Fact-checked by Grok 2 weeks ago

Two-phase flow

Two-phase flow is the simultaneous movement of two distinct, immiscible phases—most commonly a and a gas—within a conduit, , or , where the phases interact through interfaces such as menisci, leading to complex hydrodynamic behaviors distinct from single-phase flows. This phenomenon arises under conditions where phase change, mixing, or separation occurs, such as or , and is governed by principles of including momentum, mass, and . In engineering contexts, two-phase flow is critical due to its influence on pressure drops, heat transfer rates, and system efficiency, often resulting in higher friction losses than single-phase flows by factors that can exceed unity significantly. The flow can exhibit various regimes depending on factors like phase velocities, pipe orientation, diameter, and fluid properties; common regimes include bubbly flow (dispersed small gas bubbles in a continuous liquid), slug flow (large intermittent gas pockets separated by liquid slugs), churn flow (chaotic mixing of phases in larger pipes), annular flow (liquid film along walls with a gas core), and mist or dispersed flow (liquid droplets entrained in gas). These regimes transition based on dimensionless parameters such as the Reynolds number ratio and Suratman number, with bubbly-to-slug transitions occurring around Re_G/Re_L ≈ 464 Su^{-2/3} in vertical flows. Applications of two-phase flow span multiple industries, including nuclear power systems where it affects coolant stability and critical heat flux during boiling (potentially reducing it by up to 40% due to instabilities like oscillations or flow reversal), oil and gas production for predicting pressure drops in wells and pipelines, geothermal energy extraction, and refrigeration cycles in evaporators and condensers. More recently, as of 2024, two-phase flow has gained prominence in high-flux thermal management for electronics cooling in data centers and power electronics for electric vehicles and AI hardware. In space exploration, it is vital for life support systems like water recovery and air revitalization under microgravity, where regime maps and pressure drop correlations ensure reliable operation, with ongoing microgravity experiments advancing understanding. Modeling approaches, such as homogeneous equilibrium or separated flow models, are used to predict behaviors, though challenges persist in capturing interfacial dynamics and phase interactions accurately; recent advances include multiscale numerical simulations.

Fundamentals

Definition and Scope

Two-phase flow describes the simultaneous transport of two immiscible phases, such as gas and or and , within a conduit, , or , where the phases maintain distinct identities and interact dynamically at their interfaces. This contrasts with single-phase flow by introducing complexities like between phases, interfacial tension, phase slip, and potential phase changes, which significantly alter momentum, heat, and behaviors. Common examples include steam-water mixtures in power generation systems and oil-gas transport in pipelines, where the presence of interfaces leads to non-uniform velocity profiles and enhanced mixing compared to homogeneous flows. Two-phase flow phenomena were first encountered by engineers in the with the advent of steam during the . Systematic studies emerged in the early , focusing on performance and flow instabilities, with foundational work on water circulation in forced- systems by the late 1920s. A pivotal advancement came in 1949 with the Lockhart-Martinelli correlation, which provided an empirical method to predict frictional drops in isothermal two-phase pipe flows by relating them to single-phase equivalents, influencing subsequent modeling efforts. The scope of two-phase flow primarily encompasses configurations where phases coexist without fully mixing, including dispersed flows—characterized by one phase forming discrete bubbles, droplets, or particles suspended in a continuous carrier —and separated flows, such as stratified arrangements where phases occupy distinct regions due to differences or . While the field centers on binary phase interactions, extensions to three-phase or multiphase systems are considered in specialized contexts like , though these introduce additional complexities beyond the core two-phase framework. Analysis of two-phase flow builds on foundational , presupposing familiarity with conservation laws such as the for mass balance and the Navier-Stokes equations for momentum transport across phases.

Phases and Interfaces

Two-phase flow involves the simultaneous movement of two immiscible phases, each characterized by distinct physical properties that govern their interactions. The most common configuration is gas-liquid flow, such as air-water systems, where the gas phase typically exhibits low density (e.g., around 1.2 kg/m³ for air at standard conditions) and viscosity (approximately 1.8 × 10⁻⁵ Pa·s), while the liquid phase has higher density (e.g., 1000 kg/m³ for water) and viscosity (about 0.001 Pa·s). Surface tension, a critical property at the interface, measures the cohesive forces within the liquid (e.g., 0.072 N/m for water-air), influencing bubble or droplet formation and stability. Liquid-liquid flows, like oil-water emulsions, feature immiscible fluids with comparable densities but differing viscosities, such as water (1 cP) and crude oil (up to 1000 cP), where interfacial tension (typically 0.01–0.05 N/m) promotes emulsification. Solid-liquid flows, exemplified by slurries, involve dispersed solid particles in a carrier liquid, with phase properties including solid density (e.g., 2500 kg/m³ for silica) exceeding that of the liquid, and effective viscosity increasing with particle concentration due to inter-particle interactions. The between phases is a dynamic boundary where effects arise from imbalances in across the surface, governed by the Young-Laplace equation, leading to phenomena like droplet or capillary rise. Slip velocity, the between phases at the , emerges due to differences and can enhance effective permeability in porous media flows by up to 30% under certain conditions. Interfacial tension forces minimize surface area, driving coalescence or breakup, while wettability—quantified by the (θ, from 0° for complete to 180° for non-wetting)—determines phase to solid surfaces; for instance, θ < 90° favors liquid spreading in brine-CO₂ systems. In flows involving phase change, such as boiling and condensation, interfaces become dynamic as vaporization or liquefaction occurs. Boiling initiates at nucleation sites—microscopic cavities on heated surfaces that trap vapor or gas—requiring wall superheat to activate bubble growth, with site density influencing heat transfer efficiency in systems like nuclear reactors. Condensation forms liquid films on cooler surfaces, where nucleation begins at impurities or roughness, creating transient interfaces that evolve through droplet coalescence. These processes are pivotal in applications like heat exchangers, where controlled phase changes enhance thermal performance. A key metric quantifying phase interaction intensity is the interfacial area concentration, defined as the local interfacial area per unit volume, which captures the extent of contact and thus the rates of mass, momentum, and energy exchange between phases. This parameter varies with flow conditions, peaking near walls in bubbly regimes, and is essential for two-fluid modeling to predict transfer processes accurately.

