (or ⌘ + R after emptying caches) – Hard reload (bypass cache)
de Laval Nozzle Model
Hemodynamic Modeling Based on de Laval Nozzle Theory (Written June 21, 2025)
The de Laval nozzle is a classic convergent-divergent flow passage known for accelerating compressible fluids (such as gases) to high velocities. It consists of three distinct sections: a converging inlet, a narrow throat, and a diverging outlet. In theory, as gas flows through a de Laval nozzle under steady conditions, it accelerates to sonic speed at the throat and can further accelerate to supersonic speeds in the diverging section if sufficient pressure drop is available. In cardiovascular hemodynamics, blood flow in certain regions—especially around heart valves—exhibits geometric and functional similarities to the convergent–throat–divergent pattern of a de Laval nozzle. This document examines the theoretical relevance of de Laval nozzle models to blood flow, focusing on the subvalvular (converging), valvular (throat), and supravalvular (diverging) sections of cardiovascular flow. The discussion covers three conceptual mappings of the nozzle model to hemodynamics (analogy, surrogate modeling, and direct integration), the quasi-steady assumptions used to adapt compressible flow concepts to pulsatile incompressible blood flow, and the limitations of applying such models given physiological realities.
Mapping the de Laval Nozzle Concept to Cardiovascular Flow
In cardiovascular fluid dynamics, the flow path from the heart through a valve and into the arteries can be conceptually mapped to a convergent–divergent nozzle configuration. Three levels of mapping can be considered, each progressively more concrete: (1) a qualitative analogy, (2) a surrogate modeling approach, and (3) a directly integrated structural model. These mappings envision the subvalvular region as a convergent inlet, the valvular orifice as a narrow throat, and the supravalvular region (e.g., the proximal aorta) as a divergent diffuser. In all cases, the goal is to leverage the well-understood physics of nozzle flow to gain insight into or simplify the description of blood flow dynamics through cardiac valves or arterial stenoses.
-
Analogy: The Cardiovascular Nozzle
As an analogy, the de Laval nozzle serves as a descriptive tool to illustrate how blood accelerates and decelerates through varying cross-sectional areas in the cardiovascular system. Consider the left ventricular outflow tract and aortic valve: during systole, blood is ejected from the left ventricle through the outflow tract (subvalvular region) which generally tapers toward the aortic valve. This tapering is akin to a converging nozzle section, causing the blood velocity to increase as the flow area narrows. The aortic valve orifice itself represents the throat of the nozzle – the point of minimum cross-sectional area – where the blood velocity reaches its peak and the pressure is at its lowest. Beyond the valve, the blood enters the ascending aorta and aortic root, which often expand in diameter (supravalvular region) analogous to a diverging nozzle section. Here, the blood flow slows down as the area increases, and some of the static pressure is recovered.
Through this analogy, key hemodynamic phenomena become easier to visualize. For instance, the significant pressure drop across a stenotic (narrowed) valve can be understood via the Venturi effect: as blood speeds up in the tight orifice, static pressure must decrease. Likewise, the concept of vena contracta – the narrowest actual flow stream just downstream of a valve – mirrors the throat effect in a nozzle where the flow jet contracts and reaches maximum velocity. Downstream of a normal valve or mild stenosis (with a gradually expanding aortic root), the diverging geometry allows partial pressure recovery, similar to how a diffuser recovers pressure in a nozzle flow. This simple geometric analogy helps clinicians and researchers qualitatively reason about flow acceleration, jet formation, and pressure changes in conditions like aortic stenosis or other valvular lesions.
Analogy Highlight:
Just as gas in a de Laval nozzle accelerates through a narrowed throat and then decelerates in the expansion, blood flow picks up speed as it passes through a constricted valve and slows down as the vessel widens downstream. The converging–throat–diverging structure is a useful mental model for understanding where velocities peak and pressures drop in cardiovascular flow.
-
Surrogate Modeling Tool: Simplified Nozzle-Based Calculations
Beyond qualitative description, the de Laval nozzle concept can be used as a surrogate modeling tool in theoretical hemodynamics. In this context, surrogate modeling means using the nozzle’s simplified physics and equations to approximate or represent the behavior of blood flow without directly simulating the full complexities of cardiovascular fluid dynamics. For example, engineers and physiologists might treat a heart valve or arterial stenosis as an “equivalent nozzle” to estimate flow rates and pressure drops. By applying conservation of mass and energy as one would in a nozzle, one can derive relationships such as the continuity equation and Bernoulli’s principle to link flow velocity, cross-sectional area, and pressure difference across a constriction.
In practice, this surrogate approach underlies many diagnostic and modeling techniques. A prime example is the use of the Bernoulli equation to estimate the pressure gradient across a stenotic valve from measured blood velocity. In echocardiography, the maximal jet velocity \( v_{\text{max}} \) through a narrowed valve is used to approximate the pressure drop \( \Delta p \) via the simplified relation
\( \Delta p \approx \tfrac12 \rho\, v_{\text{max}}^{2} \)
(or, for \( \rho \approx 1 \, \mathrm{g\,cm}^{-3} \), the clinical shorthand
\( \Delta p \approx 4 v^{2} \)).
This formula is essentially a Venturi-nozzle calculation assuming inviscid, quasi-steady flow. Similarly, when determining the effective orifice area of a valve, clinicians apply the continuity equation (an incompressible-flow analog to mass conservation in a nozzle):
\( A_{\text{valve}} = \dfrac{Q}{v_{\text{valve}}} \),
where \( Q \) is the flow rate and \( v_{\text{valve}} \) is the blood velocity at the valve. Such computations treat the valve region as if it were a nozzle throat connecting two larger reservoirs (the ventricle and the aorta) and yield surprisingly accurate results for many purposes.
Researchers also incorporate nozzle-like elements in analytical models or lumped-parameter simulations of the circulation. For instance, an arterial stenosis might be modeled as a short converging–diverging segment in a one-dimensional flow model, using equations from nozzle theory to capture the drop in pressure and change in velocity. This surrogate modeling is valuable because it simplifies complex 3-D fluid dynamics into tractable equations while still capturing the first-order effects of area change on flow behavior. However, it typically requires calibration or empirical correction (such as discharge coefficients or pressure-recovery factors) to account for differences between the ideal nozzle theory and real blood flow (e.g., energy losses due to viscosity and turbulence).
-
Direct Integration as a Structural Model Component
The third conceptual mapping involves directly integrating the de Laval nozzle structure into the theoretical framework of cardiovascular modeling. Rather than using the nozzle concept only as an analogy or a separate calculation tool, this approach builds a model in which the geometry and equations of a convergent–throat–divergent nozzle are embedded as part of the cardiovascular system’s structure. In other words, the section of the circulation containing a valve or stenosis is explicitly modeled as a tapered tube with a narrow throat, and the flow equations are applied accordingly along that segment.
For example, consider a one-dimensional (1D) model of blood flow in the aorta that includes the aortic valve. Using a de Laval nozzle basis, one can represent the left ventricular outflow tract, valve, and proximal aorta as a continuous channel with varying cross-sectional area \( A(x) \): decreasing from the ventricle toward the valve and increasing beyond it. By applying the Euler or Navier–Stokes equations in 1D along this channel, one directly leverages nozzle flow dynamics to compute how pressure and velocity evolve from the subvalvular region, through the valvular throat, and into the supravalvular region. In the limit of steady, inviscid flow with constant density, these equations reduce to familiar nozzle relationships (mass-flow conservation \( A\,v = \text{const} \) and energy conservation akin to Bernoulli’s law). Even when extended to unsteady or viscous conditions, the nozzle-shaped structural model provides a foundation to incorporate additional effects (like unsteady inertial terms or viscous losses) in a spatially distributed manner.
