Fact-checked by Grok 2 weeks ago

Viscosity

Viscosity is a measure of a 's to , describing the internal that opposes the relative motion of its layers. This property arises from interactions between molecules, transforming of bulk motion into through dissipative processes akin to . In , viscosity governs behaviors such as the spreading of liquids, the through pipes, and the formation of layers in , making it a fundamental parameter in . Viscosity is quantified in two primary forms: dynamic viscosity (also called absolute viscosity) and kinematic viscosity. Dynamic viscosity, denoted by the symbol μ, represents the fluid's resistance to under an applied force and is defined by Newton's law of viscosity as the ratio of (τ) to the velocity gradient (du/dy) in : τ = μ (du/dy). Kinematic viscosity, denoted by ν, is the dynamic viscosity divided by the fluid's (ρ), or ν = μ / ρ, and is commonly used to describe flows influenced by body forces such as . The SI unit for dynamic viscosity is the pascal-second (Pa·s), equivalent to one newton-second per square meter (N·s/m²), while kinematic viscosity is measured in square meters per second (m²/s). Viscosity varies significantly with temperature—for liquids, it decreases as temperature rises due to reduced molecular cohesion, whereas in gases, it increases with temperature because of higher molecular speeds. These dependencies are critical in applications ranging from lubrication in engines to atmospheric modeling. In and physics, viscosity influences phenomena like drag forces, transitions via the (Re = ρVL / μ), and material processing, where high-viscosity fluids like exhibit slow deformation compared to low-viscosity ones like . Understanding viscosity enables predictions of fluid behavior in diverse fields, including , , and .

Etymology and Definitions

Etymology

The word "viscosity" derives from the viscositas, meaning "stickiness," which stems from viscosus ("sticky") and ultimately from the Latin viscum (""). This etymological root refers to the viscous, produced from the berries of the mistletoe plant, used historically to trap birds, and the term was later extended metaphorically to characterize the clinging or resistant quality of semi-fluid substances. Early conceptual notions of viscosity trace back to ancient , where qualitative descriptions of fluid resistance appeared without the modern term. For instance, , in his Physics (circa 350 BCE), observed that objects moving through denser media like experience greater opposition to motion than in air, attributing this to the medium's inherent resistance that scales with its "thickness." The English term "viscosity" itself emerged in the late , initially denoting a general state of viscidity or glutinousness, and entered scientific usage in the amid investigations into fluid behavior. Isaac Newton's (1687) formalized proportional relationships in fluid resistance, influencing later terminology without explicitly defining viscosity as a distinct . In the , the term gained precise application in when incorporated viscous effects into his 1822 generalization of Leonhard Euler's inviscid equations, marking a pivotal formalization of viscosity as internal friction, though the word had long predated this in broader natural philosophical contexts.

Dynamic Viscosity

Dynamic viscosity, denoted by the symbol μ, is a fundamental property of that quantifies their resistance to flow under applied forces. It is defined as the ratio of the τ to the , which is the velocity gradient du/dy perpendicular to the flow direction. This relationship is expressed mathematically as \mu = \frac{\tau}{\frac{du}{dy}} where τ represents the shear stress, the internal frictional force per unit area that opposes the relative motion of adjacent fluid layers sliding past one another. In simple shear flow, such as between two parallel plates where one moves relative to the other, this gradient du/dy arises from the velocity difference across the fluid layers, and dynamic viscosity characterizes the linear proportionality between this stress and rate for Newtonian fluids. Unlike viscosity, which pertains to a fluid's resistance to uniform or , dynamic viscosity specifically addresses deformation due to without volume change, making it central to understanding frictional effects in laminar flows. This distinction ensures that dynamic viscosity focuses on tangential stresses in directional flows, excluding dilatational contributions. The concept originates from Newton's 1687 work , where he postulated that for ideal fluids, the resistive force is proportional to and independent of the , assuming constancy of at a given —a foundational assumption for Newtonian behavior. This linearity underpins the definition, distinguishing it from more complex fluid responses.

Kinematic Viscosity

Kinematic viscosity, denoted by the symbol \nu, is defined as the ratio of a fluid's dynamic viscosity \mu to its \rho, expressed as \nu = \frac{\mu}{\rho}. This measure has units of area per unit time, such as square meters per second in the SI system. Physically, kinematic viscosity represents the of within the , quantifying how readily is transported through the medium due to viscous effects. It characterizes the 's resistance to shear relative to its inertial properties, making it particularly useful in analyses where variations influence flow behavior. A key application of kinematic viscosity is in the , Re = \frac{v L}{\nu}, where v is the and L is a ; this predicts whether a will be laminar or turbulent by comparing inertial to viscous forces. The value of kinematic viscosity varies with and through the dependencies of both \mu and \rho; for instance, in ideal gases, \nu scales proportionally to T^{3/2}/p, where T is and p is , since dynamic viscosity \mu increases with while being largely independent of , and \rho decreases with rising or falling . In liquids, typically reduces \mu more significantly, leading to a decrease in \nu, whereas effects on \rho are more pronounced but often secondary.

Fluid Behavior

Newtonian Fluids

A Newtonian fluid is defined by the linear relationship between shear stress \tau and the rate of strain \dot{\gamma}, expressed as \tau = \eta \dot{\gamma}, where \eta is the dynamic viscosity that remains constant and independent of the shear rate \dot{\gamma}. This constitutive relation, known as Newton's law of viscosity, implies that the fluid's resistance to flow does not vary with the intensity of deformation, distinguishing it from more complex behaviors in other fluids. Common examples of Newtonian fluids include water, air, glycerine, and most simple gases and low-molecular-weight liquids under typical conditions of low shear. These fluids exhibit constant viscosity across a wide range of shear rates, making their flow behavior predictable in engineering applications such as pipelines and lubrication systems. The constant viscosity of Newtonian fluids has key implications for flow dynamics, particularly in enabling at low Reynolds numbers, where layers maintain smooth, parallel motion without mixing or . For instance, the drag force F_d on a small of r moving at low v through such a is given by : F_d = 6 \pi \eta r v, which quantifies viscous resistance in and colloidal systems. This linearity arises from fundamental physical mechanisms. In dilute gases, kinetic theory explains viscosity as the net momentum transfer between adjacent layers by molecules with average thermal speed \bar{v} traversing the mean free path \lambda, yielding \eta \approx \frac{1}{3} \rho \lambda \bar{v}, where \rho is density; the proportionality to the velocity gradient ensures \eta remains independent of \dot{\gamma}. For low-molecular-weight liquids, short-range intermolecular forces dominate, providing a consistent frictional resistance without rate-dependent alignment or entanglement effects.

Non-Newtonian Fluids

Non-Newtonian fluids are substances in which the viscosity does not remain constant but varies with the applied or over time, deviating from the linear relationship between and shear rate observed in Newtonian fluids. This behavior arises in complex fluids containing particles, polymers, or other microstructures that respond to deformation. Non-Newtonian fluids are broadly classified into time-independent and time-dependent categories based on how their flow resistance changes. In time-independent non-Newtonian fluids, viscosity depends solely on the instantaneous . Shear-thinning fluids, also known as pseudoplastic, exhibit decreasing viscosity as increases, allowing easier flow under ; common examples include paints and solutions. Conversely, shear-thickening or fluids show increasing viscosity with higher s, often due to or particle interactions; a classic example is a cornstarch-water , which hardens under rapid impact. Bingham plastics represent yield- fluids that behave as solids below a critical but flow like viscous liquids above it, exemplified by , which holds its shape until squeezed. Time-dependent non-Newtonian fluids, such as thixotropic ones, display viscosity that decreases over time under constant shear due to reversible structural breakdown, with recovery upon rest; this is seen in certain inks and gels. A key model for describing shear-thinning and shear-thickening behaviors is the Ostwald-de Waele power-law model, which relates \tau to \dot{\gamma} as: \tau = K \dot{\gamma}^n where K is the consistency index and n is the flow behavior index; n < 1 indicates shear-thinning, while n > 1 denotes shear-thickening. The apparent viscosity \eta_\text{app} is then defined as \eta_\text{app} = \tau / \dot{\gamma} = K \dot{\gamma}^{n-1}, which varies with , providing a measure of effective flow resistance. Everyday relevance of non-Newtonian fluids is evident in biological and industrial contexts. , for instance, acts as a shear-thinning , with viscosity dropping at higher shear rates in vessels to facilitate circulation. Polymer melts and solutions in also exhibit pseudoplastic , enabling efficient processing in and applications. These properties influence , from non-drip paints to protective body armors using materials.