Flow Regimes

Gas-Liquid Patterns

In gas-liquid two-phase flows, distinct flow patterns emerge based on the interplay of phase velocities, densities, viscosities, surface tension, pipe diameter, and orientation, influencing heat and mass transfer as well as pressure gradients in engineering systems such as pipelines and boilers. These patterns are classified into several primary regimes, each characterized by specific interfacial structures and phase distributions. Bubbly flow occurs at low gas velocities, where discrete gas bubbles are dispersed uniformly within a continuous liquid phase, with bubbles typically small and spherical due to surface tension dominance; this regime is common in vertical upward flows or large-diameter horizontal pipes. As gas flow increases, slug flow develops, featuring large, elongated Taylor bubbles that nearly fill the pipe cross-section, separated by liquid slugs containing smaller bubbles; these Taylor bubbles rise due to buoyancy, promoting efficient mixing but also pressure fluctuations. At higher velocities, churn flow appears as a transitional, highly turbulent regime with chaotic, oscillating liquid slugs and fragmented interfaces, often observed in vertical pipes where bubble coalescence and breakage intensify. Annular flow forms when gas velocity is sufficient to shear the liquid into a thin film along the pipe wall, with a high-velocity gas core possibly entraining droplets; this pattern prevails in both horizontal and vertical configurations at moderate to high gas rates. In horizontal pipes, stratified flow arises under gravity when liquid settles at the bottom and gas flows above, potentially developing waves at the interface if velocities increase; this regime is absent in vertical flows due to lack of gravitational separation. Finally, mist or dispersed flow occurs at very high gas velocities, where the liquid phase breaks into fine droplets entrained in the continuous gas, resembling a fog-like suspension. Transitions between these patterns are predicted using criteria based on superficial gas and liquid velocities, which represent the volumetric flow rates per unit cross-sectional area; for horizontal pipes, the Taitel-Dukler map delineates boundaries such as the shift from stratified to annular when Kelvin-Helmholtz instability waves destabilize the interface. In vertical pipes, a corresponding map by Taitel et al. outlines transitions like bubbly to slug via bubble crowding and coalescence, or slug to churn through flooding mechanisms. Pipe inclination significantly alters these criteria, favoring stratified patterns near horizontal orientations but promoting churn or annular in near-vertical setups; smaller diameters enhance bubbly and slug stability by restricting bubble rise, while larger diameters permit earlier stratification. Fluid properties further modulate boundaries, with higher liquid viscosity delaying bubbly-to-slug transitions by hindering coalescence, and greater density differences accelerating drift-flux effects in vertical flows. Experimentally, these patterns are identified through high-speed imaging, which captures interfacial dynamics and bubble shapes in real-time, or conductivity probes that detect phase changes via electrical resistance variations between gas (non-conductive) and liquid (conductive). In vertical flows, the churn regime often dominates at high gas velocities due to intermittent liquid bridging and gas penetration, contrasting with horizontal flows where stratified patterns persist under similar conditions. Pressure drop tends to be elevated in slug and churn patterns owing to periodic accelerations, though detailed analysis appears in subsequent sections.

Solid-Liquid and Other Combinations

In solid-liquid two-phase flows, regimes are broadly classified into homogeneous and heterogeneous patterns, differing significantly from gas-liquid flows due to the comparable densities of the phases and dominant gravitational settling effects. Homogeneous flow occurs when solid particles remain fully suspended throughout the liquid, resulting in a uniform mixture that behaves as a with enhanced viscosity; this regime is maintained at sufficiently high liquid velocities that counteract individual particle settling. Heterogeneous flow, in contrast, features particle settling to the pipe bottom, forming either a stationary or moving bed load where particles roll along the wall, or saltation where particles intermittently lift off the bed in jumping trajectories influenced by turbulence; these patterns predominate at lower velocities or higher particle concentrations, leading to stratified distributions. The settling dynamics in these regimes are governed by the terminal settling velocity of isolated particles, modified by hindered settling at elevated concentrations, where inter-particle interactions reduce the effective descent speed. The seminal quantifies this hindrance as u_s = u_t (1 - C)^{n-1}, where u_s is the hindered settling velocity, u_t is the terminal velocity, C is the volumetric solids concentration, and n is an empirical exponent (typically 4.65 for low Reynolds numbers, decreasing to around 2.4 at higher values) that depends on particle Reynolds number and accounts for drag augmentation from neighboring particles. This effect is critical in homogeneous regimes to prevent segregation and in heterogeneous ones to predict bed formation thresholds. The further distinguishes heterogeneous transport by layering the flow into a lower heterogeneous zone with settled particles and an upper homogeneous suspension, incorporating hindered settling to estimate layer velocities and overall pressure gradients. Liquid-liquid two-phase flows exhibit dispersed and separated patterns, driven primarily by viscosity contrasts rather than density differences, with emulsification potential altering regime stability through droplet coalescence or breakup. In dispersed flow, the less viscous liquid forms droplets suspended in the continuous more viscous phase (or vice versa), favored when the viscosity ratio m = \mu_d / \mu_c (dispersed to continuous phase) is near unity, promoting uniform distribution and minimal phase separation under moderate flow rates. Separated patterns, such as core-annular flow, arise at high viscosity ratios (m \gg 1), where the viscous liquid cores the pipe surrounded by a lubricating annular film of the less viscous phase, reducing wall shear and enabling efficient transport of high-viscosity oils; stability depends on interfacial tension and flow rates, with waves at the interface potentially leading to emulsification if shear exceeds critical thresholds. Transitions between these patterns are influenced by viscosity ratios exceeding 10, where core-annular dominates to minimize energy dissipation, and emulsification risks increase with prolonged high-shear exposure. Gas-solid two-phase flows, as in pneumatic transport, feature dilute and dense phase regimes, characterized by particle suspension in gas streams with choking risks in vertical configurations unlike the slip-dominated gas-liquid cases. Dilute phase flow involves low solids loading (typically <15 kg solids per kg gas) where particles accelerate individually with the gas, resembling turbulent suspension transport at high velocities (>15-20 m/s). Dense phase flow occurs at higher loadings, with particles forming clusters or plugs that propagate intermittently, reducing velocity fluctuations but increasing pressure drops. In vertical risers, choking manifests as a transition from dilute to dense flow when gas velocity falls below a critical choking velocity (often 3-6 m/s depending on particle size), causing particle accumulation, voidage collapse, and potential blockage due to insufficient drag to suspend the solids. Regime transitions across these combinations hinge on particle size (finer particles favor homogeneous suspension, coarser ones promote heterogeneous bedding), solids concentration (higher values enhance hindrance and stratification), and flow rates (increased liquid or gas velocity shifts toward suspension-dominated patterns). Unique to solid-involved flows are erosion risks from high-velocity particle-wall impacts, and deposition hazards that can initiate blockages in low-velocity heterogeneous regimes. Such patterns underpin industrial applications like slurry pipelines for mineral transport, where maintaining homogeneous flow minimizes energy use.

