Fact-checked by Grok 2 weeks ago

Heat transfer coefficient

The heat transfer coefficient, often denoted as h or \alpha, is a proportionality factor that characterizes the rate of convective between a and an adjacent , as expressed in : q'' = h (T_s - T_f), where q'' is the , T_s is the surface temperature, and T_f is the temperature. This coefficient quantifies the effectiveness of exchange in processes involving flow over surfaces, such as in , exchangers, and cooling systems, and its value depends on (e.g., , thermal conductivity), flow characteristics (e.g., , ), and surface . In engineering applications, the convective heat transfer coefficient typically ranges from 10–100 W/m²·K for gases like air in free convection to 500–10,000 W/m²·K for liquids like water in forced convection, with higher values in turbulent flows due to enhanced mixing and a thinner boundary layer. For composite systems involving multiple layers or fluids, the overall heat transfer coefficient U is used, defined by Q = U A \Delta T, where Q is the total heat transfer rate and it accounts for both convective resistances on either side and conductive resistances through walls, calculated as \frac{1}{U} = \frac{1}{h_i} + \sum \frac{\delta}{k} + \frac{1}{h_o} for plane walls. Dimensionless groups like the (Nu = \frac{[h](/page/H+) L}{k}), , and are for predicting [h](/page/H+) in various flow regimes, correlations for in systems. The for both [h](/page/H+) and U is watts per square meter per (W/·), reflecting the per area per temperature difference, with common conversions to like Btu/(h·ft²·°F).

Fundamentals

Definition and Newton's Law of Cooling

The heat transfer coefficient, denoted as h, is defined as the proportionality factor between the convective heat flux at a solid surface and the temperature difference driving the heat transfer process. In the context of Newton's law of cooling, the rate of heat transfer q from a surface of area A to a surrounding fluid is given by q = h A (T_s - T_\infty), where T_s is the temperature of the surface and T_\infty is the temperature of the fluid far from the surface. This formulation quantifies the convective heat transfer under conditions where the fluid motion enhances heat removal compared to pure conduction. Isaac Newton first articulated the foundational idea in 1701 in his paper "Scala graduum Caloris," published in the Philosophical Transactions of the Royal Society, where he stated that the rate of cooling of a warm body suspended in cooler air is proportional to the difference in temperature between the body and the surrounding air. This empirical observation laid the groundwork for convective heat transfer analysis, though Newton's original experiments involved natural convection with limited control over flow conditions. Modern interpretations refine the law to address non-linear temperature dependencies that arise in cases of large temperature gradients, where radiation or variable fluid properties may influence the process. Physically, h represents the effectiveness of convection in bridging the thermal resistance between the surface and the bulk fluid, encapsulating the combined effects of molecular diffusion near the surface and bulk fluid motion. Its magnitude depends on intrinsic fluid properties, such as thermal conductivity, viscosity, and density, as well as extrinsic flow characteristics like velocity and turbulence intensity, which determine how efficiently heat is carried away from the surface. Higher values of h indicate more vigorous convective transport, often achieved through forced flow or enhanced surface features. The derivation of h conceptually stems from equating the conductive heat flux at the fluid-solid interface, as described by Fourier's law of conduction, to the overall convective heat flux. Fourier's law gives the local heat flux as q'' = -k \frac{\partial T}{\partial y} \big|_{y=0}, where k is the fluid's thermal conductivity and y is the direction normal to the surface. Setting this equal to the convective expression q'' = h (T_s - T_\infty) yields h = -\frac{k \frac{\partial T}{\partial y} \big|_{y=0}}{T_s - T_\infty}, illustrating how h normalizes the near-wall temperature gradient to the far-field driving potential. This approach assumes steady-state conditions, constant thermophysical properties, negligible radiative contributions, and a linear temperature profile approximation valid for small Biot numbers.

Local versus Average Coefficients

The local coefficient, h(x), represents the convective at a specific point on a surface and is defined by as h(x) = \frac{q''(x)}{T_s(x) - T_\infty}, where q''(x) is the local , T_s(x) is the local surface , and T_\infty is the free-stream . This coefficient varies spatially along the surface due to the of the thermal in convective flows; as the thickens in the downstream direction, the at the wall decreases, resulting in a corresponding reduction in h(x). For instance, in external flows over a flat plate, h(x) is highest near the leading edge where the is thinnest and progressively diminishes along the plate length. In contrast, the average heat transfer coefficient, \bar{h}, provides a spatially integrated measure suitable for overall performance evaluation and is computed as the surface integral of the local coefficient divided by the total surface area: \bar{h} = \frac{1}{A} \int_A h \, dA. For one-dimensional flows, such as along a plate of length L, this reduces to \bar{h} = \frac{1}{L} \int_0^L h(x) \, dx. Local coefficients are employed in detailed design scenarios, such as predicting temperature hotspots or optimizing non-uniform surface conditions, whereas average coefficients facilitate practical calculations of total rates across entire surfaces. In developing flows, like the entrance region of a duct, h(x) starts high due to thin layers and decreases toward a constant fully developed value, making the average \bar{h} typically higher than the local value at the trailing edge. For cases where the bulk fluid temperature varies significantly along the surface, such as in heat exchangers, the average coefficient is often determined using the (LMTD) to for non-constant driving potentials: the total Q = \bar{h} A \Delta T_{lm}, with \Delta T_{lm} = \frac{\Delta T_1 - \Delta T_2}{\ln \left( \frac{\Delta T_1}{\Delta T_2} \right)}, where \Delta T_1 and \Delta T_2 are the temperature differences at the and outlet. This approach yields an effective \bar{h} that represents the integrated performance without requiring point-wise integration of varying local values.