Physical Mechanisms

Momentum Transport

Viscosity manifests as the transport of across adjacent layers that exhibit relative motion, arising from intermolecular forces that enable exchange perpendicular to the primary . In a shearing , molecules from a faster-moving layer collide with those in a slower layer, effectively diffusing downward and equalizing velocities over time. This process is analogous to phenomena, where viscosity acts as the for , smoothing out velocity gradients and resisting shear. A classic illustration occurs in through a , such as Poiseuille flow between parallel plates, where the transport due to viscosity results in a parabolic profile. The maximum is at the center, decreasing symmetrically toward the walls due to the , with the profile governed by the balance between pressure-driven and viscous of . This parabolic shape emerges directly from solving the simplified Navier-Stokes equations under steady, incompressible conditions, highlighting how viscous transfer enforces the velocity variation. The quantitative relation for this in simple shear is captured by Newton's law of viscosity, expressed as the \tau = -\eta \frac{du}{dy}, where \eta is the dynamic viscosity and \frac{du}{dy} is the velocity gradient normal to the flow. This form parallels Fick's first law for mass diffusion (J = -D \frac{dc}{dy}) and Fourier's law for heat conduction (q = -k \frac{dT}{dy}), underscoring the unified framework of processes in continua, with \eta serving as the diffusivity . Within the full Navier-Stokes equations, viscous effects enter through the of the stress tensor, specifically the term \nabla \cdot (\eta \nabla \mathbf{v}) (for constant \eta), which represents the net diffusive flux of momentum and acts to dampen velocity fluctuations. This term is essential for describing the evolution of velocity fields in viscous flows. In practical applications, such momentum transport dominates in boundary layers adjacent to solid surfaces, where it generates by slowing fluid near the wall and creating low-momentum regions. Strategies for drag reduction, such as injecting low-viscosity fluids or using additives, target these layers to enhance momentum transfer away from the wall, thereby reducing and overall resistance.

Molecular Origins in Gases

In dilute gases, viscosity originates from the diffusive transport of between adjacent layers through intermolecular collisions, as described by kinetic theory. Molecules traveling across velocity gradients carry excess from faster-moving layers to slower ones, resulting in a net that opposes the flow. This microscopic mechanism underpins the macroscopic viscous resistance observed in gases. The foundational estimate from kinetic theory for the shear viscosity \eta of a dilute gas is given by \eta \approx \frac{1}{3} \rho \lambda v_{\rm avg}, where \rho is the mass density, \lambda is the between collisions, and v_{\rm avg} is the average molecular speed. This expression arises from considering the flux of molecules across a , each transporting on the order of m v_z v_x (with m the and v_x, v_z components), averaged over the Maxwell-Boltzmann distribution and assuming a velocity gradient over the scale of \lambda. James Clerk Maxwell derived this form in his pioneering work on gas , confirming its proportionality to molecular speed and path length. A key feature of this model is the of \eta from gas (or ) at constant , valid for dilute conditions where the is much larger than molecular sizes. The \lambda \approx 1/(\sqrt{2} \pi d^2 n) inversely scales with n (and thus \rho = m n), so the product \rho \lambda remains constant, yielding \eta independent of n. This counterintuitive result, predicted by , was experimentally verified and distinguishes gaseous viscosity from that in denser fluids. With , v_{\rm avg} \propto \sqrt{T} from equipartition, while \lambda is temperature-independent in the basic model, leading to \eta \propto T^{1/2}. The simple hard-sphere model, assuming molecules as rigid spheres of diameter d with no long-range forces, provides a good first approximation for monatomic gases like or at moderate conditions but has limitations. It underpredicts the temperature dependence (T^{1/2}) compared to experiments, where real gases show a stronger increase due to attractive intermolecular potentials that effectively reduce the collision cross-section at higher temperatures. Corrections for real gases, such as the model, modify the effective collision cross-section \sigma \propto d^2 (1 + S/T), where S is a characteristic temperature reflecting attractive forces, yielding \eta \propto T^{1/2} / (1 + S/T). This semi-empirical adjustment, originally proposed by William , improves accuracy for polyatomic gases like air over wide temperature ranges without resorting to full quantum treatments. For bulk viscosity \zeta, which arises in compressible flows involving volume dilation, kinetic theory predicts \zeta \approx 0 for dilute monatomic gases. Without internal (e.g., or ), there is no relaxation time for energy redistribution during or , so no additional dissipative beyond effects. This result holds in the hard-sphere and is confirmed by the Chapman-Enskog solution for low-density monatomic gases.

Molecular Origins in Liquids

In liquids, viscosity arises primarily from the strong intermolecular interactions and the motion required for molecules to flow past one another in a dense medium, contrasting with the dilute collision-dominated transport in gases. Unlike gases, where viscosity stems from via infrequent binary collisions, liquids exhibit viscosity due to the caged of molecules surrounded by neighbors, leading to higher resistance to as increases. This dense-phase behavior results in activation barriers for flow, where molecules must overcome hurdles to rearrange, often modeled through extensions of kinetic theory adapted for correlated motions. The Enskog theory, originally developed for dense gases, has been extended to s by incorporating corrections for frequent collisions and spatial correlations in high-density regimes, predicting that viscosity η increases with due to enhanced collision rates and reduced mean free paths. In these modifications, such as the modified Enskog theory (MET), the viscosity is expressed as η = η_0 * Y, where η_0 is the low-density limit from Chapman-Enskog theory, and Y is a density-dependent Enskog correction factor accounting for pair correlations that amplify momentum transfer in crowded environments. This extension successfully describes how higher liquid densities elevate viscosity by promoting more entangled molecular trajectories, as validated for simple fluids like near saturation. Temperature dependence in liquids follows an Arrhenius-like form, η = A exp(E_a / RT), where A is a pre-exponential factor, E_a is the activation energy for viscous flow reflecting the energy barrier to molecular rearrangement, R is the gas constant, and T is absolute temperature; this exponential increase in viscosity with decreasing temperature underscores the role of thermal energy in overcoming intermolecular attractions. For many organic liquids, E_a correlates with molecular size and polarity, typically ranging from 10-30 kJ/mol, as derived from empirical fits to experimental data. This model holds well above the glass transition but deviates at lower temperatures where cooperative effects dominate. Intermolecular forces significantly influence liquid viscosity, with hydrogen bonding in creating a dynamic network that enhances resistance to flow by forming transient bridges between molecules, leading to 's anomalously high viscosity compared to non-hydrogen-bonding liquids of similar mass. In contrast, van der Waals forces dominate in nonpolar oils, where interactions between chains increase viscosity proportional to chain length and branching, as longer chains foster greater entanglement and slower relaxation times. These forces contribute to the scale of viscosity, with hydrogen-bonded systems like showing E_a ≈ 16 kJ/mol, while van der Waals-dominated oils like exhibit higher values around 25 kJ/mol. Free volume theory provides a complementary perspective, positing that viscosity rises dramatically as available free volume—the unoccupied per molecule—decreases near the T_g, where molecular freezes due to insufficient for diffusive jumps. In this , the diffusion coefficient D ∝ exp(-B / v_f), with v_f the fractional free volume and B a constant, linking viscosity inversely to D via the Stokes-Einstein relation; as drops below T_g + 50 K, v_f shrinks, causing η to span orders of magnitude from 10^2 ·s at T_g to 10^{12} ·s in the glassy state. This theory, pioneered by and Turnbull, explains the universal super-Arrhenius behavior in dense liquids approaching .