Applications

Industrial and Engineering Uses

Two-phase flow plays a critical role in the energy sector, particularly in nuclear reactors where boiling water reactors (BWRs) utilize two-phase steam-water flow as to efficiently remove heat from fuel assemblies, enhancing safety and performance during operation. In steam turbines, of wet steam forms two-phase flows that can reduce efficiency due to non-equilibrium effects but are essential for energy extraction, with studies showing that droplet injection can control flow structures to minimize losses. Similarly, oil-gas transport in pipelines relies on two-phase flow regimes such as annular and slug patterns to move multiphase mixtures over long distances, requiring careful pressure management to maintain steady transport. In chemical processing, multiphase reactors employ gas-liquid two-phase flows to facilitate reactions like and oxidation, where bubble columns and trickle beds promote and mixing for producing chemicals and polymers. Distillation columns often feature froth flows—two-phase dispersions of vapor and liquid on trays—that enhance separation efficiency, with advanced profiling techniques measuring effective froth height to optimize column design and capacity. Refrigeration and (HVAC) systems depend on two-phase refrigerant flows in and condensers, where phase change from to vapor absorbs in , and releases it, enabling efficient management in cycles like vapor-compression systems. Numerical analyses confirm that varying velocities and temperatures in these components influence rates, underscoring the need for precise flow control. As of 2025, advancements include (EOR) using CO2 foam flows, where ultra-dry CO2-in-water foams improve sweep efficiency and enable by stabilizing two-phase displacements in reservoirs. In electronics cooling, microchannel two-phase flows with refrigerants like HFE-7100 handle ultra-high heat fluxes up to 345 W/cm², leveraging to dissipate from high-power chips while mitigating instabilities. Design challenges in these applications center on flow assurance, particularly preventing —which causes pressure surges and vibrations—and blockages in pipelines, often addressed through terrain modeling and valve controls to stabilize two-phase dynamics.

Environmental and Natural Contexts

Two-phase flows occur prominently in geophysical processes, such as volcanic eruptions where gas exsolves from , creating a separated gas-liquid or gas-solid mixture that drives explosive dynamics. In volcanic eruptions, the separation of gas bubbles from the leads to rapid acceleration of the mixture, influencing eruption styles from effusive to highly explosive. For instance, during the 1980 eruption (involving dacitic ), the ascent of gas-saturated through the conduit involved nonequilibrium two-phase flow, with fragmentation occurring as pressure dropped, producing pyroclastic flows that traveled up to 8 km from the vent, and a lateral blast that devastated areas up to 25 km away. Landslides and debris flows represent solid-liquid two-phase systems, where saturated soil or rock mixes with water, resulting in high-density, high-velocity downslope movements. These flows exhibit complex interactions between solid particles and interstitial fluid, governed by frictional and collisional stresses, as modeled in generalized two-phase frameworks that account for phase segregation and momentum exchange. In atmospheric and oceanic settings, two-phase flows arise during cloud formation, where water vapor condenses into liquid droplets within turbulent air, forming a dispersed gas-liquid mixture that affects radiative properties and . Ocean waves entrain air into during breaking, producing whitecaps characterized by bubbly two-phase flows with void fractions up to 0.5 near the surface, enhancing and momentum transfer across the air-sea interface. These natural oceanic processes contribute to global carbon cycling by facilitating CO2 dissolution. Hydrological systems feature two-phase flows in river , where nondilute suspensions of solid particles in create stratified profiles with lag between phases, influencing bed and deposition. During floods, form hyperconcentrated two-phase mixtures, with concentrations exceeding 50% by volume, leading to increased and altered resistance compared to clear floods. Environmental impacts of two-phase flows include dispersion in bubbly regimes, such as rivers or coastal waters, where entrained air bubbles enhance mixing and vertical transport of contaminants, prolonging exposure times in ecosystems. In carbon sequestration efforts, injecting CO2 into saline aquifers creates immiscible gas-liquid two-phase flows, where capillary trapping immobilizes up to 20% of the injected CO2 as residual ganglia, mitigating leakage risks over geological timescales. Field measurements of these natural two-phase systems pose challenges due to and transient dynamics, often requiring integrated and modeling approaches.

Flow Characteristics

Void Fraction and Quality

In two-phase flows, the void fraction, denoted as \alpha, quantifies the volume occupied by the gas relative to the total volume of the mixture. It is defined as \alpha = \frac{V_g}{V_g + V_l}, where V_g and V_l are the volumes of the gas and liquid , respectively. This parameter, often time-averaged due to fluctuations in instantaneous values, serves as a key indicator of distribution and is essential for characterizing behavior in channels. Closely related is the quality, x, which represents the mass fraction of the vapor (or gas) phase in the total mixture. It is expressed as x = \frac{m_g}{m_g + m_l}, where m_g and m_l are the masses of the gas and liquid phases. provides a thermodynamic perspective on phase composition, particularly in evaporating or condensing flows, and differs from void fraction due to the density contrast between phases. The interconnection between void fraction and quality arises through the slip ratio S, defined as the ratio of gas to . The relationship is given by \alpha = \left[1 + \frac{1 - x}{x} \cdot \frac{\rho_g}{\rho_l} \cdot S \right]^{-1}, where \rho_g and \rho_l are the gas and densities. This equation accounts for velocity differences between phases, with S > 1 typically observed in separated flows due to or effects. Under the homogeneous flow assumption, where phases travel at the same velocity (S = 1), the relation simplifies to \alpha = \frac{x}{x + (1 - x) \frac{\rho_g}{\rho_l}}. This model assumes perfect mixing, akin to a single pseudo-fluid, and yields higher void fractions for low-density gas phases compared to separated flow scenarios. It is most applicable to or regimes but overpredicts \alpha in flows with significant slip. Void fraction is commonly measured using quick-closing valves, which isolate a section to capture the instantaneous volumes; the gas volume fraction is then determined from the drained liquid volume relative to the known section volume. Another established technique is gamma densitometry, employing absorption to infer \alpha from beam , calibrated against known densities via I = I_0 e^{-\mu z}, where I is the transmitted intensity, \mu the , and z the path length. These methods provide reliable average values, though gamma techniques require safeguards and are sensitive to flow patterns. Void fraction profoundly influences the density \rho_m = \alpha \rho_g + (1 - \alpha) \rho_l, which in turn affects buoyancy-driven phenomena in vertical flows.

