Fact-checked by Grok 2 weeks ago

Plug flow

Plug flow refers to an idealized model of flow in chemical reactors, particularly the plug flow reactor (PFR), where the moves through a cylindrical as a series of discrete, non-mixing "plugs" with uniform composition within each plug, no axial mixing, and complete radial mixing, ensuring that all elements experience the same based on their position. In a plug flow reactor, reactants enter at one end of the tube and flow continuously toward the outlet, undergoing progressively along the length, resulting in a in the axial direction while maintaining steady-state operation. This model assumes laminar or turbulent flow with negligible in the direction of flow, treating each plug as a with varying exposure time to conditions. Unlike continuous stirred-tank reactors (CSTRs), which feature complete backmixing and uniform composition throughout, PFRs achieve higher conversion rates for most due to the absence of dilution streams, making them suitable for reactions where selectivity and yield are critical. Plug flow reactors are widely applied in industrial processes involving gas or liquid phases, such as cracking to produce , synthesis, oxidation of at high temperatures (800–1100°C), and even bioreactors for production, often configured as single long tubes or parallel arrays of shorter ones with diameters from centimeters to meters. Their advantages include simple construction without moving parts, high throughput for fast reactions, and efficient space utilization, though challenges arise in for exothermic reactions, potentially leading to hot spots, and the need for precise flow management to approximate ideal plug conditions in practice.

Fundamentals

Definition

Plug flow is an idealized model in and that describes a moving through a conduit, such as a or , with a uniform across the entire cross-section. In this model, the behaves as if it consists of discrete cylindrical plugs advancing sequentially along the path, each maintaining the same speed without any variations or between layers. This assumption eliminates gradients in the radial direction, simplifying the analysis of in steady-state conditions. The concept of plug flow emerged in the early as part of the foundational developments in , providing a straightforward for flows where viscous effects are negligible or flattens the velocity profile. It was particularly adopted to model steady-state flows in and conduits, allowing engineers to focus on axial transport without the complexities of detailed profile calculations. By mid-century, the model gained prominence in chemical reaction engineering for its utility in describing ideal behavior. Visually, plug flow can be analogized to a train of solid plugs sliding through a tube, where each element travels at the mean , preserving spatial uniformity and avoiding back-mixing or axial . This representation highlights the model's core idealization: fluid elements retain their relative positions and composition as they progress, akin to a pushing successive volumes forward. In practice, this model is referenced in the design of tubular reactors in chemical processes, where it approximates conditions with minimal longitudinal mixing.

Key Assumptions

The plug flow model in and relies on several simplifying s to idealize the behavior of motion in conduits or reactors, treating the flow as a series of discrete plugs advancing without along the flow direction. A primary is the uniform profile, where there are no radial or axial variations in ; instead, the elements move as coherent plugs at a constant speed across the entire cross-section of the flow path. This flat profile implies that all particles at a given cross-section travel at the same , resembling piston-like motion without or gradients. Another key assumption is the negligible diffusion in the axial direction, meaning there is no molecular or that would cause mixing between adjacent plugs along the flow path; however, perfect mixing is assumed in the radial direction to ensure uniformity to the flow. This separation of mixing behaviors simplifies the model by preventing backmixing or longitudinal dispersion, allowing each plug to react independently based on its . The model further assumes steady-state conditions, under which the flow rate, temperature, and composition are uniform across each cross-section and remain constant over time, though they may vary along the flow path, with no transient accumulation or depletion of species. This steady operation ensures that the input conditions at the inlet directly determine the outlet profiles without time-dependent fluctuations. Finally, the plug flow model typically assumes an incompressible fluid, where remains constant throughout the system, thereby simplifying equations by maintaining a constant and eliminating the need to account for variations. This assumption is particularly valid for liquid-phase systems but introduces approximations for compressible gases unless and effects are negligible.