Extensions to Other Systems

Bulk Viscosity

Bulk viscosity, denoted as \zeta, represents the fluid's resistance to uniform volumetric compression or expansion, in contrast to dynamic viscosity which governs resistance to deformation. In the context of compressible , it manifests as a deviation in the mechanical from its value, related to the of the through the constitutive P' = -\zeta (\nabla \cdot \mathbf{v}), where P' is the deviation and \mathbf{v} is the . This term arises in the Navier-Stokes equations for the , capturing dissipative effects during isotropic volume changes. In simple fluids such as monatomic gases in the dilute limit, bulk viscosity is negligible, approximately \zeta \approx 0, because these systems lack internal degrees of freedom or significant intermolecular forces that could lead to delayed equilibration during compression or expansion. However, in polyatomic gases and molecular liquids, bulk viscosity becomes substantial due to relaxation processes involving internal modes, such as rotational and vibrational excitations, which cannot instantaneously adjust to rapid volume changes, resulting in non-equilibrium pressure contributions. For instance, in polyatomic species like nitrogen, these mechanisms introduce a finite \zeta that scales with the complexity of molecular structure. Bulk viscosity plays a key role in acoustic propagation and shock wave dynamics. It contributes to sound absorption, where the classical attenuation coefficient includes a term proportional to \zeta, specifically \alpha = \frac{\omega^2}{2\rho c^3} \left( \frac{4}{3} \eta + \zeta + (\gamma - 1)^2 \frac{\kappa}{\rho c_p} \right), with \eta as dynamic viscosity, \kappa thermal conductivity, \rho density, c sound speed, \gamma the adiabatic index, and \omega angular frequency; the \zeta term accounts for structural relaxation damping. In shock waves, bulk viscosity influences the transition zone thickness, with higher \zeta leading to broader profiles; theoretical models show a linear dependence of the normalized shock thickness on the ratio \zeta / \eta, as observed in experiments with polyatomic gases like sulfur hexafluoride. Experimental determination of bulk viscosity primarily relies on attenuation measurements, which isolate the excess not explained by viscosity or . Acoustic techniques have been applied to various Newtonian fluids, including gases and liquids, revealing \zeta values independent of properties like or viscosity; for example, in liquid and organic solvents, attenuation data yield \zeta on the of 10^{-3} to 10^{-2} ·s across frequencies from 1 to 100 MHz. Similar experiments in polyatomic gases, such as from 77 to 300 , confirm the role of vibrational relaxation in generating measurable \zeta, with spectra fitting models that attribute up to 50% of total damping to bulk effects.

Viscosity in Solids

In solids, viscosity manifests as resistance to deformation rates under sustained , analogous to fluid shear viscosity but applied to the behavior of viscoelastic materials. In tests, where a σ is applied, the resulting strain ε(t) increases over time, and the creep compliance is defined as J(t) = ε(t) / σ, quantifying the material's time-dependent deformability. The Maxwell model, a fundamental viscoelastic framework consisting of a (elastic modulus E) in series with a (viscosity η), captures this by relating the viscous component to the : the alone yields dε/dt = σ / η, so η = σ / (dε/dt), where the total combines elastic and viscous contributions. This model highlights how solids can exhibit fluid-like flow under prolonged loading, distinguishing viscosity from pure elasticity, which involves instantaneous, recoverable deformation without time dependence. Unlike ideal elastic solids, many real solids possess a yield stress beyond which permanent deformation occurs, yet they demonstrate viscous flow characteristics particularly at elevated temperatures or over extended timescales. For instance, glacier ice deforms viscously under its own weight, flowing like a with an effective viscosity on the order of 10^{13} to 10^{14} Pa·s, enabling slow movement over geological periods without fracturing. Similarly, , a highly viscous at with η ≈ 10^{8} Pa·s, flows imperceptibly slowly, as demonstrated by the ongoing where drops form roughly every decade, illustrating solid-like rigidity masking underlying viscous behavior. Amorphous solids, such as glassy polymers, can be viewed as supercooled liquids frozen into a non-equilibrium state, where viscous dominates near the temperature T_g. In glassy polymers, viscosity reaches extraordinarily high values near T_g, typically around 10^{12} Pa·s, marking the boundary where structural relaxation times approach observable scales (e.g., 100 seconds), transitioning the material from a viscous liquid to a rigid glass. This regime is critical for applications like polymer processing, where flow resistance governs shaping and annealing. Recent advances in 2025 have introduced atomistic models that connect near-glass-transition viscosity directly to the full spectrum of atomic vibration modes, using non-affine lattice dynamics to compute shear viscosity from low-frequency vibrational contributions without relying on computationally intensive simulations. These models, validated on polymer melts like the Kremer-Grest system, reveal how collective vibrational anharmonicity enhances flow resistance, providing predictive power across temperatures where traditional methods falter.

Eddy Viscosity

Eddy viscosity, denoted as \nu_t, is an empirical concept in turbulence modeling that parameterizes the enhanced momentum transport due to turbulent eddies, analogous to molecular viscosity in laminar flows. It is defined by the relation \nu_t = \tau_t / [\rho (\partial u / \partial y)], where \tau_t represents the turbulent shear stress, \rho is the fluid density, and \partial u / \partial y is the mean velocity gradient in the direction perpendicular to the flow. This formulation arises from the Boussinesq hypothesis, which assumes that turbulent fluctuations act like an additional viscous stress on the mean flow. A foundational approach to estimating eddy viscosity is Prandtl's mixing-length , introduced in the early . According to this , \nu_t \approx l^2 |\partial u / \partial y|, where l is the mixing length, a characteristic scale representing the average distance traveled by turbulent eddies before their momentum is redistributed. Near walls, l is often taken as proportional to the distance from the surface, such as l = \kappa y with \kappa \approx 0.41. This simple model effectively captures the in flows dominated by a single length scale. In practice, eddy viscosity plays a central role in closing the Reynolds-Averaged Navier-Stokes (RANS) equations for simulating turbulent flows in engineering applications, such as fully developed pipe flows and atmospheric boundary layers. By incorporating \nu_t into the effective viscosity, RANS models approximate the Reynolds stresses as \tau_{ij} = \rho \nu_t (\partial U_i / \partial x_j + \partial U_j / \partial x_i) - (2/3) \rho k \delta_{ij}, enabling computationally efficient predictions of mean flow fields without resolving individual eddies. These models are particularly valuable for design in , , and , where high-fidelity direct simulations are infeasible. However, eddy viscosity is not a true thermophysical property of the fluid like molecular viscosity; instead, it varies spatially and temporally with local flow conditions, turbulence intensity, and geometry, often requiring ad hoc tuning or additional transport equations for its prediction. This dependence leads to limitations in non-equilibrium flows, such as those with strong streamline curvature, separation, or rapid distortions, where the isotropic assumption fails and models can produce unphysical results like excessive diffusion or instabilities.