Pressure Drop and Velocity Profiles

In two-phase flows, the total along a or conduit is composed of three primary components: frictional, accelerational, and gravitational. The frictional component arises from stresses at the and between phases, often quantified using a two-phase multiplier \phi^2 = \Delta P_{tp} / \Delta P_l, where \Delta P_{tp} is the two-phase frictional and \Delta P_l is the for flowing alone under the same conditions. The accelerational component results from changes in the density due to phase distribution variations, particularly in flows where the void fraction increases, such as during or . The gravitational component depends on the liquid holdup, which determines the effective weight of the in vertical or inclined flows. The frictional pressure drop is commonly predicted using the Lockhart-Martinelli parameter, defined as X^2 = (\Delta P_l / \Delta P_g), where \Delta P_g is the for gas flowing alone. This parameter facilitates correlations for the two-phase multiplier, such as \phi_l^2 = 1 + C/X + 1/X^2, where C is an empirical constant typically set to 20 for turbulent-turbulent flow conditions in both phases. Originally developed for isothermal, adiabatic gas-liquid flows, this approach has been validated across a range of pressures and flow rates, providing a foundational for engineering predictions. Velocity profiles in two-phase flows are characterized by differences between the gas and phases, often described using superficial velocities j_g and j_l, which represent the velocities each would have if flowing alone in the conduit. The actual velocities u_g and u_l exceed these due to interactions, with the slip velocity u_s = u_g - u_l accounting for relative motion, typically positive as the gas moves faster. In annular regimes, prevalent at high gas fractions, the velocity profile features a relatively flat, high-speed gas core surrounded by a turbulent film near the wall, where the film's velocity decreases parabolically toward the wall due to viscous effects. Pressure drops in two-phase systems are typically measured using differential pressure transducers, which detect changes across test sections or orifices with high sensitivity to dynamic fluctuations. A notable is , observed in nozzles where the two-phase mixture reaches sonic velocity at the throat, limiting mass flow rates regardless of downstream conditions and often leading to critical flow states.

Modeling Approaches

Analytical and Empirical Models

Analytical and empirical models provide simplified frameworks for predicting two-phase flow behavior by incorporating key assumptions about phase interactions, velocities, and properties, enabling practical calculations without full computational simulations. These models balance theoretical derivations with experimental data to estimate parameters such as void fraction, , and flow regimes, often assuming isothermal or conditions to reduce complexity. They are foundational in engineering design, offering closed-form solutions that approximate real-world phenomena in pipes and channels. The homogeneous equilibrium model (HEM) treats the two-phase mixture as a single pseudo-fluid with uniform velocity and between phases, simplifying analysis for flows where rapid phase interactions occur, such as or . In this approach, the mixture density is calculated as \rho_m = \alpha \rho_g + (1 - \alpha) \rho_l, where \alpha is the void fraction, \rho_g the gas density, and \rho_l the liquid density; the mixture viscosity and other properties are similarly averaged. HEM is particularly suited for critical flow scenarios, like discharge or vessel blowdown, where it predicts maximum under isentropic expansion assumptions. This model was originally developed for predicting the maximum in single-component two-phase mixtures. The drift- model addresses relative velocities by decomposing the total into a and a drift component, providing a more accurate representation of void in dispersed flows compared to homogeneous assumptions. The void is given by \alpha = \frac{j_g}{C_0 (j_g + j_l) + u_d}, where j_g and j_l are the superficial gas and velocities, C_0 is the distribution coefficient (typically around 1.2 for flows, accounting for lateral profiles), and u_d is the drift (e.g., bubble rise relative to the ). This model originates from averaging volumetric concentrations in two- systems, emphasizing kinematic effects like . It performs well in vertical or inclined flows, capturing non-uniform distributions without requiring separate momentum equations. Separated flow models, such as the Lockhart-Martinelli approach, treat gas and phases as flowing independently in , interacting only through interfacial , which is ideal for stratified or annular regimes where phases maintain distinct velocities. The Lockhart-Martinelli parameter X = \sqrt{\frac{(dP/dz)_l}{(dP/dz)_g}} compares single-phase gradients for liquid and gas alone, enabling correlations for two-phase multipliers \phi_l^2 = 1 + \frac{C}{X} + \frac{1}{X^2} (with C \approx 20 for turbulent-turbulent ) to predict total \frac{dP}{dz}_{TP} = \phi_l^2 \frac{dP}{dz}_l. This seminal correlation was derived from isothermal air-liquid data in horizontal pipes. The Chisholm parameter extends this by providing values of C for arbitrary phase combinations and regimes (e.g., C=12 for turbulent-viscous, C=10 for viscous-turbulent, C=5 for viscous-viscous), improving applicability across viscosities and densities. These models are widely used for frictional in pipelines. Empirical correlations, like the , map flow pattern transitions based on superficial velocities normalized by phase densities and , using correction factors \lambda = \sqrt{ \frac{\rho_g}{\rho_{air}} \frac{\rho_l}{\rho_w} } for the gas and \psi = \frac{\sigma_w}{\sigma} \left[ \frac{\mu_l}{\mu_w} \frac{\rho_w}{\rho_l} \right]^{1/3} for the liquid (with reference air-water conditions: \rho_w = 1000 kg/m³, \rho_{air} = 1.23 kg/m³, \sigma_w = 0.072 N/m, \mu_w = 0.001 ·s). The is plotted with axis x = \frac{j_l \sqrt{\rho_l / \rho_w} }{\lambda} and vertical axis y = \frac{j_g \sqrt{\rho_g / \rho_{air}} }{\psi}, delineating regimes such as , , and annular. Developed for oil-gas flows in large pipes, it aids initial design assessments. However, the has limitations in non- orientations, where gravity effects alter patterns, and at low s, where influences are underrepresented, leading to inaccuracies outside oilfield conditions. These models find wide application in pipeline design to estimate holdup and losses efficiently.