Direct integration of the nozzle model is particularly useful for investigating theoretical scenarios or boundary conditions that are otherwise difficult to capture with simple lumped models. For instance, one can examine how a given valve area and aortic geometry together influence the peak achievable flow rate or the pressure pulse waveform by solving the flow governing equations in this structured geometry. This approach bridges the gap between a purely conceptual analogy and a full computational fluid dynamics simulation: it uses the structure of the nozzle to simplify the domain while retaining the essential spatial variability of area that drives the hemodynamics. It should be noted, however, that building such a model still demands careful attention to the realism of assumptions (e.g., whether to treat the fluid as incompressible, how to include wall flexibility, etc.), and often hybridizes the ideal nozzle theory with empirical corrections to remain physiologically accurate.
Quasi-Steady Assumptions for Pulsatile Flow Adaptation
Adapting compressible nozzle models to cardiovascular flow requires acknowledging two major differences: blood is effectively incompressible (density does not change appreciably with pressure), and blood flow is pulsatile (unsteady) rather than strictly steady. The concept of quasi-steady flow is often invoked to justify using steady-state relationships from compressible flow in the transient, incompressible context of the cardiovascular system. A quasi-steady assumption means that at each instant of time, the flow behaves as if it were in a steady state corresponding to the instantaneous conditions, neglecting the effects of fluid acceleration or history. This assumption is reasonable if the flow’s time-varying nature is slow enough that the fluid has time to adjust to near-steady conditions within each phase of the cycle.
In practice, the quasi-steady approach in hemodynamics allows one to apply formulas like the Bernoulli equation or continuity equation instantaneously during a heartbeat. For example, when blood accelerates through a valve during systole, a quasi-steady model assumes that the relationship between pressure and velocity at that moment is given by the steady Bernoulli principle \( \Delta p = \tfrac12 \rho\, v^{2} \) for the instantaneous velocity even though the velocity is changing over time. The justification lies in the timescales: if the cardiac cycle frequency is moderate and the Womersley number (a dimensionless number quantifying unsteady flow effects) is not too high, inertial and unsteady boundary-layer effects may be small compared to the dominant convective acceleration and pressure drop. Under those conditions, the flow at each time behaves nearly like a snapshot of a steady nozzle flow.
Using quasi-steady assumptions, one effectively borrows the compressible-flow nozzle framework by treating blood as a nearly incompressible fluid (Mach number \( M = v/c \approx 0 \)) and by treating each phase of the pulsatile cycle independently. The compressibility aspect is removed by assuming constant density, which simplifies the nozzle equations to their incompressible form. The time-dependent aspect is handled by updating the flow parameters (e.g., valve area, instantaneous flow rate) as functions of time while still using steady relations at each time step. In essence, the compressible-flow equations are reduced to an incompressible, time-parametrized form. This approach has been quite successful for capturing peak pressure drops and flow rates – for instance, predicting the maximum jet velocity through a stenotic orifice using the instantaneous pressure difference from ventricle to aorta. It also underlies many clinical indices where dynamic measurements are interpreted through steady formulas.
However, it is important to recognize that quasi-steady adaptation is an approximation. When flow changes rapidly (such as at the very onset of systole or during valve closure), unsteady inertial effects and fluid accumulation can produce pressure deviations that the steady formulas do not capture. For example, an accelerating flow requires extra pressure (an inertial term) beyond the steady Bernoulli prediction to account for the kinetic-energy buildup. Likewise, flow separation or reattachment can lag behind instantaneous flow changes. Despite these nuances, quasi-steady models remain a cornerstone in theoretical hemodynamics because they greatly simplify analysis. They allow the use of well-known nozzle and Venturi relationships in a pulsatile setting, provided one remains mindful of their approximate nature and corrects for unsteady discrepancies when necessary (e.g., via an added unsteady term or empirical factors at high-acceleration phases).
Theoretical and Physiological Limitations
While the de Laval nozzle model provides useful insights, several assumptions inherent in the nozzle theory do not hold exactly in the cardiovascular system. Applying this model to blood flow therefore comes with important limitations:
Aspect |
Ideal de Laval Nozzle Assumption |
Hemodynamic Reality |
Compressibility |
Fluid is compressible; density can drop as pressure drops (significant in gas flows). Flow can reach Mach 1 (choked flow) at the throat under sufficient pressure drop, and supersonic flow beyond. |
Blood is effectively incompressible (density nearly constant). No true choking occurs; flow rate can continue to increase with higher pressure gradient (until other factors like turbulence or cavitation intervene). Pressure waves travel fast (elastic walls give pulse wave velocity ~5–10 m/s), but flow itself remains subsonic with negligible density change. |
Viscosity and Energy Loss |
Ideal nozzle theory often assumes inviscid (or isentropic) flow for simplicity, ignoring friction and turbulence. Flow remains attached in the diverging section if design is optimal. |
Blood has non-negligible viscosity and can experience turbulence, especially downstream of a severe constriction. Energy losses occur due to viscous friction and turbulent eddies (e.g., in the post-stenotic region). Unlike the ideal nozzle, a significant portion of kinetic energy may be dissipated as heat or turbulent mixing, reducing pressure recovery. |
Wall Properties |
Nozzle walls are rigid and fixed in shape. The geometry (convergent angle, throat area, divergent angle) is static and predetermined. |
Vessels and heart valve structures are flexible and dynamic. Vessel walls are compliant (they expand and contract with pressure), and valve orifices change area over time (valve leaflets open and close). The effective geometry of the “nozzle” in vivo is not fixed, but varies with pressure and flow during the cardiac cycle, making the flow dynamics more complex than a static nozzle model. |
Flow Regime |
Typically analyzed under steady or quasi-steady conditions. In compressible flow, established formulas assume a steady-state flow or a controlled expansion process. |
Cardiac flow is pulsatile and undergoes acceleration and deceleration each heartbeat. True steady flow never actually occurs in vivo. Unsteady inertial forces and time-varying flow separation can lead to deviations from steady-state predictions. The quasi-steady approach is an approximation that may break down at high acceleration phases or high-frequency oscillations. |
Boundary Conditions |
Upstream and downstream conditions are usually simple: e.g., a large reservoir feeding the nozzle and discharge into a constant-pressure ambient environment. Reflections or feedback from downstream are often neglected in basic nozzle theory. |
In the circulation, the upstream “reservoir” (the ventricle) and downstream system (arterial tree) are both dynamic and coupled. Downstream pressure is not constant but varies with arterial compliance and peripheral resistance. Reflected waves from the arterial tree can travel back toward the valve, affecting local pressure and flow. These complex boundary interactions are beyond the scope of standard nozzle equations and require more sophisticated modeling. |
Geometry Simplification |
The nozzle is usually axisymmetric and smoothly contoured for optimal flow expansion (especially in engineering applications to avoid flow separation and shocks). |
Real cardiac and vascular geometries can be irregular or asymmetric. For example, a stenosis in an artery might not have a smooth gradual divergence; it could be an abrupt narrowing followed by a sudden expansion, leading to flow separation that a smooth nozzle model would not predict. Additionally, features like branch vessels (e.g., coronary arteries near the aortic valve or arterial bifurcations) break the simple tubular geometry assumption. |
These limitations highlight that while nozzle-based models can capture the core mechanism of how area variation influences flow, they oversimplify many aspects of cardiovascular hemodynamics. It is crucial to account for viscosity (e.g., by including resistance or loss coefficients), wall elasticity (by coupling the nozzle model with compliance elements), and unsteady effects (by including added mass or unsteady pressure terms) when trying to apply the de Laval nozzle concept quantitatively to blood flow. In many cases, hybrid models are used – for instance, combining Bernoulli’s equation with an empirical discharge coefficient to correct for energy losses, or using a nozzle model for the mean behavior and superimposing wave dynamics for pulse propagation.