Dimensionless Framework

Nusselt Number and Its Relation to h

The Nusselt number, denoted as \mathrm{Nu}, is defined as the dimensionless form of the heat transfer coefficient h, given by the equation \mathrm{Nu} = \frac{h L}{k}, where L is the characteristic length of the surface (such as the plate length or pipe diameter) and k is the thermal conductivity of the fluid. This formulation arises in convective heat transfer problems to normalize the convective resistance against conductive resistance across the fluid boundary layer. Physically, the Nusselt number represents the ratio of the total heat transfer (convective plus conductive) to the purely conductive heat transfer across a fluid layer of thickness equal to the characteristic length L. A value of \mathrm{Nu} = 1 indicates heat transfer dominated by conduction alone, as in a stagnant fluid, while \mathrm{Nu} > 1 signifies enhancement due to convection, with higher values corresponding to more effective convective mixing within the boundary layer. This interpretation underscores the Nusselt number's role in quantifying the relative importance of convection in boundary layer flows. The emerges from of the convection equations, where the heat transfer coefficient h is expressed in terms of fluid properties, , and . By applying the to the energy equation and boundary conditions, the dimensionless temperature gradient at the wall yields \mathrm{Nu} as the key group linking convective to conduction, directly tied to the thermal . This approach connects to Prandtl's theory, where effects scale with velocity and temperature profiles near the surface. Similar to the heat transfer coefficient, the Nusselt number can be defined locally or as an average over a surface. The local \mathrm{Nu}_x at position x is \mathrm{Nu}_x = \frac{h_x x}{k}, capturing variations along the flow direction due to development. The average Nusselt number \overline{\mathrm{Nu}}_L over length L is then obtained by integration: \overline{\mathrm{Nu}}_L = \frac{1}{L} \int_0^L \mathrm{Nu}_x \, dx, which parallels the averaging for h. From the definition, the heat transfer coefficient is explicitly recovered as h = \frac{\mathrm{Nu} \, k}{L}, allowing empirical or theoretical correlations for \mathrm{Nu} (often as functions of Reynolds and Prandtl numbers) to directly provide values of h for engineering calculations. These correlations, derived from experiments or similarity solutions, thus bridge dimensionless analysis to practical convective heat transfer rates. The formulation assumes constant fluid thermal conductivity k and negligible viscous effects, typically valid when the Eckert number \mathrm{Ec} \ll 1, ensuring that does not significantly alter the . These assumptions align with low-speed flows and moderate differences in standard problems.

Key Dimensionless Groups (Reynolds, Prandtl, Grashof)

In the context of , dimensionless groups provide a framework for scaling and correlating convective heat transfer coefficients by capturing the essential physics of fluid flow and thermal transport without dependence on specific units. The Reynolds, Prandtl, and Grashof numbers are pivotal in this framework, particularly for expressing the (), which directly relates to the heat transfer coefficient h through \mathrm{Nu} = \frac{h L}{k}, where L is a and k is the fluid thermal conductivity. These groups arise from and experimental correlations, enabling predictions of h across different geometries, fluids, and flow conditions. The Reynolds number (\mathrm{Re}), introduced by Osborne Reynolds in his 1883 experimental investigation of pipe flow transitions, quantifies the ratio of inertial forces to viscous forces in a fluid flow. It is defined as \mathrm{Re} = \frac{\rho u L}{\mu} = \frac{u L}{\nu}, where \rho is fluid density, u is a characteristic velocity, L is a characteristic length, \mu is dynamic viscosity, and \nu is kinematic viscosity. Physically, low \mathrm{Re} (typically below 2300 for pipe flows) indicates laminar flow dominated by viscous effects, while high \mathrm{Re} signals turbulent flow where inertia promotes mixing and enhanced heat transfer. In forced convection correlations for the Nusselt number, \mathrm{Re} determines the flow regime and is a primary independent variable, as in the Dittus-Boelter equation for turbulent pipe flow: \mathrm{Nu} = 0.023 \mathrm{Re}^{0.8} \mathrm{Pr}^{0.4}, which scales h with flow speed and geometry. This role underscores \mathrm{Re}'s influence on boundary layer development and convective enhancement in external and internal flows. The Prandtl number (\mathrm{Pr}), named after Ludwig Prandtl for his foundational 1904 boundary layer theory, measures the relative thickness of the momentum (velocity) boundary layer to the thermal boundary layer in convective flows. It is defined as \mathrm{Pr} = \frac{\nu}{\alpha} = \frac{\mu c_p}{k}, where \alpha is thermal diffusivity and c_p is specific heat capacity at constant pressure. For gases, \mathrm{Pr} \approx 0.7; for water, \mathrm{Pr} \approx 7; and for oils, \mathrm{Pr} > 100, reflecting how momentum diffuses faster than heat in low-\mathrm{Pr} fluids, leading to thicker thermal layers and potentially reduced h. In both forced and natural convection, \mathrm{Pr} appears in Nusselt correlations to account for fluid property effects on heat diffusion relative to momentum transport, such as in the aforementioned Dittus-Boelter form or natural convection relations like \mathrm{Nu} = C (\mathrm{Gr} \mathrm{Pr})^n, where it modulates the overall convective resistance. Prandtl's boundary layer concept highlights how \mathrm{Pr} influences the interplay between conduction within the layer and bulk convection, critical for accurate h predictions across fluid types. The (\mathrm{Gr}), attributed to Franz Grashof's early contributions to in the late , characterizes the ratio of -driven forces to viscous forces in natural convection, where flow arises from differences due to temperature gradients. It is defined as \mathrm{Gr} = \frac{g \beta \Delta T L^3}{\nu^2}, where g is , \beta is the thermal expansion coefficient, and \Delta T is the temperature difference driving . High \mathrm{Gr} (e.g., >10^9) promotes turbulent natural convection, enhancing h through stronger circulation, while low \mathrm{Gr} yields laminar regimes. In correlations, \mathrm{Gr} replaces \mathrm{Re} for buoyancy-dominated flows, often combined with \mathrm{Pr} into the (\mathrm{Ra} = \mathrm{Gr} \mathrm{Pr}) for sselt expressions, such as \mathrm{Nu} = 0.59 \mathrm{Ra}^{1/4} for laminar vertical plates, directly linking h to geometry, temperature difference, and fluid properties. This group's role is essential for applications like electronics cooling or solar collectors, where forced flow is absent.