Measurement and Units

Measurement Techniques

Viscosity measurement techniques vary depending on the type, viscosity range, and whether the behaves as Newtonian or non-Newtonian, with methods designed to apply controlled and quantify resistance to . Common approaches include , rotational, falling sphere, and oscillatory methods, each leveraging fundamental principles to derive viscosity from measurable parameters like , , or . These techniques are calibrated against standard to ensure accuracy, often achieving precisions better than 1% for low-viscosity liquids. Capillary viscometers are widely used for low-viscosity Newtonian fluids, such as gases and oils, by forcing the through a narrow and measuring the under a difference. The method relies on the Hagen-Poiseuille , which assumes laminar, fully developed flow in a cylindrical : \Delta P = \frac{8 \eta L Q}{\pi r^4}, where \Delta P is the , \eta is the dynamic viscosity, L is the capillary length, Q is the , and r is the ; solving for \eta yields viscosity directly from experimental measurements of \Delta P and Q. This technique is effective for viscosities ranging from 0.1 to 100 ·s, with automated versions enabling high-throughput analysis in pharmaceutical applications. Rotational viscometers, particularly those employing Couette geometry with coaxial s, measure viscosity by rotating an inner within a stationary outer one filled with the fluid and quantifying the resulting . For Newtonian fluids in this setup, the T balances the viscous , given by T = \frac{4 \pi \eta \Omega h r_i^2 r_o^2}{r_o^2 - r_i^2}, where \Omega is the , h is the (or ), and r_i, r_o are the inner and outer radii, respectively; viscosity is then computed from measured and speed. This method suits moderate to high viscosities (up to 10^6 mPa·s) and can handle opaque samples, making it versatile for industrial fluids like paints and polymers./20%3A_Miscellaneous/20.04%3A_Viscosity/20.4.02%3A_The_Couette_Viscometer) The falling sphere viscometer determines viscosity by observing the terminal velocity of a descending through a transparent column under gravity, applicable to low-viscosity Newtonian liquids like or oils. Based on for the drag F_d = 6 \pi \eta r v balancing the buoyant weight, the terminal v satisfies \eta = \frac{2 r^2 g (\rho_s - \rho_f)}{9 v}, where r is the radius, g is , and \rho_s, \rho_f are the densities of the and ; high-speed or timing tracks v for precise . This is simple and absolute, offering accuracies of 0.5-2% for viscosities below 100 mPa·s, though wall effects require corrections for finite tube diameters. Oscillatory rheometers extend viscosity assessment to non-Newtonian and high-viscosity materials, such as gels and viscoelastic polymers, by applying sinusoidal and measuring the stress response to derive dynamic moduli. The loss modulus G'' relates to the viscous component via the dynamic viscosity \eta' = G'' / \omega, where \omega is the , allowing of -rate-dependent without excessive deformation. These instruments, often using parallel-plate or cone-plate geometries, operate in the linear viscoelastic (strain < 1%) and are essential for complex fluids where steady may induce structural changes. Recent advances in microfluidic devices have enabled viscosity measurements on microliter-scale samples, particularly for biological fluids and high-value materials, overcoming limitations of traditional methods in sample volume and portability. For instance, chip-based viscometers integrate pressure-driven in microchannels to apply Poiseuille-like principles, achieving rapid (seconds) assessments with 10 μL volumes and accuracies of 5% for protein solutions up to 10 ·s. These post-2020 innovations, including temperature-controlled rheometers, facilitate monitoring of non-Newtonian effects like in complex mixtures.

Units and Dimensions

The standard unit for dynamic viscosity in the () is the pascal-second (Pa·s), which is equivalent to the per square meter (N·s/m²). In the centimeter-gram-second (CGS) system, the corresponding unit is the poise (P), defined as the -second per square centimeter (dyne·s/cm²). The conversion between these units is given by 1 P = 0.1 Pa·s. Kinematic viscosity, defined as the ratio of dynamic viscosity to fluid , has the SI unit of square meters per second (m²/s). In the CGS system, it is measured in stokes (), where 1 St = 10^{-4} m²/s. Dimensional analysis yields the expression for the dimension of dynamic viscosity as [\eta] = M L^{-1} T^{-1}, where M represents , L length, and T time. Historical units for viscosity, particularly in the petroleum industry, include Saybolt Universal Seconds (SUS), which quantify the time in seconds for a fixed volume of fluid to flow through a standardized orifice and are primarily applied to oils.

Theoretical Prediction

Chapman–Enskog Theory for Gases

The Chapman–Enskog expansion provides a systematic perturbation method to solve the Boltzmann equation for dilute gases, deriving expressions for transport coefficients such as viscosity directly from the intermolecular potential function. Developed independently by Sydney Chapman in 1916–1917 and David Enskog in 1917, and later refined in their collaborative work, the theory expands the velocity distribution function in powers of the Knudsen number (the ratio of mean free path to macroscopic length scale), assuming small gradients in velocity, temperature, and density. To first order, this yields the Navier–Stokes constitutive relations, with viscosity expressed as a function of molecular mass, temperature, and collision parameters derived from the potential. The first-order approximation for the shear viscosity η of a is given by \eta = \frac{5}{16 \sigma^2} \sqrt{\frac{\pi m k_B T}{\Omega^{(2,2)}}}, where m is the , k_B is Boltzmann's , T is the , σ is the characteristic collision from the intermolecular potential (e.g., Lennard-Jones), and Ω^{(2,2)} is the collision integral for viscosity, which depends on the reduced T^* = k_B T / ε (with ε the potential well depth). This formula arises from integrating the linearized Boltzmann collision operator over the perturbation to the Maxwellian distribution. For simple potentials like (where Ω^{(2,2)} = 1), the viscosity simplifies to η = (5/16 σ²) √(π m k_B T), independent of in the dilute . The temperature dependence of viscosity follows η ∝ T^s, where the exponent s varies with the intermolecular potential: s = 0.5 for , approaching s = 1 for long-range inverse-power potentials, and typically s ≈ 0.6–0.8 for realistic Lennard-Jones potentials used in . This arises because the collision integral Ω^{(2,2)} decreases with increasing T^*, softening the effective repulsion at higher temperatures. The theory has been validated extensively for like , , , and , where predictions using Lennard-Jones parameters match experimental viscosities to within 1–2% over wide ranges at low densities. For example, for near , the computed η ≈ 22.7 μPa·s aligns closely with measured values, confirming the accuracy of the first-order expansion for monatomic systems. Extensions to polyatomic gases incorporate internal via the Wang Chang–Uhlenbeck formalism, which modifies the to include rotational and vibrational energy distributions, enabling predictions of both and viscosities while retaining the Chapman–Enskog structure. This approach has been applied successfully to diatomic gases like and oxygen, adjusting collision integrals for anisotropic potentials.

Models for Liquids and Mixtures

For pure liquids, the Eyring theory provides a fundamental activated-rate approach to viscosity, treating flow as a thermally activated process where molecules overcome an energy barrier to shear. The theory posits that the shear viscosity \eta is given by \eta = \frac{h N_A}{V} e^{\Delta G / RT}, where h is Planck's constant, N_A is Avogadro's number, V is the , \Delta G is the of activation, R is the , and T is the . This model, derived from absolute reaction rate theory, successfully correlates viscosity with temperature dependence in simple liquids like and hydrocarbons, emphasizing the role of molecular rearrangements in dense fluids. Building on Eyring's framework, the significant structure theory of liquids models the liquid state as a quasi-lattice with a fraction of gas-like and solid-like , enabling predictions of transport properties including viscosity. In this approach, viscosity arises from the balance between vibrational (solid-like) and translational (gas-like) contributions, with the theory expressing \eta through partition functions that partition the liquid's configurational space. Developed in the , it has been applied to compute viscosities of metals and liquids, offering insights into how structural influences resistance beyond simple models. For mixtures, particularly liquid blends, the Arrhenius mixing rule approximates the logarithm of the mixture viscosity as a weighted sum of the pure-component logarithms: \ln \eta_{\text{mix}} = \sum x_i \ln \eta_i, where x_i are fractions and \eta_i are pure viscosities. This empirical relation, rooted in reaction rate theory, performs well for miscible non-polar liquids like blends at moderate concentrations, capturing the exponential temperature sensitivity of flow. In contrast, for gas mixtures, the Wilke equation extends kinetic theory by weighting pure-component viscosities with collision factors: \eta_{\text{mix}} = \sum y_i \frac{\eta_i}{\sum y_j \phi_{ij}}, where y_i are fractions and \phi_{ij} account for molecular interactions; though derived for dilute gases, it informs hybrid models for vapor-liquid mixtures. In suspensions of rigid particles in liquids, the Einstein predicts an increase in viscosity due to hydrodynamic interactions: \eta = \eta_0 (1 + 2.5 \phi), where \eta_0 is the viscosity and \phi is the volume fraction of spheres. Valid for dilute regimes up to approximately 5% solids by volume, this relation highlights the rotational and translational perturbations caused by suspended particles, as derived from low-Reynolds-number hydrodynamics. Extensions beyond this limit incorporate higher-order terms for denser suspensions, but the establishes the baseline scaling for colloidal systems like paints and slurries. For electrolyte solutions, the Jones-Dole equation models relative viscosity as \eta_r = 1 + A c^{1/2} + B c, where c is concentration, A reflects ion-ion interactions (often negligible in dilute limits), and B the ion-solvent effects. Proposed in , it quantifies how s like NaCl increase viscosity through shells, with positive B values indicating structure-making ions and negative for structure-breaking ones, applicable up to about 1 M. Recent computational advances, such as quantitative structure-activity relationship (QSAR) models for solutions, leverage on molecular descriptors to predict viscosities; for instance, combining simulations with experimental data achieve high accuracy for lubricants and other polymers, addressing gaps in traditional empirical fits for complex solvents. These steady-state models assume time-independent Newtonian behavior and have limitations in capturing transient effects, such as in non-Newtonian mixtures where viscosity evolves with history.