To mitigate these differences when modeling blood flow using nozzle analogies, one can:
-
Include viscous losses:
Add resistance terms or discharge coefficients to represent energy dissipation due to viscosity and turbulence.
-
Include compliance:
Couple the model with elastic wall elements (e.g., Windkessel or spring-damper components) to simulate how vessel expansion buffers pressure changes.
-
Account for unsteady effects:
Incorporate unsteady inertial terms or use the unsteady form of Bernoulli’s equation to capture the additional pressure needed for accelerating blood and any phase lags in flow development.
-
Validate geometry assumptions:
Use empirical data or higher-fidelity simulations to adjust the model if the actual vessel geometry causes flow separations or asymmetries not predicted by an ideal nozzle shape.
Geometry and Function of a De Laval Nozzle
Figure: Diagram of a convergent–divergent (de Laval) nozzle, showing how flow velocity \(v\) increases from the converging inlet to the diverging outlet while pressure \(p\) and temperature \(T\) decrease along the length. The narrowest cross-section (the throat) is where the flow reaches sonic speed (Mach \(1\)). Upstream of the throat, in the converging section, the flow is subsonic and accelerates as the area decreases. Downstream of the throat, in the diverging section, the flow becomes supersonic and continues to accelerate as it expands into a larger area.
A de Laval nozzle is a tube that is pinched in the middle, consisting of a converging section followed by a diverging section. This hourglass-shaped geometry is carefully designed to accelerate a compressible fluid (gas) to high speeds, including supersonic velocities. In the convergent part of the nozzle, the cross-sectional area decreases toward the throat, causing the subsonic incoming gas to speed up (much like a Venturi tube). At the throat (the minimum area), the gas reaches sonic conditions (Mach \(M=1\)) under choked-flow conditions. Beyond this point, the diverging section expands the flow: the increasing area allows the gas to continue accelerating if it has reached Mach \(M=1\) at the throat, thereby transitioning into a supersonic flow regime (Mach \(M>1\)). The de Laval nozzle thus converts the fluid’s internal energy into directed kinetic energy. As the gas accelerates through the nozzle, its pressure and temperature drop significantly due to the isentropic expansion, while its velocity increases dramatically. This principle is exploited in applications such as rocket engines and turbines to produce high-velocity jets for thrust.
-
Converging Section (Subsonic Inflow)
The converging section of the nozzle is the upstream part where the cross-sectional area decreases toward the throat. In this region, the flow is subsonic (Mach < 1). For subsonic gases, a smaller area causes the flow to accelerate (since the mass flow is constant, the fluid must speed up to get the same mass through a tighter space). This is analogous to squeezing an incompressible fluid through a pipe contraction – the fluid speeds up – but with compressible gas there is the added effect of density decreasing as well. Throughout the convergent section, velocity increases, Mach number rises (approaching 1.0 at the throat), static pressure drops, and static temperature drops. The drop in pressure from the inlet to the throat provides the driving force that accelerates the gas. This phenomenon is essentially a manifestation of the Venturi effect in compressible flow: as the channel narrows, kinetic energy increases at the expense of pressure energy.
Near the inlet of the convergent section, the flow Mach number is low, so compressibility is minor and the pressure is relatively high. As the gas moves through the narrowing passage, it accelerates and its Mach number may increase from, say, \(M \approx 0.2\) or \(0.3\) up toward \(M \approx 1\) at the throat. Because \(M<1\), the relation
\( \frac{dA}{A} = (M^2 - 1)\frac{dv}{v} \)
tells us \(dA<0\) gives \(dv>0\), confirming the acceleration. The pressure drop in the convergent nozzle can be estimated from Bernoulli-like considerations (with compressibility corrections via the isentropic relations). By the time the gas reaches the throat, it has been compressed and accelerated enough that its Mach number is nearly 1 and the flow is choked if sonic conditions are achieved. In summary, the converging section acts to pre-accelerate the flow subsonically, preparing it to reach Mach 1 at the throat.
It is worth noting that if the pressure ratio across the nozzle (upstream total pressure vs. downstream pressure) is not large enough, the flow may remain entirely subsonic throughout and never choke. In that case, the nozzle simply behaves like a subsonic Venturi tube – flow accelerates in the converging part and then decelerates in the diverging part (with pressure recovering). Only when the flow actually reaches Mach 1 at the throat does the diverging section serve its supersonic acceleration role. In practical nozzle operation (such as rocket engines), the upstream pressure is sufficiently high compared to the exit pressure that the throat will choke. At choke, the throat pressure is fixed at the critical value (\( \approx 0.528,p_0 \) for air), and a further decrease in exit pressure cannot increase the mass flow through the throat – it will only allow expansion to supersonic speeds in the diverging section.
-
Nozzle Throat and Choked Flow
The throat is the minimum-area point of the nozzle and serves as the pivotal location for the flow behavior. Under design conditions, the throat is where the gas velocity reaches sonic speed \( M = 1 \). This is the condition of choked flow, meaning the mass flow has hit its maximum for the given upstream stagnation conditions. At the throat, the flow’s Mach number is exactly 1, and the local pressure, temperature, and density are the critical values \( p^{*}, T^{*}, \rho^{*} \) as given by the isentropic formulas above. The occurrence of choked flow effectively decouples the flow upstream of the throat from the conditions downstream – information (pressure disturbances) cannot travel upstream through a sonic point. In other words, once choked, the mass flow is controlled solely by upstream conditions (total pressure, temperature, throat area) and is insensitive to further drops in downstream pressure.
The physics at the throat can be understood by considering that as the flow accelerates toward Mach 1, the pressure gradient needed to continue accelerating the gas becomes very steep. To achieve exactly \( M = 1 \) at the throat, a certain minimum pressure ratio is required (the back pressure must be low enough). When this is met, the throat “locks in” sonic conditions. At Mach 1, the flow speed equals the local speed of sound, so the gas cannot respond to small downstream pressure fluctuations – the sonic point acts like a one-way barrier for pressure signals. This is why beyond a certain point, further lowering the pressure after the nozzle does not increase the flow rate; it only influences what happens in the diverging section.
In summary, the throat is where the critical flow occurs. Engineers often design the throat size \( A_* \) to set the desired mass-flow rate (since \( \dot{m} \) at choke is proportional to \( A_* \) and \( p_0 /\sqrt{T_0} \) as shown earlier). The flow properties at the throat are also useful reference points: for instance, the Mach 1 condition is used to start the expansion fan for supersonic-flow calculations in the diverging part. It is also important to ensure the nozzle is operated at or above the critical pressure ratio for choked flow; otherwise the nozzle won’t produce supersonic exit velocities. If the throat is not choked (i.e., \( M \lt 1 \) there), the nozzle behaves just like an orifice or Venturi with subsonic flow throughout, and the advantages of the de Laval design (producing supersonic jets) are not realized.