Mathematical Modeling

Velocity Profile Derivation

In the plug flow model, the velocity profile is idealized as uniform across the flow channel's cross-section, with the axial u in the radial . This arises from conditions of intense radial mixing, such as in highly turbulent flows at large s, where the velocity becomes nearly flat in the core region, approaching a plug-like distribution as the Reynolds number tends to infinity. The derivation begins with the Navier-Stokes equations for steady, incompressible, axisymmetric flow in cylindrical coordinates, assuming fully developed conditions where the axial does not vary along the flow direction (\partial u / \partial z = 0) and there are no radial or azimuthal components. The simplified axial equation is then \frac{\mu}{r} \frac{d}{dr} \left( r \frac{du}{dr} \right) = \frac{dp}{dz}, where \mu is the dynamic viscosity, r is the radial coordinate, u is the axial , and dp/dz is the axial . Under the plug flow assumptions, u is independent of r, so du/dr = 0, which nullifies the viscous term on the left-hand side and implies dp/dz = 0 for balance. In practice, this idealization neglects viscous effects in the balance for the profile, with any computed separately via empirical correlations using the average . The constant velocity magnitude follows from the for , ensuring mass conservation without radial variations: u = Q / A, where Q is the and A is the cross-sectional area. This yields a flat profile u(z) = constant across the cross-section. The no-slip condition at the walls is disregarded in plug flow, as it would introduce radial gradients; instead, infinite radial mixing is assumed to diffuse instantly, eliminating boundary layers and enforcing uniformity even near the walls. In a circular of D, the cross-sectional area is A = \pi D^2 / 4, so the uniform is u = 4Q / (\pi D^2), equivalent to the applied throughout.

Design Equations for

In plug , the uniform across the conduit cross-section leads to simplified expressions for key parameters, particularly in the of pressure losses and effects in pipes or channels. These equations are derived under the assumption of steady, fully developed with no variations, making them useful for approximating high-Reynolds-number turbulent flows where the velocity flattens. For applications, the plug flow assumption enables the derivation of the reactor design equation from the balance. For a species A, the is \frac{dF_A}{dV} = r_A, where F_A is the molar flow rate of A, V is the reactor volume, and r_A is the rate of production of A (negative for consumption). Integrating gives the performance equation V = F_{A0} \int_0^{X_A} \frac{dX_A}{-r_A}, where F_{A0} is the inlet molar flow rate and X_A is the fractional conversion. For constant volumetric flow rate and density, this simplifies to \tau = C_{A0} \int_{C_A}^{C_{A0}} \frac{dC_A}{-r_A}, with \tau = V / v_0 the space time and v_0 the inlet volumetric flow rate. This models the progressive reaction along the reactor length under uniform velocity u = v_0 / A. The pressure drop per unit length along the conduit is described by the Darcy-Weisbach equation: \frac{\Delta P}{L} = \frac{f \rho V^2}{2D} where f is the Darcy friction factor, \rho is the fluid density, V is the uniform , and D is the of the conduit. In ideal plug flow approximations for turbulent regimes, f is often treated as constant, independent of the , to reflect the negligible influence of viscous effects in the core flow. The wall shear stress \tau_w arises from the frictional interaction at the conduit boundary and is obtained through a momentum force balance on a differential fluid element: the net pressure force driving the flow equals the shear force resisting it at the wall. For a circular pipe, this yields \tau_w = \frac{D}{4} \frac{\Delta P}{L}. This expression connects the macroscopic pressure gradient directly to the local shear at the wall, providing a fundamental relation for conduit design under plug flow conditions. To characterize the near-wall dynamics in plug flow approximations, the friction velocity u^* is defined as u^* = \sqrt{\frac{\tau_w}{\rho}}. This scale quantifies the intensity of and near the boundary, serving as a reference for nondimensionalizing variables in analyses where plug flow idealizes the bulk motion. In turbulent approximations approaching plug flow, the viscous sublayer thickness \delta_s—the region adjacent to the wall where viscous effects dominate—can be estimated as \delta_s \approx \frac{5 \nu}{u^*}, with \nu denoting the kinematic viscosity of the fluid. This approximation delineates the thin layer where the flow transitions from wall-dominated laminar behavior to the uniform plug-like core, aiding in the assessment of boundary influences on overall flow resistance.