Examples and Data

Water

serves as a Newtonian liquid in viscosity studies, exhibiting ideal linear response to without time-dependent effects, making it a standard reference for and theoretical comparisons. Its viscosity properties are extensively documented, providing a foundation for understanding fluid behavior in aqueous systems. The dynamic viscosity (η) of displays a pronounced dependence, reaching a maximum of approximately 1.79 mPa·s at 0°C, where strengthened bonding creates a more rigid molecular network that resists flow. As rises, thermal agitation weakens these bonds, reducing η to 1.002 mPa·s at 20°C and further to 0.282 mPa·s at 100°C. This behavior includes an anomalous maximum near 0°C, arising from structural transitions in the liquid phase where cooling promotes the formation of transient, ordered clusters that enhance intermolecular cohesion beyond simple thermal expectations. Kinematic viscosity (ν = η / ρ), which accounts for density (ρ), couples these effects and is particularly relevant for applications involving gravitational flow. Water's density peaks at 4°C (approximately 1000 kg/m³), leading to a nuanced temperature profile for ν; for example, it measures 1.787 × 10^{-6} m²/s at 0°C, 1.004 × 10^{-6} m²/s at 20°C, and 0.294 × 10^{-6} m²/s at 100°C, reflecting the interplay between decreasing η and varying ρ. Under pressure, water's viscosity remains largely unaffected up to 100 , with increases typically less than 1% at ambient temperatures, owing to its low that preserves the hydrogen-bonded structure. This stability underscores water's role as a reliable medium in high-pressure contexts.

Air

Dry air serves as a reference for gas-phase viscosity due to its prevalence in atmospheric and engineering applications. The dynamic viscosity of dry air at 20°C and 1 atm is approximately 18.1 μPa·s. This value aligns well with predictions from the , which derives transport properties from kinetic theory for dilute gases like air. The dynamic viscosity of air increases with temperature, approximately following a T^{0.7} dependence as described by the Sutherland formula: \mu = \mu_0 \left( \frac{T}{T_0} \right)^{3/2} \frac{T_0 + S}{T + S}, where \mu_0 is the reference viscosity at temperature T_0, and S \approx 110.4 K is the Sutherland constant for air. At standard temperature and pressure (STP, defined here as 20°C and 1 atm for atmospheric contexts), the kinematic viscosity \nu = \mu / \rho is approximately $15.1 \times 10^{-6} m²/s, where \rho \approx 1.20 kg/m³ is the density of dry air. The presence of water vapor slightly increases the viscosity of air, with measurements showing an enhancement of up to a few percent at high relative humidities (e.g., 90%) and moderate temperatures (20–50°C), due to the higher viscosity of water vapor compared to dry air components. This effect is generally small and often negligible for many applications unless humidity exceeds 50%. In the Earth's atmosphere, air viscosity varies primarily with temperature under the U.S. Standard Atmosphere model, decreasing in the troposphere (up to ~11 km altitude) as temperature drops from 15°C at sea level to -56°C, then increasing in the stratosphere due to rising temperatures. Dynamic viscosity remains nearly independent of pressure (altitude-induced density changes), with values ranging from ~17 μPa·s at sea level to ~12 μPa·s at 11 km and back toward ~18 μPa·s at 20 km.

Other Common Substances

Beyond water and air, viscosity varies widely among other common substances, influenced by molecular structure, temperature, and shear conditions. For instance, everyday fluids like exhibit dynamic viscosities typically ranging from 2 to 10 ·s at , though values can reach up to 23 ·s depending on content and floral origin, and honey displays non-Newtonian behavior where viscosity decreases under shear. Engine oils, critical for , have viscosities of approximately 0.005 to 0.015 ·s at operating temperatures (around 100°C) for typical grades, but this is highly temperature-sensitive, dropping significantly as heat increases to maintain flow in engines. Biological fluids such as human blood show apparent viscosities of 3 to 4 mPa·s at physiological rates, with shear-thinning properties that reduce resistance during circulation. Among gases, at () has a viscosity of approximately 1.5 × 10^{-5} ·s, while exhibits a similar low value of about 2.0 × 10^{-5} ·s at 20°C with notably weak dependence compared to polyatomic gases. Industrial materials like melts and molten demonstrate much higher viscosities, often in the range of 10^3 to 10^6 ·s during , reflecting their entangled molecular networks that impede . Recent chemical advances in viscosity reducers for heavy oils, including multi-effect formulations, have achieved reductions of 50% to 90% at elevated temperatures (50–90°C), enhancing in reservoirs. The following table summarizes representative viscosity values for these substances under typical conditions, illustrating the broad spectrum from low-viscosity gases to highly viscous melts:
SubstanceTypical Viscosity (Pa·s)Conditions/NotesSource
Honey2–10At 20–25°C; non-Newtonian, shear-thinning
Engine Oil (SAE grades)0.005–0.015At 100°C for SAE 10–50; strong temperature dependence
Human Blood0.003–0.004At physiological shear rates (~100 s^{-1}); shear-thinning
CO_2 Gas1.5 × 10^{-5}At STP (0°C, 1 atm)
Helium Gas2.0 × 10^{-5}At 20°C; minimal T dependence
Polymer Melts10^3–10^5At processing temps (150–250°C)
Molten Glass10^3–10^6At working temps (800–1400°C)
Heavy Oil (with reducers)50–90% reduction from baseline (~10^3–10^4)At 50–90°C with chemical additives

Order of Magnitude Estimates

Viscosity values across different states of matter provide a sense of scale for fluid behavior, with gases exhibiting the lowest magnitudes. For typical gases like air at room temperature and atmospheric pressure, dynamic viscosity is on the order of $10^{-5} Pa·s. Liquids span a broader range, from approximately $10^{-3} Pa·s for water to around 1–10 Pa·s for viscous substances like syrup. Amorphous solids, such as glasses near their transition temperatures, display viscosities exceeding $10^{6} Pa·s, effectively behaving as rigid over short timescales despite their fluid-like nature at geological scales. The overall range of viscosities in fluids spans more than 20 orders of magnitude, from superfluid with an effective viscosity below $10^{-15} ·s to highly viscous materials like at approximately $10^{8} ·s. This vast spectrum underscores viscosity's role in distinguishing flow regimes, from inviscid superfluids to near-solid-like resistances. Such estimates draw from examples like those in common substances, providing baselines for broader material classes. Temperature significantly influences viscosity scaling. In gases, viscosity generally increases with temperature, following an approximate relation \eta \propto T^{1/2} derived from kinetic theory, as molecular collisions enhance momentum transfer. For liquids, viscosity decreases exponentially with rising temperature, often modeled as \eta = A \exp(B/T), where higher overcomes intermolecular forces. In nanofluids, and concentration introduce additional effects, where effective viscosity can deviate from bulk values by factors of 10–20% due to interfacial interactions and . These order-of-magnitude estimates prove useful in for rapid assessments, such as approximating the \mathrm{Re} = \rho v L / \eta to predict laminar or turbulent flow without precise data, facilitating initial design iterations in fluid systems.