-
Diverging Section (Supersonic Expansion)
Downstream of the throat lies the diverging section, which opens up gradually. When the flow is choked at the throat \( M = 1 \) there, the gas entering the diverging section is at Mach just above 1 and supersonic expansion can occur. In the diverging section, the area \( A \) increases with distance, and since the flow is now supersonic \( (M \gt 1) \), the relation
\(\dfrac{dA}{A} = (M^{2} - 1)\,\dfrac{dv}{v}\)
tells us \( dA \gt 0 \) yields \( dv \gt 0 \): the supersonic flow accelerates further as the area widens. This is a counter-intuitive flip from subsonic behaviour – a bigger cross-section pushes a supersonic fluid to go faster. Physically, what happens is that the high-pressure, high-temperature gas that was choked at the throat now undergoes a rapid expansion into the larger volume of the diverging nozzle. This expansion to lower pressure causes a strong conversion of thermal energy into kinetic energy: the gas cools and its pressure drops while its velocity increases to supersonic levels.
Throughout the diverging section, Mach number keeps rising above 1. For example, it might go from \( M = 1 \) at the throat to \( M = 2, 3 \), or higher by the nozzle exit, depending on how much the area expands and how low the exit pressure is. Correspondingly, the static pressure \( p \) falls continuously along the expansion, and static temperature \( T \) also falls. The flow speed can reach extremely high values (for instance, in rocket nozzles, exhaust velocities on the order of 2–3 km/s are common, corresponding to Mach numbers of 5 or more, depending on gas properties). The pressure ratio between the nozzle inlet (stagnation pressure) and exit is the driving factor for how far the expansion goes; a larger pressure drop (pushing towards a vacuum at exit) allows more expansion and higher exit Mach.
If the nozzle is correctly expanded, the exit pressure will equal the ambient pressure. In many practical cases, especially for rockets at high altitude, the expansion may be under-expanded or over-expanded. Under-expanded means the gas could expand more (exit pressure is still above ambient), so the flow will continue to fan out outside the nozzle. Over-expanded means the expansion went too far (exit pressure is below ambient); in this case, a supersonic jet will experience compression shocks outside (or even inside) the nozzle to adjust the pressure upward to ambient. In extreme over-expansion, flow separation can occur inside the diverging section, causing a shock structure and potentially unsteady behaviour. Designers avoid severe over-expansion that might cause the flow to detach from the walls. Ideally, at the design condition, the diverging section expands the gas just enough that exit pressure matches ambient, yielding maximal thrust with a uniform supersonic flow at exit.
In summary, the diverging section of a de Laval nozzle transforms the sonic flow from the throat into a high-Mach supersonic jet by allowing it to expand against a falling pressure. The combination of the convergent section (to accelerate subsonically to \( M = 1 \)) and the divergent section (to accelerate supersonically beyond \( M = 1 \)) is what enables a de Laval nozzle to efficiently produce extremely fast flows. Thermodynamically, most of the enthalpy drop (and thus temperature drop) of the gas occurs in the diverging section, as this is where the bulk of internal energy is converted to kinetic energy. By the nozzle exit, the gas pressure is much lower than at the inlet, and the gas has cooled significantly (often to a fraction of its stagnation temperature), having expended its energy to achieve a high-velocity stream.
Isentropic Flow Governing Equations
-
Assumptions and One-Dimensional Flow
To analyze flow through a de Laval nozzle, we typically make the following ideal assumptions for simplicity:
- Steady, one-dimensional flow: The flow is in steady state (does not change with time) and is confined to essentially one dimension along the nozzle’s axis (flow properties are uniform across any cross-section).
- Ideal gas, isentropic process: The gas is treated as ideal, and the expansion through the nozzle is isentropic (no heat transfer and no friction or dissipative losses). Thus, entropy remains constant and the process is reversible and adiabatic.
- No external forces aside from pressure: The flow is horizontal along the nozzle axis, so gravity or other body forces are negligible; pressure forces and inertia dominate.
- Compressible flow: The gas velocity can reach high values (typically Mach $>0.3$), so density changes with pressure are significant and must be accounted for (unlike in incompressible flow).
Under these conditions, the flow can be modeled by the classical equations of compressible fluid dynamics, which relate how area variations influence velocity, density, pressure, and Mach number in the nozzle.
-
Continuity and Mass Flow Rate
Because the flow is steady and one-dimensional, the mass flow rate
\( \dot m \) remains constant at every cross-section of the nozzle.
The continuity equation expresses this as:
$$
\dot m \;=\; \rho\,A\,v \;=\; \text{constant}
$$
where \( \rho \) is the local gas density, \( A \) is the cross-sectional
area, and \( v \) is the flow velocity at that cross-section.
This means that as the area \( A \) changes along the nozzle, the
product \( \rho A v \) must remain constant. In sub-sonic flow a
reduction in \( A \) forces an increase in \( v \); unlike an
incompressible liquid, however, a gas also changes its density
\( \rho \) as it accelerates. The mass flow can be rewritten in terms
of Mach number \( M \) and the speed of sound \( a \) as
\( \dot m = \rho A M a \). For a perfect gas,
\( a = \sqrt{\gamma R T} \) (with \( \gamma \) the specific-heat ratio
and \( R \) the gas constant).
Choked flow.
There is a maximum mass flow rate that the nozzle can pass, achieved
when the flow at the throat becomes sonic (\( M = 1 \)).
Once choked, further lowering the downstream pressure cannot increase
\( \dot m \). Combining continuity with the isentropic relations for an
ideal gas gives the mass flux at the throat:
$$
\frac{\dot m}{A^{\*}}
\;=\;
\rho^{\*}\,a^{\*}
\;=\;
p_{0}\,
\sqrt{\frac{\gamma}{R\,T_{0}}}\,
\left(\frac{2}{\gamma+1}\right)^{\frac{\gamma+1}{2(\gamma-1)}},
$$
where \( A^{\*} \) is the throat area and \( p_{0}, T_{0} \) are the
stagnation pressure and temperature upstream. Quantities marked with
a superscript \( ^{\*} \) are evaluated at sonic (critical) conditions.
This expression shows that \( \dot m \) reaches its ceiling once the
throat is choked.
-
Area–Mach Number Relationship
One of the most important results for variable-area flow is the relationship between cross-sectional area and the Mach number of the flow. By differentiating the continuity equation and using the compressible-flow relations, one can derive a differential relation between area change and velocity (or Mach-number change):
$$
\frac{dA}{A} \;=\; (M^{2}-1)\,\frac{dv}{v}.
$$
- If \(M<1\) (subsonic flow), then \(M^{2}-1\) is negative, so \(dA/A\) and \(dv/v\) have opposite signs. This means a reduction in area (\(dA<0\)) causes \(dv>0\), i.e. velocity increases – a converging duct accelerates subsonic flow. Conversely, expanding the area slows down subsonic flow.
- If \(M>1\) (supersonic flow), \(M^{2}-1\) is positive, so \(dA/A\) and \(dv/v\) have the same sign. An increase in area (\(dA>0\)) now causes \(dv>0\), i.e. a diverging duct accelerates supersonic flow. A converging section would slow down a supersonic flow (which is how supersonic wind tunnels use a diverging diffuser to decelerate flow).