Natural Convection Correlations

Vertical Flat Plates and Planes

In natural convection heat transfer from vertical flat plates and planes, buoyancy-driven flow arises from density variations caused by temperature differences between the surface and the surrounding , leading to the formation of a along the plate. The velocity boundary layer, where fluid velocity adjusts from zero at the wall to the free-stream value, develops concurrently with the thermal boundary layer, where gradients occur. For low fluids (Pr ≪ 1), such as liquid metals, the thermal boundary layer is thicker than the velocity boundary layer due to higher relative to momentum diffusivity, resulting in broader profiles and reduced local heat transfer rates compared to high-Pr fluids like oils. Empirical correlations for the average heat transfer coefficient h on vertical flat plates express it through the average \overline{\mathrm{Nu}}_L = \frac{h L}{k}, where L is the plate height and k is the fluid thermal conductivity. The \mathrm{Ra}_L = \mathrm{Gr}_L \mathrm{Pr} governs the flow regime, incorporating the \mathrm{Gr}_L (ratio of buoyancy to viscous forces) and \mathrm{Pr} (ratio of momentum to ). For over $10^4 < \mathrm{Ra}_L < 10^9, the correlation for isothermal surfaces is \overline{\mathrm{Nu}}_L = 0.59 \mathrm{Ra}_L^{1/4}, while turbulent transition occurs around \mathrm{Ra}_L \approx 10^9, yielding \overline{\mathrm{Nu}}_L = 0.10 \mathrm{Ra}_L^{1/3} for higher Rayleigh numbers. These relations follow the general form h = \frac{k}{L} C \mathrm{Ra}_L^n, with constants C and exponent n varying by regime (C = 0.59, n = 1/4 for laminar; C = 0.10, n = 1/3 for turbulent). For uniform heat flux boundary conditions, where the surface heat flux q'' is constant rather than the temperature, the correlations adopt similar forms but require adjustment, often using an effective temperature difference derived iteratively from q'' = h (T_s - T_\infty); dedicated expressions, such as those approximating \overline{\mathrm{Nu}}_L \approx 0.60 \mathrm{Ra}_L^{1/4} for laminar cases, account for the altered temperature profile and yield heat transfer coefficients about 5-15% higher than isothermal predictions. These distinctions arise because uniform flux maintains a monotonically increasing surface temperature along the plate, thickening the boundary layer differently than the fixed-temperature case. The correlations stem from boundary layer analyses and experiments with gases like air and liquids such as water and oils, validated across Prandtl numbers from 0.7 to 10,000 and plate heights up to several meters; they provide engineering accuracy of ±10-20% within the specified Rayleigh ranges, with the Churchill-Chu framework encompassing both laminar and turbulent regimes for broad applicability.

Horizontal Cylinders and Plates

In natural convection from horizontal surfaces, the orientation significantly influences flow stability and heat transfer rates. For upward-facing hot plates or downward-facing cold plates, the buoyancy forces create an unstable configuration where lighter fluid rises, promoting the development of thermal plumes that enhance mixing and heat transfer compared to stable orientations. For horizontal plates, the average Nusselt number for the upper surface of a hot plate (or lower surface of a cold plate) is correlated as Nu_L = 0.54 Ra_L^{1/4} for 10^4 < Ra_L < 10^7, where Ra_L is the Rayleigh number based on the characteristic length L, typically taken as the plate's side length for square plates or the average of length and width for rectangular ones. This relation, recommended by McAdams, yields the average heat transfer coefficient h via h = (Nu_L k)/L, with k as the fluid thermal conductivity, and applies to fluids like air where Prandtl numbers are around 0.7. For broader applicability, Churchill extended similar forms to account for edge effects and varying plate sizes, emphasizing L as the hydraulic diameter or area-to-perimeter ratio for non-rectangular shapes to capture diameter or length influences on boundary layer development. Horizontal cylinders exhibit analogous behavior, with the plume forming along the top surface. The widely adopted correlation by for the average over the cylinder surface, valid for a broad range of (10^{-2} < Ra_D < 10^{12}) and (0.5 < Pr < ∞), is \overline{\mathrm{Nu}}_D = \left\{ 0.60 + \frac{0.387 \mathrm{Ra}_D^{1/6}}{\left[1 + \left( \frac{0.559}{\mathrm{Pr}} \right)^{9/16} \right]^{8/27}} \right\}^2, where the subscript D denotes the cylinder diameter as the characteristic length. This expression provides the average h as h = (\overline{\mathrm{Nu}}_D k)/D and outperforms earlier laminar-only models by incorporating turbulent transitions. McAdams' earlier recommendations for cylinders align closely for moderate Ra but are limited to narrower ranges. These correlations are essential in engineering applications such as horizontal pipes conveying fluids in ambient air, where natural convection governs cooling, and extended surface heat exchangers like finned tubes, optimizing designs for passive thermal management without forced flow. The Rayleigh number determines the regime applicability, with lower values indicating laminar plume-dominated flow.

Spheres and Vertical Enclosures

In natural convection around spheres, the heat transfer coefficient h is determined using Nusselt number correlations based on the Rayleigh number \mathrm{Ra}_D, where the characteristic length is the sphere diameter D. For laminar flow at low Rayleigh numbers, a widely used correlation is \overline{\mathrm{Nu}}_D = 2 + 0.43 \mathrm{Ra}_D^{1/4}, valid for \mathrm{Ra}_D \lesssim 10^9 and applicable to gases like air. This form approaches the conduction limit of \mathrm{Nu}_D = 2 as \mathrm{Ra}_D \to 0, reflecting pure diffusive heat transfer across the stagnant fluid layer surrounding the sphere. At higher Rayleigh numbers leading to turbulent flow (\mathrm{Ra}_D > 10^9), the correlation extends to forms like \overline{\mathrm{Nu}}_D = 2 + 0.50 \mathrm{Ra}_D^{1/4}, accounting for enhanced mixing and instability, though more comprehensive expressions incorporate dependence for broader fluids. The heat transfer coefficient is then obtained as h = \overline{\mathrm{Nu}}_D k / D, where k is the fluid thermal . For vertical enclosures, such as those formed by parallel plates with heated and cooled vertical walls separated by a small gap, natural convection is characterized by counterflowing boundary layers along the walls, leading to an effective thermal conductivity k_\mathrm{eff} that quantifies the enhanced heat transfer across the enclosure. The characteristic length is the plate spacing L, and for aspect ratios H/L \approx 1 (where H is the enclosure height), a key correlation is k_\mathrm{eff}/k = 0.22 \left[ \Pr \mathrm{Ra}_L / (0.2 + \Pr) \right]^{0.28}, applicable for $10^4 < \mathrm{Ra}_L < 10^7 and moderate Prandtl numbers typical of air or water. This Nusselt number-based form, where \mathrm{Nu}_L = h L / k = (k_\mathrm{eff}/k) \cdot (H/L), assumes H \approx L for square-like cavities, yielding h = (k_\mathrm{eff}/k) \cdot k / H. Confinement in narrow gaps reduces h compared to isolated plates due to boundary layer interference, suppressing circulation and limiting the convective enhancement to near-conductive levels at very small L. These correlations originate from experimental studies on isothermal boundaries and are validated for applications like building insulation panels and electronic component cooling, where buoyancy-driven flows dominate without external forcing.