References

  1. [1]
    DEFINITION OF VISCOSITY
    Viscosity is a measure of a fluid's resistance to flow. It describes the internal friction of a moving fluid.
  2. [2]
    [PDF] Viscosity.pdf - Galileo
    Viscosity is, essentially, fluid friction. Like friction between moving solids, viscosity transforms kinetic energy of (macroscopic) motion into heat energy.
  3. [3]
    Viscosity
    The shearing of the flow of air creates a layer of air near the surface, called a boundary layer, which, in effect, changes the shape of the object. The flow of ...
  4. [4]
    Viscosity - Etymology, Origin & Meaning
    Originating from late 14th-century Old French and Medieval Latin, "viscosity" means the state of flowing slowly or stickiness, also denoting internal ...
  5. [5]
    EarthWord – Viscosity | U.S. Geological Survey - USGS.gov
    Apr 11, 2016 · Etymology: Speaking of molasses, viscosity comes from the Latin viscosus, which meant “sticky.”
  6. [6]
    [PDF] Aristotle's Physics: A Physicist's Look - Research
    The third term is the dissipative force due to the resistance of the fluid. (water or air) in which the body is immersed; ⃗v is the velocity of the body, and C ...
  7. [7]
    [PDF] 200 Years of the Navier-Stokes Equation - arXiv
    Jan 25, 2024 · The inclusion of viscous forces into the equations of fluid motion has been first proposed by Claude-Louis Navier in 1822 [1].
  8. [8]
    Fluid Properties: Dynamic Viscosity
    Dynamic Viscosity. SYMBOL: m. DEFINITION: Shear stress (t) in a Newtonian fluid is linearly proportional to the time rate of angular strain (dV/dy). The ...
  9. [9]
    [PDF] CHAPTER 6 Viscosity & The Stress Tensor
    Viscosity: a measure of a fluid's resistance to deformation by shear stress. For liquids, viscosity corresponds to the informal concept of ”thickness”. A fluid ...
  10. [10]
    [PDF] Variational Approach to the Volume Viscosity of Fluids
    The shear or dynamic viscosity of a fluid can be mea- sured unambiguously in a variety of viscometric apparatuses. By imposing a specific velocity gradient ...
  11. [11]
    [PDF] Review of fluid dynamics
    This is a constitutive equation for a Newtonian fluid; λ is the bulk viscosity and. µ is the dynamic shear viscosity. Note that in an incompressible fluid, the ...
  12. [12]
    [PDF] Brief History of Fluid Mechanics - Joseph Majdalani
    In 1687, Isaac Newton (1642-1727) postulated his laws of motion and the law of viscosity of the linear fluids now called newtonian.
  13. [13]
    [PDF] What is Viscosity?
    The very definition of viscosity implies the existence of what is called laminar flow: the movement of one layer of fluid past another with no transfer of ...Missing: physics | Show results with:physics
  14. [14]
    Fundamental Properties of Fluids – Introduction to Aerospace Flight ...
    What is kinematic viscosity? Dynamic viscosity is a measure of shear resistance; i.e., viscosity matters only in the presence of motion or dynamics. In many ...
  15. [15]
    [PDF] CHAPTER 3 THE CONCEPT OF VISCOSITY
    The behavior of a fluid in flow is very much related to two intrinsic properties of the fluid: density and viscosity. For example, a solid body moving through a ...
  16. [16]
    [PDF] Viscosity
    Newton's law of viscosity (17-1) is a linear relation between the stress and the velocity gradient, only valid in a particular geometry. As for Hooke's law in.<|control11|><|separator|>
  17. [17]
    Newtonian Fluid - an overview | ScienceDirect Topics
    Newtonian fluids are fluids that exhibit a directly proportional relationship between the stress and strain. The behaviour of non-Newtonian fluids can be ...
  18. [18]
    [PDF] Fluid Mechanics II Viscosity and shear stresses
    For fluids the shear stress τ is a function of the rate of strain dγ/dt. The property of a fluid to resist the growth of shear deformation is called viscosity.Missing: η | Show results with:η<|control11|><|separator|>
  19. [19]
    Newtonian fluids – Knowledge and References - Taylor & Francis
    Examples of Newtonian fluids are water, oil, air, and glycerine. The absolute or dynamic viscosity of various substances can be obtained from data in the ...
  20. [20]
    Viscosity of Newtonian and Non-Newtonian Fluids - RheoSense
    This means that the viscosity of Newtonian fluids will remain a constant (see Figure 2) no matter how fast they are forced to flow through a pipe or channel ( ...
  21. [21]
    Viscosity of an ideal gas - tec-science
    Sep 27, 2020 · With the help of the kinetic theory of gases, the viscosity of ideal gases can be calculated. ... η=13⋅ρ⋅λ⋅¯vT. This derivation exploited the fact ...Missing: \bar
  22. [22]
    Newtonian Liquids - an overview | ScienceDirect Topics
    Liquids are defined as Newtonian when their behavior is purely viscous. This means that their viscosity depends uniquely on temperature and not on shear rate, ...
  23. [23]
    [PDF] Understanding Rheology of Structured Fluids - TA Instruments
    Shear thickening often result from material instability and structure rearrangements or phase separation. Thixotropy. For many fluid materials, viscosity is ...
  24. [24]
    [PDF] Non-Newtonian Fluids: An Introduction
    1. Shear- thinning or pseudoplastic behavior. 2. Visco-plastic behavior with or without shear-thinning behavior. 3. Shear- thickening or dilatant ...
  25. [25]
    Power Law Model - an overview | ScienceDirect Topics
    The power-law model proposed by Ostwald and de Waele approximates the relationship between the shear stress and shear rate (Mezger, 2014):. (6.3) τ = c · γ p.Missing: original | Show results with:original
  26. [26]
    [PDF] A New Viscosity Model for Non–Newtonian Fluids : Part I
    One of the widely employed simplified models in specific types of research is the Ostwald–de Waele's relation (Ostwald,. 1929), also known as the power law ...
  27. [27]
    Blood Rheology: Key Parameters, Impact on Blood Flow, Role ... - NIH
    Blood is a shear-thinning fluid, meaning that its viscosity decreases when shear rate increases. Connes et al. (2013b) shown that 15 min of cycling exercise ...
  28. [28]
    [PDF] Momentum Transport, Viscosity
    Feb 15, 2011 · Another important implication of fluid viscosity is the synovial fluid that lines the cartilage of knee joints and helps to lubricate the joint ...
  29. [29]
    Viscosity - Galileo
    Sep 28, 1997 · Viscosity is the transfer of momentum from one part of a fluid to an adjacent part, and in a dilute gas this transfer will be more effective if the collisions ...Missing: behavior | Show results with:behavior
  30. [30]
    [PDF] Navier-Stokes Equations - geo
    The viscous stress term in the Navier-Stokes equation can be seen as the divergence of the diffusive flux of momentum where the momentum diffusion ...
  31. [31]
    [PDF] Boundary-Layer Control - MIT
    This drag is largely due to viscous shear forces and is called skin-friction ... drawing the low-momentum layers from the bottom of the boundary layer ...
  32. [32]
    [PDF] Drag Reduction by Polymers - Physics Courses
    The effect of solvent viscosity on drag reduction is still uncertain ... total (viscous plus turbulent) axial momentum transport in the radial ...
  33. [33]
  34. [34]
    [PDF] LII. The viscosity of gases and molecular force - Stanford University
    May 8, 2009 · William Sutherland. To cite this article: William Sutherland (1893) LII. The viscosity of gases and molecular force ,. Philosophical Magazine ...
  35. [35]
    On the Viscosity or Internal Friction of Air and Other Gases - jstor
    CLERK MAXWELL, M.A., F.B.S.. Received November 23, 1865,-Read February 8 ... 268 ON THE VISCOSITY OR INTERNAL FRICTION OF AIR AND OTHER GASES. In MT ...
  36. [36]
    [PDF] 2. Kinetic Theory