- If \(M=1\), the equation gives \(dA=0\) for any finite \(dv\); in fact, at exactly Mach 1, an infinitesimal area change would require an infinite change in velocity – this is the choke condition at the throat.
This behavior explains the convergent–divergent nozzle design: subsonic gas accelerates in the convergent part (to reach \(M=1\) at the throat), and then the supersonic gas accelerates further in the divergent part. The precise quantitative relation between area and Mach number for an isentropic flow can be derived by combining the continuity and energy equations. The result is an expression for the area ratio \(A/A^{\*}\) as a function of Mach \(M\) (where \(A^{\*}\) is the area when \(M=1\)):
$$
\frac{A}{A^{\*}}
\;=\;
\frac{1}{M}\,
\left[
\frac{2}{\gamma+1}\,
\Bigl(1+\tfrac{\gamma-1}{2}\,M^{2}\Bigr)
\right]^{\frac{\gamma+1}{2(\gamma-1)}}.
$$
This area–Mach-number relation comes from the isentropic-flow equations. It indicates that for a given area ratio \(A/A^{\*}\), there are two possible Mach solutions: one subsonic (\(M<1\)) and one supersonic (\(M>1\)). At the throat, \(A/A^{\*}=1\) and indeed \(M=1\). For \(A > A^{\*}\), the flow could be either subsonic or supersonic, but the nozzle’s upstream and downstream pressure conditions select the physically realized branch (subsonic acceleration in the first part, then the supersonic branch beyond the throat). Notably, as \(A/A^{\*}\) becomes very large in the divergent section, \(M\) on the supersonic branch increases toward a high value (the flow can approach very high Mach numbers if the pressure ratio allows). On the sub-sonic branch, as \(A\) increases (beyond the throat in a diffuser), \(M\) would decrease toward low values. In practice, the convergent–divergent nozzle operates by following the subsonic branch up to \(M=1\) at the throat, then transitioning to the supersonic branch in the expansion.
-
Pressure, Density, and Temperature Relations
The isentropic flow assumption provides important relations between pressure, density, temperature, and Mach number along the nozzle.
As the gas accelerates and expands, its static pressure \(p\) and static temperature \(T\) drop, converting internal energy into kinetic energy,
while the total (stagnation) values \(p_{0}\) and \(T_{0}\) remain constant (in an adiabatic reversible process).
The following relations tie the local Mach number \(M\) to the ratios of static properties to their stagnation (reservoir) values:
$$
\frac{p}{p_{0}}
\;=\;
\Bigl(1+\tfrac{\gamma-1}{2}\,M^{2}\Bigr)^{-\tfrac{\gamma}{\gamma-1}},
\qquad
\frac{\rho}{\rho_{0}}
\;=\;
\Bigl(1+\tfrac{\gamma-1}{2}\,M^{2}\Bigr)^{-\tfrac{1}{\gamma-1}},
\qquad
\frac{T}{T_{0}}
\;=\;
\Bigl(1+\tfrac{\gamma-1}{2}\,M^{2}\Bigr)^{-1}.
$$
These equations show, for example, that as \(M\) increases (due to acceleration in the nozzle),
the pressure ratio \(p/p_{0}\) decreases and the temperature ratio \(T/T_{0}\) also decreases.
At the throat (\(M=1\)) they reduce to the critical (sonic) ratios:
- \(p^{\*} = p_{0}\,\Bigl(\tfrac{2}{\gamma+1}\Bigr)^{\tfrac{\gamma}{\gamma-1}}\),
- \(T^{\*} = T_{0}\,\Bigl(\tfrac{2}{\gamma+1}\Bigr)\),
- \(\rho^{\*} = \rho_{0}\,\Bigl(\tfrac{2}{\gamma+1}\Bigr)^{\tfrac{1}{\gamma-1}}\),
(An asterisk denotes sonic conditions.) Thus, the pressure at the throat is a fixed fraction of the upstream total pressure
(e.g.\ for \(\gamma = 1.4\), \(p^{\*} \approx 0.528\,p_{0}\)), and similarly for temperature and density.
Throughout the convergent section the flow remains sub-sonic (\(M < 1\)), so pressure falls modestly and Mach rises toward 1.
Once the flow becomes supersonic in the diverging section, \(M\) grows beyond 1 and the static pressure and temperature drop more sharply with further expansion.
Because the flow is isentropic, the total values \(p_{0}\) and \(T_{0}\) stay constant from inlet to exit; all of the drop in static pressure and temperature
is converted into kinetic energy.
Hemodynamic Analogy in Cardiovascular Flows
The principles of convergent–divergent nozzle flow have interesting parallels in cardiovascular hemodynamics, especially when considering blood flow through regions of varying cross-section such as heart valves and arterial segments. In the circulatory system, blood is an incompressible fluid, and vessel geometry can create nozzle-like effects. Notably, the flow through a stenotic heart valve (or any narrowed orifice in the vasculature) is often compared to flow through a nozzle: the blood accelerates through the constriction and then decelerates downstream. We can draw an analogy between the nozzle sections and specific anatomical regions:
Subvalvular region (ventricular outflow tract or vessel upstream of a stenosis): This is analogous to the converging section of a nozzle. As blood approaches a narrow valve opening or arterial stenosis, it is funneled from a wider chamber or vessel into a smaller orifice. The cross-sectional area available for flow decreases, so the blood velocity increases as it nears the valve. For example, in aortic stenosis, blood in the left ventricular outflow tract (LVOT) might have a velocity on the order of \(1~\text{m/s}\) or less, but as it is squeezed into the small valve orifice, it accelerates markedly. The flow is still subsonic in the hemodynamic sense (blood is effectively incompressible and \(M\approx 0\) with respect to the speed of sound in tissue), but the acceleration is analogous to a gas accelerating in a convergent nozzle. Accompanying this velocity increase, there is a pressure drop from the ventricle to just before the valve (much like the pressure drop in a convergent nozzle) – this is observed clinically as the proximal pressure in the ventricle being higher than the pressure in the outflow tract or just at the valve entrance when flow is occurring.
Valvular orifice (the valve opening or stenotic throat): This is analogous to the nozzle throat. The valve opening represents the point of minimum effective flow area – often referred to as the effective orifice area (EOA) in valvular stenosis. Here, the blood achieves its maximum velocity as it passes through the tightest constriction. In a severe aortic stenosis, for instance, the jet velocity through the valve can reach \(4~\text{m/s}\) or more. The flow is characterized by a large pressure gradient across the valve: upstream (ventricular) pressure is much higher than downstream (aortic) pressure during ejection, which corresponds to the conversion of pressure into kinetic energy in the jet. While there is no literal “Mach 1” condition (blood flow never comes anywhere close to the speed of sound in liquid, which is ~1500 m/s), one can think of an analogous “critical flow” concept in that the flow through the orifice is limited by the orifice area and driving pressure. If the valve area is very small, it acts as a limiting nozzle – the heart can only push so much cardiac output through for a given pressure. In fluid dynamics terms, the valve orifice causes a vena contracta effect: the blood jet actually contracts slightly just downstream of the physical orifice and reaches a peak velocity at the vena contracta (this is akin to the nozzle throat where velocity peaks). The pressure right at the vena contracta is lowest. Clinically, the maximum pressure drop (ΔPmax) occurs between the LVOT (just before the valve) and the vena contracta just beyond the valve.