Forced Convection Correlations

External Flows over Bodies

In external flows over bodies, forced convection heat transfer coefficients are determined using the Nusselt number (Nu), which relates the convective heat transfer to conduction across the fluid boundary layer, with the coefficient given by h = \frac{Nu \cdot k}{L}, where k is the fluid thermal conductivity and L is a characteristic length (e.g., plate length or body diameter). The Reynolds number (Re) is based on the free-stream velocity U_\infty and characterizes the flow regime, influencing boundary layer development and thus h. Laminar flows occur at low Re (e.g., Re < 5 \times 10^5 for flat plates), where viscous forces dominate and the boundary layer remains smooth, leading to lower h; transitional regimes (around 5 \times 10^5 < Re < 10^6) exhibit intermittent turbulence, increasing h; and turbulent flows at higher Re promote mixing, significantly enhancing h through chaotic eddies. Prandtl number (Pr) effects are captured in correlations, typically showing h increasing with Pr^{1/3} for Pr > 0.6. For a flat plate in laminar parallel to the surface, the local at distance x from the is Nu_x = 0.332 Re_x^{1/2} Pr^{1/3} for constant wall temperature and Pr ≥ 0.6, derived from the similarity solution to the boundary layer energy equation. The average over length L is then Nu_L = 0.664 Re_L^{1/2} Pr^{1/3} for Re_L < 5 \times 10^5, doubling the local value at the trailing edge due to growth. These apply to isothermal or uniform heat flux conditions in gases and liquids, providing h values that scale with U_\infty^{1/2}, emphasizing the role of developing thermal s in external flows. Cross-flow over cylinders, common in heat exchanger design, uses the Churchill-Bernstein correlation for the average Nusselt number:
Nu_D = 0.3 + \frac{0.62 Re_D^{1/2} Pr^{1/3}}{\left[1 + (0.4/Pr)^{2/3}\right]^{1/4}} \left[1 + \left(\frac{Re_D}{282000}\right)^{5/8}\right]^{4/5}
for Re_D Pr > 0.2, valid across laminar, transitional, and turbulent regimes up to Re_D ≈ 10^7 and 0.5 < Pr < 10^2, based on diameter D. This empirical fit unifies prior data, showing h rising sharply with Re_D due to boundary layer separation and wake formation, with turbulent effects dominating at Re_D > 10^3.
For spheres in uniform cross-flow, the Whitaker correlation gives the average as
Nu_D = 2 + \left[ 0.4 Re_D^{1/2} + 0.06 Re_D^{2/3} \right] [Pr](/page/PR)^{0.4} \left( \frac{\mu_\infty}{\mu_s} \right)^{1/4}
for $3.5 \leq Re_D \leq 7.6 \times 10^4, $0.71 \leq [Pr](/page/PR) \leq 380, and $1 \leq \frac{\mu_\infty}{\mu_s} \leq 3.2, where \mu_\infty and \mu_s are the viscosities at free-stream and surface temperatures, respectively (the is 1 for constant properties). This extends the stagnant limit (Nu = 2) with enhancement. Here, h = Nu_D k / D reflects symmetric growth until separation at the rear, with lower h than cylinders due to reduced and wake intensity.
In cylinder cross-flows, unsteadiness from (prominent at 40 < Re_D < 10^5) periodically disrupts the wake, enhancing local h on the rear surface by up to 20-30% through improved mixing, though average correlations like Churchill-Bernstein account for this statistically.

Internal Flows in Pipes and Ducts

In internal flows through pipes and ducts, forced convection heat transfer is characterized by correlations expressing the Nusselt number Nu as a function of the Reynolds number Re and Prandtl number Pr, applicable to fully developed conditions where the velocity and temperature profiles are established. These correlations are essential for predicting the convective heat transfer coefficient h in applications such as heat exchangers and cooling systems, distinguishing laminar regimes (typically Re < 2300) from turbulent ones (Re > 10,000). The characteristic length is the pipe diameter D for circular geometries, enabling the relation h = \frac{k}{D} \mathrm{Nu}, where k is the fluid thermal conductivity evaluated at the bulk mean temperature T_b, defined as the arithmetic average of inlet and outlet temperatures. For laminar flow in circular ducts with a fully developed parabolic velocity profile, the attains constant values independent of Re and Pr once the thermal profile is also fully developed. Under a constant wall temperature boundary condition, Nu = 3.66, derived from the exact solution to the Graetz problem for hydrodynamically developed flow. For constant wall heat flux, the value is Nu = 4.36, reflecting the uniform energy input that sustains a linear axial temperature rise in the bulk fluid. These constants hold for gases and liquids with moderate Prandtl numbers. In the turbulent regime, mixing enhances , and the Dittus-Boelter correlation provides a simple empirical expression: \mathrm{Nu} = 0.023 \, \mathrm{Re}^{0.8} \, \mathrm{Pr}^{n} where n = 0.4 for heating the fluid (wall hotter than bulk) and n = 0.3 for cooling, valid for Re > 10,000, 0.7 < Pr < 160, and pipe lengths with L/D > 10 to ensure fully developed conditions. This , based on experimental data for smooth tubes, captures the dominant influence of on the . For improved accuracy across transitional and low-turbulence cases (3000 < Re < 10^5, 0.5 < Pr < 2000), the Gnielinski refines predictions by incorporating a friction factor term, yielding results within ±10% of measurements in many engineering scenarios. Entrance effects significantly elevate h in the developing region near the pipe inlet, where the thermal is thin and growing. The thermal entry length, over which decreases to its fully developed value, is quantified by the Graetz number Gz = \frac{\mathrm{} , \mathrm{} , D}{L}, with developing flow prevailing for Gz > 10 and higher (up to 2-3 times the fully developed value) due to the thinner . For laminar flows, the entry length is approximately L/D ≈ 0.05 , emphasizing the need for longer pipes in low- applications to achieve conditions. For non-circular ducts, such as rectangular or annular channels, the hydraulic diameter D_h = \frac{4A}{P} (where A is cross-sectional area and P is wetted perimeter) serves as the effective length scale in Re, Pr, and Nu definitions, allowing standard circular-pipe correlations to be applied with reasonable approximation. However, duct shape influences secondary flows and friction, necessitating shape-specific correction factors for Nu in precise calculations, particularly in laminar regimes where deviations can exceed 20% without them. Fluid properties remain evaluated at T_b for consistency across geometries. Overall, these correlations provide engineering accuracy of ±25% for purposes, balancing simplicity and reliability, though experimental validation is recommended for conditions like high or variable .