Applications

In Pipe Flow and Fluid Mechanics

In pipe flow and fluid mechanics, the plug flow model serves as a simplifying approximation for turbulent flows in long pipelines, where the velocity profile flattens due to high Reynolds numbers (typically Re > 10^4), enabling straightforward calculations of flow rates and energy losses without accounting for detailed velocity gradients. This assumption treats the fluid as moving with uniform velocity across the pipe cross-section, akin to a solid plug, which is particularly valid in fully developed turbulent regimes where the boundary layer near the wall is thin relative to the pipe radius. By neglecting radial variations, engineers can apply empirical correlations like the Darcy-Weisbach equation to estimate pressure drops, facilitating efficient pipeline sizing and optimization for applications such as fuel transport. In multiphase flows, plug flow modeling is applied to slug flow regimes, where intermittent gas bubbles propagate through a continuous , spanning much of the pipe cross-section and behaving like plugs. This pattern is prevalent in oil and gas pipelines, especially in near-horizontal configurations, where gas bubbles coalesce and advance faster than the surrounding , influencing overall and requiring models to predict slug characteristics for safe operations. (CFD) simulations using volume-of-fluid methods have validated these models by quantifying slug lengths, velocities, and holdups as functions of superficial velocities, aiding in the design of pipelines to mitigate issues like vibrations or blockages. The plug flow assumption offers significant advantages in simulations of pipe flows by reducing through uniform velocity profiles, making it suitable for preliminary designs where full resolution is unnecessary. In CFD analyses, this simplification avoids the need for advanced models in early-stage evaluations, allowing faster iterations while still capturing essential behaviors like bulk transport and approximate energy dissipation. For instance, in water distribution systems, plug flow approximates turbulent conditions to estimate head losses along mains without detailed modeling of velocity profiles, using standard friction formulas to balance supply pressures and demands across networks. This approach integrates with network solvers to simulate steady-state , ensuring reliable delivery while minimizing computational demands for operational planning.

In Chemical Reactors

In the plug flow reactor (PFR), fluid elements advance through the reactor as discrete plugs without axial mixing, ensuring that each element experiences a uniform reaction time equivalent to the mean \tau = V/Q, where V is the reactor volume and Q is the . This model assumes ideal plug flow conditions, leading to a concentration profile that varies predictably along the reactor length based on reaction kinetics. The design equation for a PFR handling a reaction follows from the differential mole balance, expressed as -\frac{dC_A}{dV} = \frac{[k](/page/K) C_A}{Q}, where C_A is the concentration of reactant A, [k](/page/K) is the rate constant, and Q is the constant . Integrating this under constant volumetric flow yields the outlet concentration ratio \frac{C_A}{C_{A0}} = \exp(-[k](/page/K) \tau), which highlights how reactor size and flow rate directly control . The residence time distribution (RTD) in an ideal PFR is a centered at \tau, signifying perfect uniformity with no bypassing or backmixing, which enhances selectivity for sequential reactions. Plug flow reactors find extensive use in modern flow chemistry for pharmaceutical , where post-2010 advancements in micro- and meso-scale tubing have enabled of active pharmaceutical ingredients with improved safety and efficiency for exothermic processes. They are also employed in processes, such as the production of , where tubular PFRs manage increasing viscosities while achieving narrow molecular weight distributions through controlled residence times. Recent advances (as of 2024) include plug flow reactors for , enhancing production through optimized mixing and operating conditions. A representative example is the oxidation of to in multitubular reactors, where the plug flow configuration maximizes conversion of this irreversible reaction by minimizing radial gradients and ensuring complete reactant exposure.