Supravalvular region (the downstream artery or vessel): This corresponds to the diverging section in the analogy. Beyond the valve or stenosis, the blood flows into a larger cross-sectional area (the ascending aorta or pulmonary artery, or simply a wider part of the vessel after a stenosis). Consequently, the blood decelerates and some of its kinetic energy is converted back into pressure – this is the phenomenon of pressure recovery. In the artery just downstream, the cross-sectional area is greater than the valve opening, so the high-velocity jet expands, slows down, and the static pressure in the fluid partially rises again. This is analogous to a subsonic diffuser (since blood flow remains well below sonic speeds): after a constriction, the flow spreads out and pressure increases. However, unlike an ideal diffuser, there are losses due to turbulence in the blood jet mixing with the stagnant blood in the aorta, so not all pressure is recovered. Still, the net effect is that the pressure drop from the ventricle to the distal aorta (often called ΔPrecovery) is less than the initial drop at the valve itself (ΔPmax), because some pressure is regained in the expansion region. This is similar to how an ideal divergent nozzle would recover pressure if the flow remained subsonic, except in our case the flow through the valve is like a free jet expanding into the aorta rather than a confined diffuser. The compliance of the artery also plays a role: as the high-speed blood enters, the artery can elastically expand to accommodate the surge of volume (particularly in the case of the aorta which has significant elasticity). This expansion (due to compliance) effectively increases the cross-sectional area available and aids in reducing the peak pressure and smoothing the flow – an effect with no direct analog in a rigid nozzle.
Hemodynamic Flow Equations and Analogous Variables
From a mathematical perspective, blood flow through a constricted region can be described by equations analogous to those used for nozzle flow, with appropriate modifications for the incompressible and pulsatile nature of blood. Some key correspondences and differences are outlined below:
De Laval Nozzle (Compressible Gas) |
Hemodynamic Analog (Blood Flow) |
Cross-sectional area (A): Fixed geometric area of nozzle at any section. |
Effective area (A): Cross-sectional area of blood flow path (e.g., ventricular outflow tract area, valve orifice area, arterial lumen area). In vessels, the area can change with pressure if the walls stretch (compliance). |
Velocity (v): Fluid velocity increases in convergent section and (for supersonic flow) in divergent section. |
Velocity (v): Blood velocity increases as it approaches and passes through a narrowing (e.g., higher velocity through a stenotic valve). After the constriction, the velocity decreases as the flow spreads out in a larger area (and the kinetic energy dissipates or converts to pressure). |
Continuity (mass flow): \( \dot m = \rho A v \) is constant along the nozzle. Density \( \rho \) can vary (gas compressibility). At choked flow, \( \dot m \) reaches a maximum limit for given upstream conditions. |
Continuity (volume flow): \( Q = A v \) is (approximately) constant along a closed flow circuit (incompressible assumption). Blood density is essentially constant, so volume flow is conserved. There is no direct “choking” in incompressible flow, but the maximum flow through a valve is limited by the orifice size and the available pressure drop – at very high flow, most energy goes into kinetic and further increases require disproportionately large pressure drops. |
Mach number (M): \( M = \dfrac{v}{a} \) (ratio of flow speed to local sound speed). Governs compressibility effects. \( M=1 \) at throat is the critical condition. Supersonic flow (\( M>1 \)) in divergent section means pressure disturbances cannot travel upstream. |
Wave speed ratio: Blood flows at \( v \) much lower than the pulse wave speed \( c_{\text{pulse}} \) in arteries. One could define an analog \( M_h = v / c_{\text{pulse}} \), but under normal conditions \( M_h \ll 1 \) (e.g., \( v \sim 1\,\text{m/s} \) vs. \( c_{\text{pulse}} \sim 5\mbox{–}10\,\text{m/s} \)). Thus, blood flow is highly subsonic in this sense; pressure waves travel fast relative to the flow, and upstream/downstream remain in communication (except in pathological extremes). There is no true supersonic regime in blood flow — information (pressure) can propagate upstream through the blood almost instantaneously compared to the flow speed. |
Pressure and energy: Bernoulli/Energy equation for isentropic flow: the drop in static pressure across a nozzle equals the gain in kinetic energy (neglecting potential and losses). Stagnation pressure \( p_0 \) is constant. For a large reservoir feeding a nozzle, \( p_0 \) (reservoir) and \( p_{\text{exit}} \) determine the flow velocity. Example: \( \dfrac{1}{2}\rho v^2 = p_{\text{reservoir}} - p_{\text{throat}} \) for a choked throat (roughly). |
Pressure and energy: Bernoulli’s principle can be applied to blood flow (with corrections for viscosity). The pressure drop \( \Delta P \) across a stenosis is approximately \( \tfrac{1}{2}\rho (v_2^2 - v_1^2) \), where \( v_1 \) is upstream blood velocity and \( v_2 \) is the high velocity through the orifice. In practice, if \( v_1 \ll v_2 \), this simplifies to \( \Delta P \approx \tfrac{1}{2}\rho v_2^2 \). Clinically, this is used as \( \Delta P (\text{mmHg}) \approx 4\,v_2^2 (\text{m}^2/\text{s}^2) \) (since \( \rho_{\text{blood}}\approx1060\,\text{kg/m}^3 \)). This is analogous to the conversion of pressure to kinetic energy in a nozzle. However, unlike the ideal nozzle (frictionless, no loss), blood flow experiences viscous losses and turbulence, so some pressure is irrecoverably lost (converted to heat or sound). Stagnation pressure in the ventricle is not fully recovered in the artery due to these losses. |
Density (\( \rho \)) and compressibility: Gas density decreases in the nozzle as pressure drops (significant compressibility). This allows large velocity increases without needing as high a volume flow increase. |
Density and compliance: Blood density is essentially constant (incompressible fluid), so density does not change with pressure. Instead, the compliance of the vessel plays a similar role: the vessel can expand to accommodate more blood. In equations, one might say the effective cross-sectional area can increase with pressure: \( A_{\text{artery}}(P) = A_0 + \frac{dA}{dP}\Delta P \). The vessel compliance \( C \) (volume change per pressure change) helps buffer pressure – when blood accelerates through a narrow valve and enters the artery, the artery’s expansion absorbs some kinetic energy, reducing the pressure drop needed for a given flow. This is a key difference: gas compressibility stores energy as increased density, whereas arterial compliance stores energy via wall stretch (increased volume). |
Expansion and pressure recovery: In a correctly expanded nozzle (supersonic case aside), a divergent section can slow the flow and raise static pressure (pressure recovery) if flow remains subsonic or after shocks. In supersonic flow, expansion fans lower pressure and increase velocity; pressure recovery would require a normal shock or subsonic diffuser section. |
Expansion and pressure recovery: After a stenotic valve, the blood enters a larger area in the aorta or artery, causing the jet to slow down and partial pressure recovery to occur. For example, the pressure drop between the ventricle and the distal aorta is less than the drop right at the valve, because some kinetic energy is converted back to pressure as the blood disperses and slows. The arterial walls, being elastic, also recoil (Windkessel effect) which further helps restore pressure and maintain flow. Unlike a designed diffuser, the recovery in vivo is not complete (some energy is lost to turbulence), but it is significant. This is why the pressure gradient measured by catheter (across the whole region) can be smaller than that inferred from the jet velocity by Doppler (which corresponds to the maximum drop at the vena contracta). |
In the above comparison, we see that many qualitative behaviors align: a narrowing causes acceleration and pressure drop, and a subsequent expansion allows deceleration and partial pressure rise. The major distinctions arise from compressibility versus compliance. In gases, the ability to compress (change density) is what allows the flow to reach sonic speeds and beyond, and to have a choke point. In blood (which is nearly incompressible), the analogous “give” in the system is the elasticity of the vessels. The pulse wave speed in an elastic artery (which is the speed at which pressure disturbances travel) acts like the sound speed in a gas. This wave speed \( c \) is given by Moens–Korteweg equation, \( c = \sqrt{\frac{Eh}{2\rho R}} \) for a vessel (where \( E \) is elastic modulus, \( h \) wall thickness, \( R \) radius), or in terms of area compliance \( C_A = \frac{dA}{dP} \), one can show \( c^2 \approx \frac{A}{\rho, dA/dP} \). In the human arteries, \( c \) is on the order of 5–10 m/s (much faster than blood flow), so the “Mach number” \( M_h = v/c \) stays very low (0.1–0.2 under peak flows). This means blood flow is subcritical and information travels upstream easily, preventing any choking phenomenon under normal conditions. Only in extreme situations (such as severe flow limitation in collapsible airways or perhaps shock waves in traumatic aortic ruptures) would a Mach-1-like condition occur in biological flows – these are rare and involve different mechanics (e.g., in airflow during a cough, the concept of a “choke point” does emerge).