Special Correlations and Cases

Thom Correlation for Boiling

The Thom correlation offers an empirical expression for estimating the nucleate boiling heat transfer coefficient in pool boiling scenarios involving at . It is given by the formula h = 5.56 (\Delta T)^3 where h is the heat transfer coefficient in W/m²K and \Delta T is the wall superheat (temperature difference between the heated surface and the saturation temperature) in °C. This relation is valid within the range of 5 < \Delta T < 25°C, corresponding to the fully developed nucleate boiling regime where bubble formation and detachment enhance heat transfer without reaching transition or film boiling. The correlation stems from experimental measurements on horizontal heating surfaces, capturing the influence of bubble dynamics on convective enhancement during . Bubble , growth, and departure create mixing in the , leading to the characteristic cubic dependence on superheat observed in the fitted data. Substituting into , the is obtained as q = h \Delta T = 5.56 (\Delta T)^4 W/m², which remains applicable up to but not exceeding the , beyond which departure from occurs and deteriorates. This correlation is limited to as the under low-pressure conditions near atmospheric levels and does not extend to flow regimes or other liquids, where additional factors like velocity or fluid properties alter the behavior. Developed in the , the Thom correlation emerged from research aimed at supporting design, particularly for predicting surface temperatures in boiling systems to ensure safe operation margins.

Radiation and Extended Applications

The radiative heat transfer coefficient, denoted as h_r, provides a linearized approximation for quantifying radiative heat flux in terms of a temperature difference, analogous to convective or conductive forms, thereby simplifying combined-mode analyses. This coefficient arises from the Stefan-Boltzmann law, which governs the net radiative heat flux q_{\text{rad}} between a surface at temperature T_s and its surroundings at T_{\text{sur}} as q_{\text{rad}} = \varepsilon \sigma (T_s^4 - T_{\text{sur}}^4), where \varepsilon is the surface emissivity and \sigma = 5.67 \times 10^{-8} W/m²·K⁴ is the Stefan-Boltzmann constant. To linearize this nonlinear expression for practical engineering calculations, the flux is recast as q_{\text{rad}} = h_r (T_s - T_{\text{sur}}), yielding the derivation: h_r = \varepsilon \sigma (T_s + T_{\text{sur}}) (T_s^2 + T_{\text{sur}}^2) This form assumes small temperature differences relative to absolute temperatures, ensuring the approximation holds within typical engineering ranges (e.g., |T_s - T_{\text{sur}}| \ll T_s \approx T_{\text{sur}}). Key assumptions include gray-body behavior, where emissivity \varepsilon is constant and independent of wavelength, and diffuse radiation, meaning emitted and reflected rays are isotropic rather than specular. The radiative coefficient h_r is particularly relevant in scenarios with negligible convection, such as vacuum environments or rarefied gases, where radiation dominates heat transfer. In space applications, for instance, spacecraft thermal control relies solely on for heat rejection in the , with h_r guiding surface coatings and designs to manage temperatures near 2.7 K cosmic . When is present but low, as in insulating gases or natural at low velocities, the total effective heat transfer coefficient combines modes additively: h = h_{\text{conv}} + h_r, allowing unified treatment in overall thermal resistance calculations. An analogous concept exists for conduction across thin films or layers, where the conductive heat transfer coefficient is defined as h_{\text{cond}} = k / \delta, with k as thermal conductivity and \delta as layer thickness; however, this is rarely used as a standalone coefficient outside specialized thin-film contexts, as conduction is typically handled via Fourier's law directly. Modern extensions of the heat transfer coefficient concept incorporate advanced materials and geometries to enhance performance empirically. In nanofluids—colloidal suspensions of nanoparticles in base fluids—post-2000 research has demonstrated enhancements in convective heat transfer coefficients by 10–50% at low particle concentrations (e.g., 1–5 vol.%), attributed to mechanisms like and , though results vary by fluid type and flow regime. Similarly, microchannel flows, prevalent in electronics cooling since the early 2000s, yield significantly higher coefficients (often 10,000–100,000 W/m²·K) due to increased surface-to-volume ratios and , with recent optimizations like hybrid surfaces achieving up to 2–3 times conventional values. These empirical advances address limitations in traditional correlations by integrating nanoscale effects and miniaturization, expanding applicability to high-heat-flux systems.

Engineering Applications

Heat Transfer in Pipe Walls

In pipe systems, the convective heat transfer coefficient h at the inner and outer surfaces is combined with the conductive resistance of the to determine the overall rate from the internal at T_i to the external at T_o. This involves both convective processes at the fluid- interfaces and steady-state radial conduction through the cylindrical . For radial conduction in a cylindrical wall, the rate Q under steady-state, one-dimensional conditions follows from Fourier's law, yielding Q = \frac{2 \pi k L (T_{wi} - T_{wo})}{\ln(r_o / r_i)}, where k is the thermal conductivity of the wall material, L is the length, r_i and r_o are the inner and outer radii, and T_{wi} and T_{wo} are the inner and outer wall temperatures. The corresponding thermal resistance due to is R_\text{wall} = \frac{\ln(r_o / r_i)}{2 \pi k L}. This logarithmic form accounts for the varying cross-sectional area with radius in cylindrical geometry. When the wall is thin—typically when the thickness \delta = r_o - r_i is much smaller than the mean radius—the conduction can be approximated as planar, with resistance R_\text{wall} \approx \frac{\delta}{2 \pi r_m k L}, where r_m = (r_i + r_o)/2, or equivalently, an effective conductance h_\text{cond} = k / \delta. In this case, the inner wall temperature T_w satisfies T_w \approx \frac{h_i T_i + (h_o T_o) / (k / \delta)}{h_i + h_o / (k / \delta)}, allowing the local convective heat flux at the inner surface q_i = h_i (T_i - T_w) to incorporate the wall's temperature drop. For thicker walls, the full radial solution is required to avoid underestimating the resistance, as the logarithmic term grows with \delta / r_i. The overall heat transfer rate, focusing on the wall's contribution, is Q = \frac{T_i - T_o}{R_\text{conv,i} + R_\text{wall} + R_\text{conv,o}}, where R_\text{conv,i} = 1/(h_i A_i), A_i = 2 \pi r_i L, R_\text{conv,o} = 1/(h_o A_o), and A_o = 2 \pi r_o L; here, h_i and h_o are the inner and outer convective coefficients, often derived from correlations for flows. The wall term dominates when k is low or \delta is large relative to the convective resistances. Common pipe materials exhibit a wide range of thermal conductivities that influence the wall's role: pipes have k \approx 45 W/m·K, making conduction resistance minor for typical thicknesses, while pipes offer k \approx 385 W/m·K for even lower resistance in high-conductivity applications. Insulating materials wrapped around pipes, such as , have k \approx 0.04 W/m·K at , significantly increasing R_\text{wall} to enhance thermal isolation.