Comparisons and Limitations

Versus Laminar and Turbulent Flows

Plug flow represents an idealized model in where the is assumed uniform across the pipe cross-section, contrasting sharply with the parabolic profile observed in laminar pipe flow. The Hagen-Poiseuille equation describes the laminar profile as u(r) = 2V \left(1 - \left(\frac{r}{R}\right)^2 \right), where V is the average , r is the radial from the centerline, and R is the pipe radius; this results in maximum at the center twice the average and zero at the wall due to no-slip conditions. By ignoring this variation, plug flow overestimates the uniformity of momentum and mass transport, particularly in low regimes (Re < 2300) where viscous forces dominate and shear rates are highest near the wall. In turbulent flows, occurring at higher Reynolds numbers (typically Re > 4000), the velocity profile flattens significantly due to radial mixing from eddies, approaching the uniform profile of plug flow more closely, especially at Re > 10,000 where the core region exhibits near-constant . However, even in these conditions, a thin viscous sublayer persists adjacent to the wall—typically on the order of y^+ ≈ 5 in wall units—where laminar-like linear velocity gradients govern the , a detail neglected by the plug flow assumption for analytical simplicity. This approximation thus underpredicts wall shear effects while capturing the bulk flow behavior adequately in highly turbulent scenarios. The plug flow model assumes fully developed conditions and breaks down in entrance or developing regions of the , where the evolves over an entry length proportional to the pipe diameter and inversely to ; in real turbulent flows, this transition introduces deviations, with the following the Blasius correlation f \approx 0.316 \, \mathrm{Re}^{-1/4} for smooth pipes, rather than the constant value implied by ideal plug flow. Such deviations highlight the model's limitations in non-equilibrium flows, where axial development and intensity alter uniformity. Regarding performance impacts in pipes, plug flow predicts wall lower than in —where the parabolic profile yields a maximum of \frac{4V}{R}—but higher than in the turbulent core, where momentum transfer occurs primarily through turbulent stresses beyond the sublayer; this affects estimates, often requiring corrections for accurate .

Versus Continuous Stirred-Tank Reactors

Plug flow reactors (PFRs) and continuous stirred-tank reactors (CSTRs) represent two fundamental ideal models in , differing primarily in their mixing characteristics. In a PFR, there is no axial mixing, corresponding to a high (Pe >> 1), where the is defined as Pe = uL/D with u as fluid velocity, L as reactor length, and D as axial dispersion coefficient; this ensures that elements of fluid move through the reactor with uniform and composition gradients along the flow direction. In contrast, a CSTR assumes complete mixing throughout the vessel, resulting in uniform composition equivalent to the outlet conditions and a broad distribution. For the same reactor volume and residence time τ, these mixing differences lead to distinct performance in terms of reactant conversion, particularly for first-order irreversible reactions (A → products, rate = kC_A). The PFR achieves higher conversion because reactants are exposed to progressively lower concentrations without backmixing, following the design equation derived from the mole balance: X_{A,\text{PFR}} = 1 - e^{-k\tau} where X_A is the fractional conversion of A. The CSTR, however, operates at the lower outlet concentration, yielding: X_{A,\text{CSTR}} = \frac{k\tau}{1 + k\tau} This results in the PFR requiring a smaller volume than the CSTR for the same conversion in positive-order kinetics, as illustrated in Levenspiel plots where the PFR volume is the integral under the 1/(-r_A) vs. X curve, while the CSTR volume is a rectangle at the outlet condition. The mixing profiles also influence reaction selectivity, especially in complex kinetic schemes. For series reactions (A → B → C, where B is desired), the PFR's lack of backmixing maintains higher intermediate concentrations early in the reactor, favoring higher selectivity to B compared to the CSTR, which dilutes concentrations uniformly and promotes over-conversion to C. Conversely, for parallel reactions (A → B and A → C with different orders), the CSTR's uniform conditions can enhance selectivity to the product from the lower-order path, whereas the PFR's gradients may favor the higher-order path; thus, CSTRs are often preferred when uniform reactant concentrations are needed to optimize side-reaction suppression. Scale-up considerations further highlight these contrasts. PFRs, typically , scale linearly by increasing length or units while preserving the plug flow profile, making them suitable for high-throughput continuous processes like cracking. CSTRs, however, require adjustments to and volume to maintain mixing uniformity at larger scales, often complicating and potentially leading to hot spots; this makes PFRs generally more straightforward for capacity expansion in flow-dominated systems.