Numerical and Computational Methods

Numerical and computational methods play a crucial role in simulating the complex dynamics of two-phase flows, enabling the prediction of interfacial behavior, phase interactions, and regime transitions in scenarios where analytical solutions are infeasible. These approaches solve the governing equations for , , and while accounting for phase coupling through interfacial terms. Key methods include the two-fluid model, volume of fluid techniques, and Eulerian-Lagrangian formulations, each suited to different flow regimes and computational demands. Validation against experimental data ensures accuracy, particularly for regime transitions like bubbly to . The two-fluid model (TFM) treats each phase as interpenetrating continua, deriving separate conservation equations for , , and per phase k. The volume fraction \alpha_k represents the occupancy of phase k, with \sum \alpha_k = 1. The momentum equation for phase k is given by: \frac{\partial (\alpha_k \rho_k \mathbf{u}_k)}{\partial t} + \nabla \cdot (\alpha_k \rho_k \mathbf{u}_k \mathbf{u}_k) = -\alpha_k \nabla P + \nabla \cdot \boldsymbol{\tau}_k + \mathbf{F}_k where \rho_k is the , \mathbf{u}_k the , P the (shared or phase-specific), \boldsymbol{\tau}_k the stress tensor, and \mathbf{F}_k includes interfacial forces like and . Closure relations for interfacial transfers, such as , are essential for solvability. This model, foundational since its formulation in the 1970s, excels in simulating dispersed and separated flows in engineering applications like nuclear reactors. The volume of fluid (VOF) method captures sharp by advecting a field \alpha that indicates the presence of each within computational cells, solving a single set of momentum equations with variable properties. reconstruction ensures geometric accuracy: the piecewise linear calculation (PLIC) approximates the as a plane within each cell using the and normal vector, while level-set methods implicitly track the via a for smoother evolution. Introduced in the early , VOF with PLIC is widely used for free-surface and droplet simulations due to its mass conservation and ability to resolve topological changes like and coalescence. Eulerian-Lagrangian approaches are ideal for dispersed two-phase flows, where the continuous phase (e.g., liquid) is modeled Eulerianly via Navier-Stokes equations, and the dispersed phase (e.g., bubbles or particles) is tracked Lagrangianly as point masses following Newton's laws, incorporating forces like and . Inter-phase coupling updates the continuous phase momentum via source terms from dispersed trajectories. Implementations in commercial CFD software, such as ANSYS Fluent's Discrete Phase Model, handle high particle loadings efficiently for applications like spray combustion. This hybrid method reduces computational cost compared to fully Eulerian models for low-volume-fraction dispersoids. As of 2025, advances integrate to enhance closures in multiphase CFD, using data-driven models to predict subgrid-scale interfacial transfers and improve RANS/ accuracy in bubbly or droplet-laden turbulent flows, achieving up to 10-fold simulation speedups. High-fidelity direct numerical simulations (DNS) resolve microscale interfaces, revealing fine details like bubble deformation and wake interactions, with grid requirements scaling as \mathcal{O}(\mathrm{Re}^{9/4}) for high Reynolds numbers. These simulations validate coarser models and inform closure developments. Overall, numerical methods are routinely validated against experiments for regime transitions, demonstrating good agreement in void fraction and velocity profiles during shifts from annular to .

Advanced Phenomena

Acoustics and Wave Propagation

In two-phase bubbly mixtures, the is dramatically reduced compared to that in the pure phase due to the compressibility of the gas bubbles. The effective c is given by Wood's formula for low-frequency waves in a homogeneous : c = \left[ \left( \frac{\alpha}{\rho_g c_g^2} + \frac{1 - \alpha}{\rho_l c_l^2} \right) \left( \alpha \rho_g + (1 - \alpha) \rho_l \right) \right]^{-1/2}, where \alpha is the void fraction, \rho_g and c_g are the gas and speed, and \rho_l and c_l are the corresponding properties. This approximation assumes equilibrium between phases and neglects dynamic interactions between bubbles, which more advanced models incorporate through additional terms for viscous and thermal effects. For instance, in -air mixtures, the speed drops from approximately 1500 m/s in pure to below 100 m/s at a void fraction of 5%. The void fraction \alpha strongly influences the 's acoustic properties, with even small gas volumes causing substantial changes in wave behavior. , defined as Z = \rho c where \rho is the , exhibits mismatches at bubble-liquid interfaces, leading to and of waves. Attenuation of in bubbly flows arises primarily from by individual , viscous in the liquid, and across the bubble walls. enhances near the Minnaert frequency, the natural oscillation of a bubble given by f = \frac{1}{2\pi R} \sqrt{\frac{3\gamma P}{\rho_l}}, where R is the bubble radius, \gamma is the polytropic exponent (approximately 1.4 for air), P is the ambient pressure, and \rho_l is the liquid density; this frequency typically falls in the audible range for millimeter-sized bubbles. Pressure waves in pipes containing two-phase flows propagate at the reduced mixture sound speed, influencing transient behaviors such as water hammer and system stability. In engine applications, two-phase nozzle flows generate excess noise through bubble-induced pressure fluctuations and interactions, contributing to broadband acoustic emissions beyond single-phase predictions. A notable application is void detection in nuclear reactors, where acoustic measurements of sound speed variations enable non-invasive estimation of local void fractions in coolant channels. As of 2025, advances include Monte Carlo-based modeling for ultrasonic propagation through bubbly mixtures, improving predictions of attenuation in multiphase flows. Experimentally, pulse-echo techniques measure the as a function of void fraction by transmitting ultrasonic pulses through the mixture and analyzing the time-of-flight of echoes from reflectors or interfaces. These methods provide high for transient bubbly flows, with spatial accuracy down to millimeters, and are validated against direct visualization for void fractions up to 20%.