Overall Heat Transfer Coefficient

The overall heat transfer coefficient, denoted as U, quantifies the combined thermal resistance to heat flow across a multi-layer system, such as the wall of a separating two fluids. It enables the heat transfer rate q to be expressed compactly as q = U A \Delta T_{lm}, where A is the reference surface area and \Delta T_{lm} is the log-mean difference defined for counterflow configurations as \Delta T_{lm} = \frac{\Delta T_1 - \Delta T_2}{\ln(\Delta T_1 / \Delta T_2)}, with \Delta T_1 and \Delta T_2 being the differences at the two ends of the exchanger. This formulation simplifies the analysis of by treating the entire system as an equivalent single resistance. The value of U derives from a one-dimensional thermal resistance network in series, analogous to electrical circuits, where the total resistance determines the . For a clean system without , the reciprocal is \frac{1}{U} = \frac{1}{h_i} + R_{wall} + \frac{1}{h_o}, with h_i and h_o as the inner and outer convective coefficients, and R_{wall} as the conduction through the separating ; fouling resistances can be added separately as additional terms \sum R_f to account for surface deposits. The individual h_i and h_o are obtained from correlations specific to the flow conditions. Since resistances are defined per unit area, U must specify a basis: U_i uses the inner area A_i, while U_o uses the outer area A_o, ensuring the product equality U_i A_i = U_o A_o for consistent heat transfer rates. For a plane wall of thickness L and thermal conductivity k, the expression simplifies to U = \frac{1}{\frac{1}{h_1} + \frac{L}{k} + \frac{1}{h_2}}, assuming uniform properties and heat flow perpendicular to the surfaces. In cylindrical geometries, such as tubes in shell-and-tube heat exchangers, the wall conduction term adjusts for radial variation: for U_i based on inner radius r_i, it becomes R_{wall} = \frac{r_i \ln(r_o / r_i)}{k}, where r_o is the outer radius, yielding \frac{1}{U_i} = \frac{1}{h_i} + \frac{r_i \ln(r_o / r_i)}{k} + \frac{r_i}{r_o h_o}. These adjustments account for the increasing area with radius in conduction through the wall. Shell-and-tube heat exchangers commonly employ this coefficient for , with typical U values ranging from 100 to 1000 W/m²K, varying with fluid types (e.g., higher for water-water systems around 800–1500 W/m²K, lower for organics at 100–300 W/m²K) and operating conditions like and . The derivations assume steady-state operation, where temperatures do not vary with time, and negligible axial conduction, justifying the one-dimensional radial flow approximation.

Thermal Resistance from Fouling

Fouling introduces an additional thermal resistance layer on heat transfer surfaces, primarily in heat exchangers, where deposits accumulate and impede heat flow, effectively reducing the local heat transfer coefficient or the overall heat transfer coefficient U. This resistance, known as the fouling factor R_f, is quantified as R_f = \frac{x}{k_f}, where x is the thickness of the fouling deposit and k_f is its thermal conductivity, typically much lower than that of the base material, leading to a measurable increase in overall thermal impedance. Common types of fouling include , which arises from of inverse solubility salts like in systems; , caused by microbial growth and formation in aqueous environments; and corrosion fouling, resulting from chemical reactions between the fluid and the surface that produce oxide or other deposits. These mechanisms vary by fluid composition, , and , with scaling predominant in cooling water circuits and in untreated open-loop systems. Typical fouling factors for water-based systems range from 0.0001 to 0.001 m²K/W, depending on water quality and conditions; for instance, treated boiler feed water above 325 K exhibits R_f \approx 0.0002 m²K/W, while untreated water may reach 0.00035 m²K/W. These values are compiled in standards like those from the (TEMA), which provide empirical guidelines based on decades of industrial data since 1941. The impact of fouling manifests as a progressive decline in the overall heat transfer coefficient, often reducing U by 20–50% over operational time in exchangers, which necessitates increased energy input or reduced throughput to maintain process temperatures; in one study of a shell-and-tube , U dropped from 1232.6 W/m²K (clean) to 862 W/m²K (fouled), a 43.2% decrease. Monitoring typically involves online estimation of h or U via and flow sensors to detect buildup early. In the overall thermal resistance network, fouling is incorporated as \frac{1}{U} = \frac{1}{h_i} + R_{f,i} + R_{\text{wall}} + R_{f,o} + \frac{1}{h_o}, where R_{f,i} and R_{f,o} denote inner and outer surface fouling resistances, respectively, added to convective and conductive terms; this modification ensures design accounts for anticipated deposit accumulation without altering clean-surface correlations. Mitigation strategies include design oversizing by incorporating conservative R_f values from standards to provide a margin, such as increasing surface area by 10–20% for fouling-prone services, and periodic via methods like hydroblasting or chemical to restore performance. Recent advances in fouling prediction utilize (CFD) models, such as transient simulations with dynamic meshing to forecast deposit growth rates under varying flow conditions, enabling proactive scheduling of in the .