To conclude, modeling the subvalvular, valvular, and supravalvular regions after a de Laval nozzle provides a useful conceptual and mathematical framework. The subvalvular outflow tract is like a convergent nozzle section accelerating the blood toward the valve. The valve or stenotic orifice is like a throat that creates a high-velocity jet (with an associated pressure drop). The supravalvular region (the receiving artery) behaves like a diffuser where the blood expands, slows down, and recovers some pressure. Equations analogous to those of nozzle flow – continuity (flow conservation) and energy (Bernoulli) – can be applied: for instance, using the continuity equation \( A_1 v_1 = A_2 v_2 \) to relate velocities in the ventricle and through the valve, or Bernoulli’s equation to relate pressure differences to velocity changes. Vessel compliance enters these equations by allowing the effective area to vary with pressure and by smoothing the pulsatile flow. Ultimately, while blood does not achieve supersonic speeds, the nozzle analogy helps in understanding phenomena like the formation of a high-speed jet through a stenosis, the drop and partial recovery of pressure, and the factors (area, pressure, compliance) that govern flow through cardiac valves and arterial narrowings.
Written on June 21, 2025
Integrating de Laval nozzle theory into lumped-branch hemodynamic simulation (Written June 22, 2025)
1 · Scope and motivation
Precise representation of pulsatile trans-valvular flow remains a central challenge in physics-based cardiovascular modelling.
Classic lumped-branch systemic models (LBSM) partition the circulation into resistive (R), inertial (L) and compliant (C) elements; yet the rapid,
area-dependent acceleration that dominates near the cardiac valves is handled only approximately.
By contrast, a de Laval converging–diverging nozzle describes compressible flow through a variable-area conduit with analytically
tractable momentum, energy and continuity relations.
The present note examines how nozzle concepts can clarify, surrogate or formally extend LBSM valve segments while
respecting physiological constraints.
2 · Re-examining the classical de Laval formulation
-
Canonical area–Mach relation
For quasi-one-dimensional, isentropic flow of a calorically perfect gas, the throat-normalised area \(A/A^{*}\) and local Mach
number \(M\) satisfy :contentReference[oaicite:0]{index=0}
\[
\frac{A}{A^{*}}=\frac{1}{M}\Biggl[\frac{2}{\gamma+1}\Bigl(1+\frac{\gamma-1}{2}M^{2}\Bigr)\Biggr]^{\tfrac{\gamma+1}{2(\gamma-1)}} .
\]
Because the expression is implicit in \(M\), root-finding (e.g. Newton–Raphson) is required during simulation.
-
Pressure and temperature fields
The isentropic static–to–total relations follow trivially:
\[
\frac{T}{T_{0}}=\Bigl(1+\tfrac{\gamma-1}{2}M^{2}\Bigr)^{-1},\qquad
\frac{p}{p_{0}}=\Bigl(1+\tfrac{\gamma-1}{2}M^{2}\Bigr)^{-\tfrac{\gamma}{\gamma-1}}.
\]
Choked flow occurs when \(M=1\) at the minimum area; further reductions in back pressure merely
extend the supersonic portion downstream :contentReference[oaicite:1]{index=1}.
-
Issues detected in the current JavaScript prototype
- Incorrect explicit Mach formula.
The implemented square-root form neglects the implicit dependence in the canonical relation and omits
the \(1/M\) factor.
- Throat area mis-assignment.
Astar
is inconsistently referenced as AStar
, and the special-case multiplier *100
at \(x=0\) alters the geometry.
- Non-physical scaling factors.
Multipliers of \(0.001\) (Mach) and \(1/2\) (velocity) distort dimensional coherence.
- Coordinate shift.
Translating \(x\) by \((x-0.3)\times10\) obscures the physical length scale and complicates coupling to LBSM time-bases.
- Unused update hooks.
updateComputedValues()
is empty, preventing parameter changes from propagating to derived fields.
3 · Essential code corrections (excerpt)
Mach(x) {
const A = this.A_geometry(x);
const f = M => (A/this.Astar) -
(1/M) *
Math.pow((2/(this.Gamma+1))*(1+(this.Gamma-1)*M*M/2),
(this.Gamma+1)/(2*(this.Gamma-1)));
return newtonRaphson(f, 0.2, 20); // sub-routine not shown
}
V_nozzle(x){
const M = this.Mach(x);
const T = this.T0 / (1 + (this.Gamma-1)*M*M/2);
const a = Math.sqrt(this.Gamma*287*T); // local speed of sound
return M * a;
}
4 · Mapping nozzle sections to valvular flow domains
Nozzle section | Cardiac analogue | Principal phenomena |
Converging (x < 0) | Sub-valvular tract (inflow or outflow) | Pre-acceleration, inertial load, pressure recovery |
Throat (\(A^{*}\)) | Valve orifice at maximal opening | Effective orifice area, viscous dissipation peak |
Diverging (x > 0) | Supravalvular aortic/pulmonary segment | Wave reflection, downstream compliance, jet dispersion |
5 · Levels of integration
-
Pedagogical analogy
Visualising transient valve flow as a miniature, mildly compressible nozzle highlights the transition from potential to kinetic
energy and clarifies the role of exit pressure in regurgitant states.
-
Surrogate framework
Within an LBSM time-step, substitute the valve element by a nozzle whose throat area equals the instantaneous
effective orifice area \(A_{v}(t)\).
The governing set becomes
\[
Q = A_{v}(t) \, v,\quad
\Delta p = p_{\text{up}}-p_{\text{down}}
= \rho\!\left(\frac{v^{2}}{2}\right)\!
\Bigl(f_{\gamma}(M) + \xi_{\!v}\Bigr),
\]
where \(f_{\gamma}\) derives from the isentropic nozzle relations and \(\xi_{\!v}\) covers viscous losses.
-
Formal structural coupling
A dedicated nozzle module may replace the lumped L–R pair of the valve compartment.
State variables include throat Mach number and chamber total pressure; they evolve through explicit,
semi-implicit, or operator-splitting techniques while the surrounding LBSM retains its ordinary differential
formulation in pressure and volume.