Heat and Mass Transfer Effects

In two-phase flow, mechanisms are profoundly influenced by phase interactions, particularly during where liquid-vapor transitions dominate energy exchange. regimes begin with , characterized by bubble nucleation, growth, and departure from the heated surface, which facilitates high rates through efficient removal. As wall superheat increases, isolated bubbles evolve into jet and column structures, sustaining vigorous mixing until the is approached. Transition boiling follows, marked by unstable partial vapor coverage on the surface, leading to fluctuating and generally declining efficiency due to intermittent liquid-wall contact. Beyond this, film boiling establishes a continuous vapor blanket that insulates the surface, relying primarily on conduction and for dissipation, resulting in significantly lower coefficients. The across these regimes is commonly modeled as q = h (T_w - T_{sat}), where h is the convective , T_w the wall temperature, and T_{sat} the saturation , with h varying markedly by regime—peaking in and minimizing in film boiling. The onset of transition boiling is delimited by the critical heat flux (CHF), a pivotal limit beyond which surface dryout causes rapid temperature excursions. Zuber's seminal hydrodynamic model attributes CHF to the instability of vapor jets escaping the heated surface, predicting the maximum heat flux as q_{\max} = \frac{\pi}{24} \rho_g h_{fg} \left[ \frac{g (\rho_l - \rho_g) \sigma}{\rho_g^2} \right]^{1/4}, where \rho_g and \rho_l are vapor and liquid densities, h_{fg} the latent heat of vaporization, g gravitational acceleration, and \sigma surface tension; this correlation has been validated across fluids and pressures, establishing a foundational benchmark for design. Condensation processes in two-phase flow contrast by involving vapor collapse into , with heat and coupled through interfacial phase change. , prevalent on wettable surfaces, forms a laminar whose thickness governs resistance, while dropwise modes on non-wettable surfaces yield droplets that coalesce and shed, offering 5-10 times higher rates but challenging reproducibility. Nusselt's classical analysis for vertical falling films assumes negligible and vapor , deriving the local as h = 0.943 \left[ \frac{k_l^3 \rho_l (\rho_l - \rho_g) g h_{fg}}{\mu_l \Delta T L} \right]^{1/4}, where k_l and \mu_l are and , \Delta T the driving force, and L the plate length; this expression highlights gravity's role in and remains a for predicting average coefficients in low-velocity flows. Mass transfer at phase interfaces, essential for evaporation or absorption rates, is often characterized by the Sherwood number (Sh), analogous to the Nusselt number for momentum and heat. In dispersed two-phase flows with droplets, convective effects dominate, with the Ranz-Marshall correlation capturing enhancement over pure diffusion: Sh = 2 + 0.6 Re^{1/2} Sc^{1/3}, where Re is the based on and Sc the ; this empirical form, derived from wind-tunnel experiments on evaporating drops, accounts for boundary layer thinning and is widely applied to predict interfacial rates in sprays and bubbly flows. Post-2020 microgravity studies have illuminated heat and mass transfer alterations for space propulsion and thermal management, where buoyancy absence reshapes phase distributions. International Space Station experiments with perfluorohexane in flow boiling channels (mass velocities 180-2400 kg/m²s) have shown variations in heat transfer coefficients compared to terrestrial analogs, attributed to enlarged bubbles and shifted annular-to-slug transitions that hinder liquid renewal. Pool boiling tests with FC-72 under electric fields achieved CHF enhancements to 257 kW/m² via dielectrophoretic bubble manipulation, underscoring potential for compact radiators in orbital habitats. These findings emphasize surface tension dominance and inform predictive models for cryogenic propellant systems. As of 2025, ongoing ISS experiments continue to investigate two-phase flow instabilities using AI/ML strategies to relate hydrodynamic effects to thermal transport. Enhancement strategies, such as porous coatings on boiling surfaces, amplify sites and to mitigate film blanketing. Coatings with 55-60% elevate coefficients by 33-60%, with peak gains of 216% over plain surfaces through increased wetted area and vapor escape paths; optimal thickness balances resistance and activation without flooding pores. Such modifications, tested in microchannels, also delay CHF by up to 230%, proving vital for high-flux cooling in two-phase systems. Flow regimes modulate these transfer effects, with slug flow particularly beneficial due to cyclic bubble intrusion that renews liquid films and boosts . In horizontal pipes, slug-induced mixing yields coefficients substantially higher than in stratified regimes—often doubling single-phase values—via thinned boundary layers and enhanced interfacial renewal, as mechanistic models confirm through validation against superficial velocity data.