References

  1. [1]
    HEAT TRANSFER COEFFICIENT - Thermopedia
    Heat transfer coefficient is a quantitative characteristic of convective heat transfer between a fluid medium (a fluid) and the surface (wall) flowed over ...
  2. [2]
    Convective Heat Transfer Coefficient | Definition | nuclear-power.com
    The convective heat transfer coefficient, h, can be defined as the rate of heat transfer between a solid surface and a fluid per unit surface area per unit ...
  3. [3]
    Overall Heat Transfer Coefficient: Definitions, Values & Applications
    Heat transfer through a surface like a wall can be calculated as. q = U A dT (1). where. q = heat transfer (W (J/s), Btu/h). U = overall heat transfer ...
  4. [4]
    Convection Heat Transfer - Engineering Library
    The convective heat transfer coefficient (h) is dependent upon the physical properties of the fluid and the physical situation. Typically, the convective heat ...
  5. [5]
    Newton's Law of Cooling: Follow up and exploration - ScienceDirect
    This paper contains the descriptions related to Newton's Law of Cooling. We compared the Newton's temperature measurements and physical phenomena.
  6. [6]
    [PDF] heat in history Isaac Newton and Heat Transfer - LPSM
    dT . In the literature, Newton's law of cooling is synony- mous with the following statement: The rate of cooling of a warm body at any moment is proportional ...
  7. [7]
    Newton's law of cooling and its interpretation - ScienceDirect.com
    Newton's law of cooling suggests that the intensity of energy transfer in the form of heat depends on a difference of temperatures of the interacting physical ...
  8. [8]
    [PDF] PART 3 INTRODUCTION TO ENGINEERING HEAT TRANSFER
    Equation 2.8 is the one-dimensional form of Fourier's law of heat conduction. ... known as the convective heat transfer coefficient. The units of h are W/m.
  9. [9]
    Integrity of Newton's cooling law based on thermal convection theory ...
    Sep 29, 2022 · Here we theoretically show consistent analytical expressions for the constant and convective heat transfer coefficients for various flows.
  10. [10]
  11. [11]
    [PDF] Chapter 8 Internal Flow
    From the definition of the average convection heat transfer coefficient, Equation. 6.9, it follows that. AT. In. AT. = PL hi. Ts. = constant mcp. (8.41a). Page ...
  12. [12]
    [PDF] HEAT AND MASS TRANSFER | Rosen Review
    From Fourier's Law (with the assumption ∇T ≈ ΔT/L),. q = k. L. ΔT ... The equation for convection is. q = ℎ̅ΔT → ΔT = Q̇. 1. ℎ̅A. Then, we can ...
  13. [13]
    18.5 Heat Exchangers - MIT
    It is useful to define an overall heat transfer coefficient $ h_0$ per unit length as. $\displaystyle \dot{Q} =2\pi r_2h_0(T_A-, (18..22). From (18.21) and ...Missing: average | Show results with:average
  14. [14]
    None
    ### Summary of Nusselt Number from the Document
  15. [15]
    None
    ### Summary: Derivation of Nusselt Number from Dimensional Analysis Linked to Boundary Layer Theory
  16. [16]
    What Exactly is the Nusselt Number in Convective Heat Transfer ...
    A larger value of the Nusselt number implies enhanced heat transfer by convection.”
  17. [17]
    Nusselt number to describe convective heat transfer - tec-science
    Apr 26, 2020 · The Nusselt number describes the ratio of convective heat transfer compared to pure heat conduction! At this point it is often tried to give a ...<|separator|>
  18. [18]
    None
    Error: Could not load webpage.<|separator|>
  19. [19]
    III. An experimental investigation of the circumstances which ...
    Cite this article. Reynolds Osborne. 1883III. An experimental investigation of the circumstances which determine whether the motion of water shall be direct ...
  20. [20]
    Transition and Turbulence - Princeton University
    In fluid flow, we often interpret the Reynolds number as the ratio of the inertia force (that is, the force given by mass x acceleration) to the viscous force.
  21. [21]
    [PDF] ME 418 Lecture 6 and 7 – Heat Transfer
    For natural convection, average Nusselt numbers are generally correlated in terms of Grashof or Rayleigh numbers along with Prandtl number. Where. A( is the ...
  22. [22]
    [PDF] Ludwig Prandtl's Boundary Layer
    Prandtl's paper gave the first description of the boundary-layer concept. He theorized that an effect of friction was to cause the fluid immedi- ately adjacent ...<|separator|>
  23. [23]
    Effect of Prandtl Number on Mixed Convective Heat Transfer ... - NIH
    Feb 6, 2020 · Pr is defined as the ratio of momentum diffusivity to thermal diffusivity. When Pr increases but Re is fixed, the thickness of the thermal ...
  24. [24]
    Grashoff/Raleigh Number - Rose-Hulman
    Grashof number can be interpreted as the ratio of buoyancy forces to viscous forces. Thus, low Gr means laminar flow and high Gr means turbulent flow.
  25. [25]
    Closed-Form Analytical Solutions for Laminar Natural Convection on ...
    It is only when the Prandtl number is large (Pr > 1) that the velocity boundary layer is thicker than the thermal boundary layer. Thus, the role of Prandtl ...<|control11|><|separator|>
  26. [26]
    Correlating equations for laminar and turbulent free convection from ...
    The correlation provides a basis for estimating transfer rates for non-Newtonian fluids and for inclined plates.
  27. [27]
    Numerical Analysis of Plume Structures of Fluids on a Horizontal ...
    Direct and indirect natural convection are two subclassifications of natural convection which arises due to the geometry orientation of the heated body. In ...
  28. [28]
    (PDF) Natural convective heat transfer coefficient – a review
    Aug 7, 2025 · This paper presents an extensive review of studies about the natural convective heat transfer coefficient on isolated vertical and horizontal surfaces
  29. [29]
    Theoretical and experimental study of natural convection on a ...
    The authors report a numerical and experimental study of the natural convection flow above an upward facing horizontal heated plate placed in a semi-infinite ...Missing: upper hot
  30. [30]
    Correlating equations for laminar and turbulent free convection from ...
    A simple empirical expression for the mean value of Nu over the cylinder for all Ra and all Pr is developed in terms of the model of Churchill and Usagi.
  31. [31]
    (PDF) Free Convection Heat Transfer from Horizontal Cylinders
    Oct 16, 2025 · A numerical study has been conducted for natural convection heat transfer for air from two vertically separated horizontal heated cylinders ...
  32. [32]
    [PDF] Natural convection heat transfer around a horizontal circular cylinder ...
    Natural convection heat transfer is an inexpensive, reliable and noise-free method of heat transfer. For instance, horizontal cylinder has many engineering.
  33. [33]
    [PDF] ICONE-9768 NATURAL CONVECTION IN ENCLOSURES ...