    Our derivation of the Navier-Stokes equation relied on the dilute gas approximation. However, the equation is more general than that. Indeed, it can be ...
  37. [37]
    Molecular theory of gases and liquids - Internet Archive
    Jul 18, 2022 · Hirschfelder, Joseph O., 1911-1990. Publication date: 1954. Topics: Molecular theory, Gases, Liquids, Théorie moléculaire, Gaz, Liquides, gas ...
  38. [38]
    Predicting viscosities and thermal conductivities from dilute gas to ...
    Dec 16, 2024 · In the 1980s, an extension of Chapman–Enskog theory to dense gaseous mixtures of hard spheres termed revised Enskog theory (RET) was developed.
  39. [39]
    [PDF] Calculation of the transport properties of dense fluids using modified ...
    In this work, the modified Enskog theory (MET) along with the modified linear isotherm regularity (MLIR) [14,15] are used to predict the viscosity of dense ...
  40. [40]
    Viscosity of liquids—Enskog-2σ model - ScienceDirect.com
    Enskog-2σ model has been extended to analyze the viscosity of liquids. •. The model is based on Enskog hard sphere formulation of kinetic theory.
  41. [41]
    Viscosities of Simple Liquids - Temperature Variation
    Mar 6, 2021 · If a fluid obeys Equation 5 , then a plot of viscosity versus reciprocal absolute temperature should be linear and the slope can be used to ...
  42. [42]
    A New Equation Relating the Viscosity Arrhenius Temperature and ...
    May 3, 2015 · We propose a novel equation modeling the relationship between the two parameters of viscosity Arrhenius-type equation, such as the energy (E a ) and the ...
  43. [43]
    Effects of Hydrogen Bond Networks on Viscosity in Aqueous Solutions
    Sep 5, 2024 · We found that the proportion of potential viscosity in aqueous solutions primarily increases due to the enhancement of non-hydrogen bonding interactions.
  44. [44]
    Thermodynamic scaling of the viscosity of van der Waals, H-bonded ...
    Herein we extend the scaling to characterize the and dependences of the viscosity for 12 materials (Table I). For the nonpolar liquids (octane, toluene, BCH, ...
  45. [45]
    [PDF] Free‐Volume Model of the Amorphous Phase: Glass Transition
    Free volume VI is defined as that part of the thermal expansion, or excess volume <lii which can be re- distributed without energy change.
  46. [46]
    Liquid-glass transition, a free-volume approach | Phys. Rev. B
    Aug 1, 1979 · The free-volume model, which has been useful for describing the behavior of the viscosity η of dense liquids and glasses, is extended to account for their ...
  47. [47]
  48. [48]
    [PDF] ENGINEERING VISCOELASTICITY - MIT
    Oct 24, 2001 · Figure 9: The Maxwell model. The “Maxwell” solid shown in Fig. 9 is a mechanical model in which a Hookean spring and a. Newtonian dashpot are ...
  49. [49]
    Perspective: Supercooled liquids and glasses - AIP Publishing
    Aug 23, 2012 · The first is the massive increase in the shear viscosity (more than a factor of 1014) that results in the continuous transition from a low ...
  50. [50]
    Maxwell Model - an overview | ScienceDirect Topics
    The Maxwell model is a type of simplest and basic mathematical model to describe the mechanics characteristic of viscoelastic solid material.
  51. [51]
    A modified viscous flow law for natural glacier ice - PNAS
    May 30, 2024 · We present a framework for calculating the ice-viscosity parameters from ice flow conditions, such as stress, ice temperature, and ice crystal size.
  52. [52]
    Pitch Drop experiment - School of Mathematics and Physics
    The experiment demonstrates the fluidity and high viscosity of pitch, a derivative of tar that is the world's thickest known fluid and was once used for ...Missing: glaciers | Show results with:glaciers
  53. [53]
    Viscosity and Glass Transition in Amorphous Oxides - Ojovan - 2008
    Mar 15, 2009 · Near the glass-transition temperature, the elasticity modulus of a glass G ~ 1010 Pa [45], hence at η = 1012 Pa⋅s, where the practical purpose ( ...
  54. [54]
    [PDF] On the Foundations of Eddy Viscosity Models of Turbulence
    Sep 29, 2020 · Eddy viscosity models are based on two conjectures: Boussinesq: Turbulent fluctuations have a dissipative effect on the mean flow. EV hypothesis ...
  55. [55]
    [PDF] Reynolds Equation and Mixing Length Model - Sites
    It is now known that the mixing length hypothesis works for turbulent flows, which are characterized by single length and velocity scales. The reason for its ...
  56. [56]
    [PDF] Turbulence Models and Their Application to Complex Flows R. H. ...
    RANS equations; however, the eddy viscosity is not added to the effective viscosity. The boundary layer parameters required for the solution of the en.
  57. [57]
    On the Foundations of Eddy Viscosity Models of Turbulence - MDPI
    Eddy viscosity models also fail in other cases, such as turbulent flow not at statistical equilibrium so there is a robust effort to develop improved models, ...
  58. [58]
    Two-capillary viscometer for temperatures and pressures relevant to ...
    May 22, 2024 · The assumptions for the basic Hagen–Poiseuille equation to be valid are: (i) the fluid is incompressible, (ii) the fluid is Newtonian, (iii) ...<|separator|>
  59. [59]
    Capillary Viscometer for Fully Automated Measurement of the ... - NIH
    Viscosity is calculated directly from Poiseuille's Law, given the measured difference in pressure between two ends of a capillary tube through which the ...
  60. [60]
    Stokes' law, viscometry, and the Stokes falling sphere clock - Journals
    Aug 3, 2020 · Stokes' law is the drag force on a sphere in a fluid. A falling sphere clock uses the time of a sphere falling through a fluid to measure time.
  61. [61]
    Measurement of Viscosity in a Vertical Falling Ball Viscometer
    Oct 27, 2008 · A falling ball viscometer measures viscosity by observing a falling sphere in a tube, using Newton's law and Stokes' law to calculate viscosity.
  62. [62]
    Discovery Hybrid Rheometers - TA Instruments
    Oscillatory rheology measures viscoelasticity (Storage Modulus, Loss Modulus, Tan Delta) of materials ranging from low-viscosity fluids to stiff solids in DMA ...
  63. [63]
    Portable Microfluidic Viscometer for Formulation Development and ...
    Aug 2, 2024 · Here we introduce a portable single-use analytical chip-based viscometer for determining the viscosity of protein solutions using low sample volumes of 10 μL.
  64. [64]
    Rapid Temperature-Dependent Rheological Measurements of Non ...
    Feb 15, 2022 · We here introduce a microfluidic rheometer for rapid simultaneous measurement of shear viscosity and longest relaxation time of non-Newtonian solutions at ...
  65. [65]
    NIST Guide to the SI, Chapter 4: The Two Classes of SI Units and ...
    Jan 28, 2016 · SI coherent derived unit. Derived quantity, Name, Symbol, Expression in terms of SI base units. dynamic viscosity, pascal second, Pa · s, m−1 · ...
  66. [66]
    NIST Guide to the SI, Chapter 5: Units Outside the SI
    Jan 28, 2016 · (a) The poise (P) is the CGS unit for viscosity (also called dynamic viscosity). The SI unit is the pascal second (Pa · s). (b) The stokes (St) ...
  67. [67]
    Viscosity - The Physics Hypertextbook
    Viscosity is a fluid's resistance to flow, formally the ratio of shear stress to velocity gradient. It is also called dynamic, absolute, or simple viscosity.Missing: dilute λ v_avg
  68. [68]
    Dynamic, Absolute, and Kinematic Viscosity – Definitions ...
    In the SI-system the theoretical unit of kinematic viscosity is m2/s - or the commonly used Stoke (St) where. 1 St (Stokes) = 10-4 m2/s = 1 cm2/s. Stoke comes ...
  69. [69]
    Chapman Enskog Viscosity Model — IDAES v2.1.0 - Read the Docs