References

  1. [1]
    [PDF] TWO PHASE FLOW CFD ANALYSIS OF REFRIGERANTS IN A ...
    In fluid mechanics, two-phase flow occurs in a system containing gas and liquid with a meniscus separating the two phases. Two-phase flow is a particular ...
  2. [2]
    Two-Phase Fluid Flow | Engineering Library
    At certain important locations in fluid flow systems the simultaneous flow of liquid water and steam occurs, known as two-phase flow.
  3. [3]
    [PDF] A Study Of Two-Phase Flow Regime And Pressure Drop In Vertical ...
    Dec 2, 2018 · ➢ Bubble flow: In two-phase flow, when there are a lot of small bubbles dispersed in a continuum of liquid, the deformable bubbles keep moving.
  4. [4]
    [PDF] Two Phase Flow Modeling: Summary of Flow Regimes and ...
    Two-phase flow of gas and liquid, and multiphase flow of gas, liquid and solid have been identified and given high priority in the final report of a workshop on ...
  5. [5]
    [PDF] Two-Phase Flow in Geothermal Wells - Stanford University
    These correlations have been modified to suit the condition that exist in geothermal wells. A complete analytical description of one dimensional two-phase flow ...
  6. [6]
    [PDF] Tulsa University Fluid Flow Projects - OSTI
    Two-phase liquid pipe flow is defined as the simultaneous flow of two immiscible liquids in pipes. One of the common occurrences in the petroleum industry ...
  7. [7]
    [PDF] NOTES ON TWO-PHASE FLOW, BOILING HEAT TRANSFER, AND ...
    Definition of the basic two-phase flow parameters. In this document the subscripts “v” and “ℓ” indicate the vapor and liquid phase, respectively. The.
  8. [8]
    [PDF] Module 12: Two Phase Heat Transfer and Fluid Flow
    Start-up and Steady State Operation. • Operational Transients. • Loss of Coolant Accidents. Each of these stages require many different analytical techniques ...
  9. [9]
    The History of the Steam-Generating Boiler and Industry
    The steam-generating boiler's roots go back to the late 1700s and early 1800s with the development of the kettle-type boiler, which simply boiled water into ...
  10. [10]
    [PDF] Review of two-phase flow instabilities in macro - Purdue Engineering
    Study of two-phase flow instabilities began in the late 1920 s, and in the nearly 100 years since, signifi- cant progress has been made in both experimental ...
  11. [11]
    A theoretical basis for the Lockhart-Martinelli correlation for two ...
    Lockhart, R.C. Martinelli. Proposed correlation of data for isothermal two-phase, two-component flow in pipes. Chem. Engng Prog., 45 (1) (1949), pp. 39-48.
  12. [12]
    [PDF] Numerical Modeling of Dispersed Two-Phase Flows - AerospaceLab
    This paper presents some fundamental aspects of the mathematical and numerical modeling of dispersed two phase flows. The special case of gas-particle flows.
  13. [13]
    A History of Gas-Liquid Two-Phase Flow Research and the ... - J-Stage
    The progress of the two-phase flow research has been divided into four stages in this series of the report. These periods are (I) 1948-1959, (II) 1960-1970, ...
  14. [14]
    Two-Phase Flow - an overview | ScienceDirect Topics
    Two-phase flow refers to the interactive flow of two distinct phases – each phase representing a mass or volume of matter – with common interfaces in a channel.
  15. [15]
    Two-Phase Flow - IntechOpen
    Two-phase flow involves the sum of liquid and gas mass flow rates, and flow patterns depend on fluid properties and flow rates.
  16. [16]
    The influence of interfacial slip on two‐phase flow in rough pores
    Aug 1, 2017 · We show that the slip flow can be cast in more familiar terms as a contact-angle (wettability)-dependent effective permeability to the invading ...1 Introduction · 2 Computational Model · 2.2 Fluid Dynamics
  17. [17]
    Local formulation and measurements of interfacial area ...
    Interfacial area concentration is key in two-phase flow. This paper introduces a local formulation and measurement methods, including a three-probe method. ...
  18. [18]
    A model for predicting flow regime transitions in horizontal and near ...
    Models are presented for determining flow regime transitions in two-phase gas-liquid flow. The mechanisms for transition are based on physical concepts and are ...Missing: original | Show results with:original
  19. [19]
    Two-phase flow patterns: A review of research results - ScienceDirect
    This paper is a literature review covering various aspects of two-phase flow patterns. After a description of commonly observed flow regimes in horizontal and ...
  20. [20]
    Modelling flow pattern transitions for steady upward gas‐liquid flow ...
    Models for predicting flow pattern transitions during steady gas-liquid flow in vertical tubes are developed, based on physical mechanisms suggested for each ...
  21. [21]
  22. [22]
    Hindered settling of a log-normally distributed Stokesian suspension
    Dec 12, 2024 · Also, the Richardson–Zaki correlation predicts the hindered settling functions of smaller particles poorly. For the other three models, a ...
  23. [23]
    Forced convection in two-phase core-annular flows | Journal of Fluid ...
    May 15, 2025 · The viscosity ratio accounts for the different viscosity of the flowing phases: when $m\to 0$ , the outer fluid is much more viscous compared ...
  24. [24]
    Dilute or Dense Phase Pneumatic Conveying? - AIChE
    Typically, a dilute phase line conveys less than 15 lb (~7 kg) of solids per pound (~0.5 kg) of gas at a pressure of less than 15 psig (~1 barg). Dense phase.Missing: risers | Show results with:risers
  25. [25]
    More fundamentals of dilute suspension collapse and choking for ...
    Apr 16, 2004 · Using gas-solid flows in circulating fluidized-bed risers to analogize vertical pneumatic conveying systems, the collapse of dilute ...
  26. [26]
    Prediction of the choking velocity and voidage in vertical pneumatic ...
    In vertical gas–solids flow two major flow regimes can be identified—the dilute phase flow and the dense phase flow. At high gas velocity the particles are ...
  27. [27]
    The heterogeneous to homogeneous transition for slurry flow in pipes
    Aug 6, 2025 · ... Wasp (1963) model. The LDV is. defined here as the line speed above which there is no stationary or sliding bed. For graded sands and gravels ...
  28. [28]
    [PDF] Extending estimation of the critical deposition velocity in solid–liquid ...
    The critical deposition velocity in horizontal pipe flow of liquid-solid slurries separates bed-forming and fully suspended flows. A compilation of critical ...
  29. [29]
    A Novel Dehumidification Strategy to Reduce Liquid Fraction and ...
    The condensation process occurs due to the steam expansion in the low-pressure stage and subsequently forms two-phase flows in steam turbines [6,7,8]. The ...
  30. [30]
    [PDF] Two-Phase Gas/Liquid Pipe Flow - AIChE
    Occurs because the gas expands and speeds up relative to the liquid. ▫ It depends upon fluid properties and flow conditions.
  31. [31]
    Computational Modeling of Multiphase Reactors - Annual Reviews
    Jul 2, 2015 · Multiphase reactors are very common in chemical industry, and numerous review articles exist that are focused on types of reactors, such as ...<|separator|>
  32. [32]
    Advanced flow profiler for two-phase flow imaging on distillation trays
    Feb 15, 2021 · The present work proposes a new approach for measuring the effective froth height on column trays. This approach is applied on the two-phase ...
  33. [33]
    [PDF] Chapter SM 7: Evaporators and Condensers - Purdue University
    The liquid refrigerant from the condenser passes through the expansion valve and then enters the evaporator as a low-temperature mixture of vapor and liquid. ...
  34. [34]
    [PDF] Analysis of Two-Phase Flow in Condensers and Evaporators Based ...
    This paper presents a numerical study of two-phase flow in condensers and evaporators using two-fluid models, considering different velocities and temperatures ...<|separator|>
  35. [35]
    Mechanism of enhanced oil recovery of ultra-dry CO2-in-water ...
    Oct 15, 2023 · The application of ultra-dry CO 2 foam in enhanced oil recovery (EOR) accelerates recovery and enables CO 2 utilization and sequestration.
  36. [36]
    [PDF] Development of a Hybrid Single/Two-Phase Capillary-Based Micro ...
    Conventional two-phase microchannel cooling is capable of removing high heat fluxes for current- and next-generation electronics [4], but it is difficult to ...
  37. [37]