6 · Practical coupling strategy
- Synchronise discretisation. Assign the nozzle solver the same Δt as the cardiovascular ODE integrator to eliminate interpolation.
- Map parameters.
- \(A^{*}(t)\) ← measured valve area (e.g. from echocardiography).
- \(\gamma\approx1.01\) for blood (quasi-incompressible yet finite sound speed).
- Total pressure \(p_{0}\) ← upstream ventricular pressure.
- Iterative interface. Use Picard iteration to reconcile nozzle exit static pressure with LBSM downstream
arterial pressure within each cardiac time-step.
7 · Expected benefits and limitations
- Improved prediction of early-systolic jet acceleration and peak velocity without ad-hoc Bernoulli coefficients.
- Natural accommodation of pressure-wave reflection when coupled to compliant downstream segments.
- Caveats: blood compressibility is weak; therefore \(\gamma\) approaches unity and numerical stiffness increases.
Careful non-dimensional scaling is advised.
8 · Concluding remarks
Replacing or augmenting the LBSM valve compartment with a rigorously corrected de Laval-based module
offers a physically consistent route to capture rapid trans-valvular dynamics while maintaining computational
efficiency.
Continued validation against Doppler and catheter data is encouraged before clinical deployment.
Written on June 22, 2025
Mapping de Laval-nozzle theory to pulmonary arterial hemodynamics: toward an enhanced LBSM framework (Written June 22, 2025)
Abstract
The de Laval nozzle provides a classical reference for compressible, one-dimensional, quasi-steady flow with an internal geometric contraction–expansion. The lesion–valve–post-valve (subvalvular–valvular–supravalvular) arrangement of the pulmonary outflow tract resembles this topology. By recasting nozzle relations into the Lumped-Biophysical System Model (LBSM), improved interpretability of pressure–velocity waveforms may be obtained, particularly for pulmonary arterial pressure (PAP) estimation. The following note (i) corrects internal inconsistencies in the supplied JavaScript nozzle routine, (ii) establishes a disciplined variable mapping between nozzle and cardiovascular domains, and (iii) delineates the conceptual limits of the analogy given blood’s low-Mach-number, weakly compressible behaviour.
1 Refinement of the de Laval-nozzle implementation
-
Geometry
The original A_geometry
routine introduces variable-name mismatch (AStar
vs. Astar
) and a throat multiplier (×100
) that breaks continuity. A consistent definition is recommended:
\[
A(x)=A^{*}\times\begin{cases}
1-k\,x, & x<0,\\[4pt]
1+k'\,x^{2}, & x\ge 0,
\end{cases}
\qquad x\in[-L_{\text{conv}},\,L_{\text{div}}].
\]
- All symbols maintain lower-case convention; \(A^{*}\) denotes the throat area.
- Continuity of \(A\) and its derivative at \(x=0\) is preserved (C1 continuity) if \(k=k'\,x|_{0^{-}}\), or a spline is used.
-
Mach-number evaluation
Scaling factors 0.001
(Mach), /10
(temperature), and /2
(velocity) obscure physical meaning. Removal is suggested. Using the isentropic area–Mach relation:
\[
\left(\frac{A}{A^{*}}\right)^{2}=
\frac{1}{M^{2}}
\left[ \frac{2}{\gamma+1}
\left(1+\frac{\gamma-1}{2}M^{2}\right)\right]^{\frac{\gamma+1}{\gamma-1}}.
\]
A numerical root-finder (e.g. bisection or Newton–Raphson) avoids analytic inversion error. The pressure-ratio term \(\bigl(p_{e}/p_{0}\bigr)\) must be employed for the choking check, not introduced directly into every pointwise Mach evaluation.
-
Recommended code fragment
Use a pure function machFromArea(areaRatio, γ)
returning the physical root.
Remove any unit-less scaling, and employ SI units consistently (m, kg, s).
2 Variable correspondence between nozzle flow and pulmonary outflow
Nozzle symbol | Haemodynamic analogue | Interpretation |
\(x\) | Longitudinal coordinate along RVOT–PA axis | Sub-/valvular/supra-valvular positions |
\(A(x)\) | Effective luminal cross-section \(A_{\text{eff}}(x,t)\) | Dynamic change due to wall compliance |
\(A^{*}\) | Valve orifice area \(A_{\text{valve}}\) | Minimum section (systole-dependent) |
\(M\) | \(\mathrm{Re}\)-based inertial index | Low-Mach; inertial–viscous balance |
\(p_{0},\,T_{0}\) | End-diastolic pressure / temperature baseline | Reference state upstream (RV) |
\(p_{e}\) | Downstream PA pressure \(p_{\text{PA}}\) | Exit boundary condition |
\(V\) | Blood velocity \(v(t)\) | Measured by Doppler, modelled in LBSM |
In the LBSM, the pulmonary arterial velocity function
\[
v_{\text{pul}}(t)=\alpha\,e^{-\gamma_{R} t}(1-t/ET)\,t
\]
already resembles the temporal profile obtained from integrating the nozzle momentum equation under quasi-steady conditions. Substituting \(t\mapsto x/c\) (convective travel time) establishes a direct mapping for forward-propagating waves.
3 Composite valvular model using nozzle sections
-
Sub-valvular (converging)
- Area contraction \(A(x)\lt A_{0}\) raises static pressure and accelerates the flow.
- Equivalent to the muscular infundibulum; stiffness modulation is represented by parameter \(k\).
-
Valvular throat
- Geometry fixed by leaflet opening; chosen as \(A^{*}(t)\).
- Choking condition analogue: maximum transmitral or trans-pulmonary gradient before regurgitation.
-
Supra-valvular (diverging)
- Gradual expansion relaxes velocity; wave reflection occurs at branch points.
- Parameter \(k'\) links to local compliance \(C_{\text{PA}}\).
4 Integration into LBSM
- Replace empirical \(\alpha,\,\beta,\,\gamma\) coefficients with nozzle-derived expressions using \(\gamma,\,k,\,k'\).
- Express \(A_{\text{eff}}(t)\) through measured valve orifice dynamics to capture patient-specific stenosis.
- Insert the refined velocity into the pressure ODEs of LBSM, maintaining conservation of mass–momentum.
- Validate against catheter-derived PAP and Doppler flow in a cohort with varying pulmonary vascular resistance.
5 Limitations of the nozzle analogy
- Compressibility: Blood behaves as a nearly incompressible Newtonian fluid at \(M\approx10^{-3}\). The nozzle model requires compressibility for energy conversion. This discrepancy mandates reinterpretation of Mach-related terms as dimensionless inertial indices rather than true Mach numbers.
- One-dimensionality: The nozzle neglects secondary flow and wall shear; pulmonary arteries exhibit complex three-dimensional pulsatility and viscoelastic wall behaviour.
- Rigid walls in nozzle theory: LBSM incorporates compliance via \(C\), but a pure nozzle does not. Coupling must therefore include a compliance term or employ fluid–structure interaction corrections.
- Turbulence onset: Transitional Reynolds numbers in large vessels (Re ≈ 2000) differ from gas-dynamic criteria; least-squares calibration against clinical data remains essential.
6 Concluding remarks
The nozzle framework, when prudently corrected and cautiously interpreted, affords an analytically tractable surrogate for rapid forward modelling of pulmonary flow and pressure. Embedding its geometry-dependent relations into the LBSM allows direct linkage between measurable valvular geometry and global haemodynamics, advancing patient-specific assessment of right-ventricular afterload.
Written on June 22, 2025