    Pure component gas viscosity is given by the formula in which μ is the dynamic viscosity, M is the molar mass, σ is the collision diameter of the molecule.
  70. [70]
    Kinetic Equations for Polyatomic Gases: The 17‐Moment ...
    As examples, the 17‐moment approximation is applied to near equilibrium and relaxation type flows; and all the essential features of the Wang Chang et al.
  71. [71]
    Significant structure theory of liquids | Journal of Chemical Education
    The theory of liquids presented in this paper is a model approach, is analogous to the models customarily used for solids and gases, and can be applied to a ...
  72. [72]
    Significant Liquid-Structure Theory of Viscosity and Self-Diffusion of ...
    Significant liquid-structure theory has been used to calculate the transport properties of the alkali metals. Unlike previous calculations, the volume and ...
  73. [73]
    Arrhenius Equation for Calculating Viscosity in Assessing the ... - MDPI
    Jan 16, 2024 · This article proposes using the Arrhenius model for estimating the viscosity of a mixture of two liquids in the quantitative assessment of the physicochemical ...
  74. [74]
    A Viscosity Equation for Gas Mixtures - AIP Publishing
    A general equation for viscosity as a function of molecular weights and viscosities of the pure components of the mixture.
  75. [75]
    A Local Version of Einstein's Formula for the Effective Viscosity of ...
    We prove a local variant of Einstein's formula for the effective viscosity of dilute suspensions, that is, $\mu^\prime=\mu ({{1+\frac 5 2\phi+o(\phi)}})$, ...
  76. [76]
    Advancing material property prediction: using physics-informed ...
    Mar 14, 2024 · The QSPR models reveal that including MD descriptors improves the prediction of experimental viscosities, particularly at the small data set ...Missing: QSAR | Show results with:QSAR
  77. [77]
    Machine learning approach to predict the viscosity of ... - NIH
    Oct 9, 2025 · This study is the first to use machine learning models to predict the viscosity of perfluoropolyether. This research predicts the viscosity of ...
  78. [78]
    A time-dependent power law viscosity model and its application in ...
    Feb 1, 2017 · The power law viscosity model, as expressed by Eq. (1), is the most widely used model to represent the viscosity change of semi-solid alloys during cavity ...
  79. [79]
    Water - Dynamic and Kinematic Viscosity at Various Temperatures ...
    The figures and tables below shows how water viscosity changes with temperature (°C and °F) at water saturation pressure (which for practicle use, ...
  80. [80]
    Origin of the anomaly in the behavior of the viscosity of water near 0°C
    Aug 8, 2025 · The analysis of the data has revealed an anomaly in the behavior of the viscosity of water near 0°C. The appearance of this anomaly has been ...
  81. [81]
    Pressure dependence of viscosity | The Journal of Chemical Physics
    We show that, at moderate values of pressure, viscosity increases, as a rule, with increasing pressure, provided the liquid is in stable or metastable ( ...III. A FIRST EXCEPTION... · IV. A FIRST... · STRUCTURAL CHANGES OF...
  82. [82]
    Thermophysical Properties of Fluid Systems - the NIST WebBook
    Accurate thermophysical properties are available for several fluids. These data include the following: Density; Cp; Enthalpy; Internal energy; Viscosity; Joule- ...Missing: dry 20°
  83. [83]
    [PDF] Viscosity and Thermal Conductivity of Dry Air in the Gaseous Phase
    V53 giving the kinematic viscosity have excessive scatter and show quite different density dependence from that of the data of Latto et al., although the ...<|separator|>
  84. [84]
    Air Viscosity: Dynamic and Kinematic Viscosity at Various ...
    Online calculator, figures and tables with dynamic (absolute) and kinematic viscosity for air at temperatures ranging -100 to 1600°C (-150 to 2900°F)
  85. [85]
    The viscosity of dry and humid air - ScienceDirect.com
    The paper presents the results of new measurements on the viscosity of dry air along seven isotherms in the range 25–250°C and 1 to 140 atm, subject to an ...
  86. [86]
    [PDF] Interpolation Correlations for Fluid Properties of Humid Air in the
    Knowledge of the properties of humid air, including den- sity, viscosity, thermal conductivity and specific heat, is im- portant in many areas of engineering.
  87. [87]
    [PDF] US Standard Atmosphere, 1976
    Speed of sound as a function of geometric altitude. Dynamic viscosity as a function of geometric altitude. Kinematic viscosity as a function of geometric ...
  88. [88]
    Effect of moisture content on the viscosity of honey at different ...
    Aug 9, 2025 · Honey, for instance, has a viscosity between 2.54 and 23.4 Pa⋅s (at 25 ºC, depending on moisture and sugar composition) [46] , while blood ...
  89. [89]
    Analytical Rheology of Honey: A State-of-the-Art Review - PMC
    This review summarizes the rheological behaviour of honeys of different botanical source(s) and geographical locations that has been described in the ...
  90. [90]
    Reference Correlation for the Viscosity of Carbon Dioxide
    Aug 6, 2025 · The final viscositycorrelation covers temperatures from 100 to 2000 K for gaseous CO2 and from 220 to 700 K with pressures along the melting ...
  91. [91]
    Reference Values and Reference Correlations for the Thermal ... - NIH
    At 20 °C, the viscosity of helium is 19.598 μPa s with an expanded relative uncertainty of 0.3 % at pressures to 10 MPa and an expanded relative uncertainty of ...
  92. [92]
    [PDF] Chapter 5 - Rheology of molten polymers - ResearchGate
    The viscosity of water is 10 3 Pa$s, while the viscosity of most polymer melts during their processing may vary from 102 Pa$s to 105 Pa$s. The shear stress is ...
  93. [93]
    [PDF] High-temperature viscosity of many-component glass melts
    Viscosities ranged from 0.4 to. 462 Pa s (17 Pa s average), and the temperatures ranged from 941 to. 1315 °C (1120 °C average). 2. Approximation function. The ...
  94. [94]
    Flow behaviors and residual oil characteristics of water flooding ...
    The results showed that the viscosity reduction rate of 4# viscosity reducer was more than 90% at 50–90°C, and it could form a stable oil/water emulsion when ...
  95. [95]
    Viscosity Scale - Smooth-On
    Viscosity Scale ; Honey or Corn Syrup, 2,000 - 3,000, centipoise (cps) ; Molasses, 5,000 - 10,000, centipoise (cps) ; Chocolate Syrup, 10,000 - 25,000, centipoise ...
  96. [96]
    The rheology of solid glass | The Journal of Chemical Physics
    May 12, 2010 · Of all the states of matter, a supercooled fluid approaching the glass transition has the highest value for the limiting zero frequency shear ...
  97. [97]
    [PDF] helium refining by superfluidity
    Superfluid helium has an extremely low viscosity, less than about lo-" poise, as compared with the normal viscosity for liquid helium which is greater than ...
  98. [98]
    World's slowest-moving drop caught on camera at last | Nature
    Jul 18, 2013 · The Trinity College team has estimated the viscosity of the pitch by monitoring the evolution of this one drop, and puts it in the region of 2 ...
  99. [99]
    Viscosity of liquids and gases - tec-science
    Mar 25, 2020 · The viscosity of a fluid is a measure of the internal resistance to flow! It is caused by intermolecular forces and transport of momentum within the fluid.Derivation Of Viscosity · Newton's Law Of Fluid Motion · Temperature Dependence Of...
  100. [100]
    A brief review on viscosity of nanofluids | International Nano Letters
    Oct 7, 2014 · In 1970, Nielsen [6] suggested the power law model to determine the viscosity of nanofluids of particle volume fraction more than 0.02 and the ...