    [PDF] A Study of Terrain-Induced Slugging in Pipelines Using Aspen ...
    Hydrodynamic slugging is caused by differences in the velocities of the gas and liquid phases in a two-phase flow. When the relative velocity between the two ...
  38. [38]
    Magma flow along the volcanic conduit during the Plinian and ...
    The ascent of magma during the May 18, 1980, Mount St. Helens eruption was modeled by means of a nonequilibrium two-phase flow model which accounts for the ...
  39. [39]
    A general two‐phase debris flow model - Pudasaini - AGU Journals
    Aug 1, 2012 · A general two-phase debris flow model. Shiva P. Pudasaini ... Download PDF. back. Back to Top. AGU Journals Logo. © 2025 American ...
  40. [40]
    Void fraction
    ### Summary of Void Fraction from https://www.thermopedia.com/content/276/
  41. [41]
    8.1 Behavior of Two-Phase Systems - MIT
    ... m_g. The fraction of the total mass in the vapor phase is called quality, and denoted by $ X$ : $\displaystyle X = \frac{m_g}{m_f + m_g} = In terms of the ...
  42. [42]
    Void fraction measurement
    ### Summary of Void Fraction Measurement Methods
  43. [43]
    Experimental investigation and new void-fraction calculation method ...
    Feb 1, 2021 · The void fraction is a key parameter of the two-phase flow in vertical downward pipes, particularly for calculating the mixture density, ...
  44. [44]
    Effect of Void Fraction on Pressure Drop in Upward Vertical Two ...
    Aug 5, 2025 · In gas–liquid two-phase flow, the prediction of two-phase density and hence the hydrostatic pressure drop relies on the void fraction and is ...
  45. [45]
    [PDF] Two-Phase Flow - IntechOpen
    The quality (dryness fraction) (x) is defined as the ratio of the mass flow rate of gas phase. ( g mq ) to the total mass flow rate ( mq ). g g l g m m x m.Missing: (m_g +
  46. [46]
    The liquid film and the core region velocity profiles in annular two ...
    The nondimensional velocity profile depends explicitly on the nondimensional distance from the interface, the entrainment and the nondimensional film thickness.
  47. [47]
    Two phase flow pressure drop measurement analysis in a horizontal ...
    Apr 3, 2024 · To measure the pressure drop, a differential pressure transducer is used and installed between the T-junction and downstream areas.
  48. [48]
    Two-phase choked flow of trans critical carbon dioxide at the micro ...
    Jan 14, 2025 · Choking is dictated by the sonic velocity at the outlet of the microchannel and occurs when the velocity approaches the speed of sound. The flow ...
  49. [49]
    Experimental and numerical characterisation of two-phase flow in ...
    This work applies a 2D two-phase Computational Fluid Dynamics (CFD) model based on the Volume of Fluid (VOF) approach to simulate the flow of two immiscible ...
  50. [50]
    Thermo-Fluid Dynamics of Two-Phase Flow - SpringerLink
    This book covers the fundamental physics of two-phase flow, including modeling, code development, and applications in nuclear reactors, energy, and chemical ...
  51. [51]
    A novel technique for including surface tension in PLIC-VOF methods
    We present here a new method to determine curvature accurately using an estimator function, which is tuned with a least-squares-fit against reference data.Missing: original | Show results with:original
  52. [52]
    ANSYS FLUENT 12.0 Theory Guide - 16. Multiphase Flows
    This chapter discusses the general multiphase models that are available in ANSYS FLUENT. Section 16.1 provides a brief introduction to multiphase modeling.<|control11|><|separator|>
  53. [53]
    [PDF] Lagrangian–Eulerian methods for multiphase flows
    This paper primarily focuses on the use of the LE approach as a solution method for the transport equation of the droplet distribution function (ddf) or number.
  54. [54]
    Data-driven closure model selection for multiphase CFD via matrix ...
    This paper addresses the problem of closure model selection in multiphase CFD by adapting machine-learning methods from information-filtering systems. ... enhance ...
  55. [55]
    (PDF) Machine Learning Reshaping Computational Fluid Dynamics
    Oct 28, 2025 · The findings demonstrate that ML techniques can accelerate simulations by up to 10,000 times in certain cases while maintaining or improving the ...
  56. [56]
    [PDF] Sound Propagation in Bubbly Liquids. A Review - DTIC
    Apr 5, 1989 · A derivation of phase speed from first principles indicates that Wood's equation is an accurate approximation for dilute mix- tures of gas ...
  57. [57]
    Finite amplitude wave propagation through bubbly fluids
    Defining the Mach number as the ratio between the characteristic speed of the bubble collapse and the linear speed of sound in the bubbly liquid, M a = 1 c w o ...
  58. [58]
    [PDF] Experimental investigation of the pressure distribution during - HAL
    ... bubbles causes a dramatic reduction of the speed of sound from about 1500 m/s to less than 100 m/s 21. For example, during violent wave impact on a ...
  59. [59]
    An audible demonstration of the speed of sound in bubbly liquids
    Oct 1, 2008 · In this section, the equation governing the low-frequency speed of sound in bubbly liquids is developed, based on the work of Wood. c 2 = 1 ρ κ ...INTRODUCTION · Low frequency sound... · VERIFICATION OF WOOD'S...Missing: formula | Show results with:formula
  60. [60]
    Mechanisms of bubble effects on acoustic propagation ...
    This study reveals the factors influencing bubble damping and how it regulates acoustic attenuation, providing insight into how acoustic waves propagate in gas ...
  61. [61]
    Propagation of pressure waves in two-phase flow - ScienceDirect.com
    We study the effects of pipe elasticity on the propagation velocity of the pressure wave. Pressure waves of finite amplitude progressing in the two-phase flow ...
  62. [62]
    The generation of sound by two-phase nozzle flows and its ...
    Mar 29, 2006 · This paper describes a prototype model experiment designed to test the principle that the 'excess' noise of a jet issuing from a conical ...
  63. [63]
    On the Use of Acoustic Waves in Nuclear Power Reactors to ...
    May 12, 2017 · A method for estimating the average void fraction in water-moderated power reactors is presented. The method might be useful to determine ...
  64. [64]
    [PDF] High-Speed Ultrasonic Pulse-Echo System for Two-Phase Flow ...
    By this technique, it is possible to measure moderately transient bubbly flow of the order of Is time intervals, and with a spatial resolution of 0.075 mm.
  65. [65]
    Acoustic spectrometry of bubbles in an estuarine front: Sound speed ...
    Apr 7, 2022 · The 10–20 kHz-derived sound speed approaches 1400 m/s while the 20–100 kHz-derived sound speed approaches ∼800 m/s at ∼1637Z and ∼1200 m/s at ∼ ...
  66. [66]
    A review of flow boiling heat transfer: Theories, new methods and ...
    Nucleate boiling is an important stage in pool boiling. Natural convection, transition boiling, and film boiling are the other three stages in the boiling ...
  67. [67]
    [PDF] Microgravity flow boiling experiments with liquid-vapor mixture inlet ...
    Feb 23, 2024 · This article is part of a series of studies culminating from the multi-objective Flow Boiling and Condensation.
  68. [68]
    [PDF] CFD Simulation of Laminar Film Condensation on a Vertical Surface
    The most well-known work within the first group is the Nusselt theory (1916), who was the first to analyse the film condensation on a vertical wall. This ...
  69. [69]
    [PDF] Ranz W E & Marshall W R, Jr. Evaporation from drops. Parts I &
    May 31, 1993 · An investigation was made of the factors influencing the rate of evaporation of pure liquid drops, and the rate of evaporation of water ...
  70. [70]
    Recent advances in two-phase fluid experiments in microgravity
    In 2020, Liu et al. [32] created a technique to use ground-based experiments to evaluate subcooled flow boiling in microgravity. The authors suggest conducting ...Missing: post- | Show results with:post-
  71. [71]
    Porous Coating Effect on Two-Phase Heat Transfer Behaviour of ...
    Sep 3, 2025 · Coatings with porosity levels ranging from 55 to 60% result in a rise in the coefficient of heat transfer, which can vary from 33 to 60%. In ...
  72. [72]
    Flow and heat transfer for a two-phase slug flow in horizontal pipes
    The heat transfer enhancement of two-phase slug flow compared with single-phase flow is mainly attributed to the turbulence increase in liquid by the injection ...