    Laminar natural convection in rectangular enclosures with different aspect ratios, containing lead-bismuth and lead is studied numerically in this paper. The ...Missing: k_eff 0.22
  34. [34]
    Nusselt Number - an overview | ScienceDirect Topics
    Nusselt number is a well-known dimensionless parameter that is a measure for the ratio of pure convection to pure conduction heat transfer at the solid surface.
  35. [35]
    [PDF] Heat Transfer in Circular ducts Fully Developed Flow-I Fully ...
    Nussselt number are constant. • In laminar flow, the values of Nu = 4.36 for constant wall heat flux case and is 3.66 for constant wall temperature case.
  36. [36]
    [PDF] New Nusselt number correlations for developing and fully ...
    Two correlations for local Nusselt number of laminar flow in ducts with arbitrary cross-section were developed for constant wall temperature and constant wall ...
  37. [37]
    Dittus-Boelter Equation | Correlation & Calculation - Nuclear Power
    To calculate the Nusselt number, we have to know: the Reynolds number, which is ReDh = 575600. the Prandtl number, which is Pr = 0.89.
  38. [38]
    [PDF] Simple heat transfer correlations for turbulent tube flow
    The Nusselt number determined experimentally was 10 to 20% higher than values predicted by the Dittus-Boelter equation. The discrepancy between the results of.
  39. [39]
    Validation of the Gnielinski correlation for evaluation of heat transfer ...
    Gnielinski correlation was modified using 103 data points and a new correlation was obtained, which showed a total mean error of 2.1%, and an agreement within ± ...
  40. [40]
    GRAETZ NUMBER - Thermopedia
    The Graetz number, Gz, is a nondimensional group applicable mainly to transient heat conduction in laminar pipe flow.
  41. [41]
    Graetz Number - an overview | ScienceDirect Topics
    The Graetz number is a dimensionless quantity that characterizes heat transfer in laminar flow through a closed conduit, reflecting the ratio of thermal energy ...
  42. [42]
    Heat transfer correlation for circular and non-circular ducts in the ...
    A new model to predict the Nusselt number in the entry region of smooth non-circular ducts and channels is developed, for both uniform heat flux and uniform ...
  43. [43]
    Hydraulic Diameter - an overview | ScienceDirect Topics
    Hydraulic diameter is four times the flow area divided by the duct perimeter, or four times the cross-section area divided by the circumference.
  44. [44]
    McAdams - Thom - Chen's Correlation - Nucleate Boiling
    Chen proposed a correlation where the heat transfer coefficient is the sum of a forced convection component and a nucleate boiling component. It must be noted ...
  45. [45]
    Radiation Heat Transfer Coefficient - an overview - ScienceDirect.com
    The radiation heat transfer coefficient, denoted as h_r, is defined as a parameter that expresses radiation heat transfer in terms of a temperature ...Missing: linearization | Show results with:linearization
  46. [46]
    7.0 Thermal Control - NASA
    Feb 5, 2025 · In a vacuum, heat is transferred only by radiation and conduction, with no convection. The internal environment of a fully enclosed small ...State-of-the-Art – Passive... · Sprayable Thermal Control... · Heat Pipes
  47. [47]
    Tech Brief: Radiation Basics: When Does It Matter?
    Sep 30, 2024 · Radiation is the primary means of heat transfer in vacuum environments and must be considered.
  48. [48]
    [PDF] Basic Heat Transfer and Some Applications in Polymer Processing
    k. = Q w δ. The quantity k/δ = h is called the heat transfer coefficient and it is an extremely important concept in heat transfer. This simplified model is ...
  49. [49]
    An updated review of nanofluids in various heat transfer devices
    Jun 15, 2020 · This review takes a retrospective look at the year 2019 by reviewing the progress made in the area of nanofluids preparation and the applications of nanofluids.
  50. [50]
    A comprehensive review on microchannel heat sinks for electronics ...
    Jan 11, 2024 · This paper comprehensively takes an overview of the research progress in MCHSs and generalizes the hotspots and bottlenecks of this area.
  51. [51]
    17.2 Combined Conduction and Convection - MIT
    Equation (17.21) is the thermal resistance for a solid wall with convection heat transfer on each side. For a turbine blade in a gas turbine engine, cooling ...
  52. [52]
  53. [53]
    [PDF] STEADY HEAT CONDUCTION
    In this case, the thermal resistance corresponds to electrical resistance, temperature difference corresponds to voltage, and the heat transfer rate corresponds ...<|control11|><|separator|>
  54. [54]
    Thermal Conductivity of Steel: Carbon, Alloys and Stainless Steels
    May 4, 2021 · Carbon steels have the highest thermal conductivity, averaging 45 Watts per Kelvin per meter. Heat treatment can harden steels containing over ...
  55. [55]
    Thermal Conductivity
    Their calculation for freon filled polyurethane of density 1.99 lb/ft3 at 20°C gives a thermal conductivity of 0.022 W/mK. The calculation for CO2 filled ...Missing: pipes | Show results with:pipes
  56. [56]
    Fiberglass Insulation - The Engineering ToolBox
    Thermal conductivity of fiberglass insulation - temperature and k-values. · T (oC) = 5/9 (T (oF) - 32) · 1 W/(m K) = 6.935 (Btu in)/(h ft² °F) ...
  57. [57]
    None
    ### Summary of Key Sections from the Document
  58. [58]
    Fouling Factor - an overview | ScienceDirect Topics
    Traditional heat exchanger design methodologies largely rely on highly empirical, fixed, TEMA “fouling factors” (TEMA, 1941). These are basically “safety ...
  59. [59]
    Heat Exchangers - Fouling and Reduced Heat Transfer
    Heat-transfer in heat exchangers are reduced by fouling. During operation with liquids and gases a dirt film may build up on the heat exchanger surfaces.
  60. [60]
    4 Kinds of Tube-Side Coil Fouling & How to Prevent Them
    Dec 9, 2020 · Corrosion fouling, scaling, biofouling and fouling from particulates can reduce your equipment's tube-side heat transfer.
  61. [61]
    Designing Heat Exchangers to Mitigate Fouling - AIChE
    Failing to account for fouling resistance in a heat exchanger may lead to reduced heat transfer efficiency, increased energy consumption, higher pressure drops ...<|separator|>
  62. [62]
    [PDF] The Impact of Fouling on Heat Transfer, Pressure Drop, and ... - ijrpr
    In one study, it was found that the overall heat transfer coefficient in a clean state is 1232.6, while in a fouled state, it drops to 862, indicating a 43.2% ...
  63. [63]
    (PDF) A new approach to predict fouling using a transient CFD ...
    Aug 7, 2025 · In this work, a novel approach to predict the fouling layer growth on the internal surface of a tube by means of a transient computational ...