Experimental Determination

Measurement Techniques

Tracer studies are a primary experimental method for verifying plug flow conditions in reactors by assessing the distribution (). In a pulse injection experiment, a non-reactive tracer, such as a or fluorescent , is introduced instantaneously at the , and the outlet concentration is monitored over time using probes or optical sensors. For ideal plug flow, the outlet response exhibits a sharp Dirac delta-like peak at the mean residence time τ with no tailing, corresponding to zero variance (σ² = 0) in the RTD, indicating no axial or mixing along the flow path. Residence time analysis can also employ a step tracer input, where the tracer concentration at the inlet suddenly changes from zero to a constant value at t = 0. The normalized F(t) = C_out(t)/C_in rises instantaneously from 0 to 1 at t = τ in an ideal plug flow system, reflecting uniform transit time for all fluid elements without spreading. This sharp confirms the absence of back-mixing, with experimental deviations quantified by the slope or variance of the . Velocity profiling techniques provide direct spatial measurements to confirm the uniform (flat) cross-sectional velocity characteristic of plug flow. measures point-wise velocities by detecting the Doppler shift of laser light scattered from tracer particles in the flow, allowing reconstruction of the velocity profile across the reactor diameter. In plug flow, LDV data show minimal variation in velocity, with deviations indicating parabolic (laminar) or turbulent profiles. extends this by capturing instantaneous velocity fields over a plane using double-pulsed laser illumination and particle tracking, enabling visualization of flow uniformity and detection of recirculation zones that violate plug flow assumptions. PIV has been applied in tubular reactors to quantify axial and radial velocities, verifying near-flat profiles under high flow rates. In small-scale reactors, modern microfluidic sensors enable dispersion measurement post-2000, often integrating or integrated detectors for continuous monitoring. For instance, piezoelectrically actuated injectors release tracer pulses in microchannels, with outlet spreading tracked via optical detection to compute coefficients, achieving resolutions down to milliseconds for times. These tools, combined with particle tracking in high-speed , assess plug flow in compact systems by analyzing uniformity and variance, supporting applications in lab-on-chip reactors.

Validation Methods

Validation of plug flow adherence in experimental systems relies on analyzing residence time distribution (RTD) data from tracer studies to quantify axial dispersion and deviations from ideal behavior. The Peclet number, defined as Pe = \frac{u L}{D_{ax}}, where u is the fluid velocity, L is the reactor length, and D_{ax} is the axial dispersion coefficient, serves as a key metric; values of Pe \gg 100 indicate near-plug flow with minimal backmixing. This parameter is typically calculated by fitting experimental tracer response curves using the moments method, which computes the mean residence time \bar{t} = \int_0^\infty t E(t) \, dt and variance \sigma^2 = \int_0^\infty (t - \bar{t})^2 E(t) \, dt from the exit age distribution E(t); for open systems with small dispersion, \frac{D_{ax}}{u L} \approx \frac{\sigma^2}{2 \bar{t}^2}, allowing direct estimation of Pe. To further confirm plug flow, experimental RTD curves are compared to the theoretical delta function response expected for plug flow, where all fluid elements have identical residence times. Least-squares optimization is applied to match the measured concentration-time data to the dispersed plug flow model , iteratively estimating D_{ax} and assessing goodness-of-fit; low errors validate the assumption when is negligible. Deviations are quantified using the Bodenstein number Bo = Pe = \frac{u L}{D_{ax}}, which mirrors the and measures the ratio of convective to dispersive transport; low values of Bo < 10 signal significant axial mixing, rendering the plug flow model invalid and necessitating alternative reactor models like the tanks-in-series approach. In a of a continuous oscillatory baffled crystallizer for liquid s, tracer experiments monitored via UV probes revealed near-plug behavior, with Peclet numbers ranging from 100 to 1000 corresponding to dispersion numbers below 0.01 and up to 100% adherence to ideal plug characteristics based on variance analysis.

References

  1. [1]
    Plug Flow (PFR)
    Apr 1, 2022 · Plug flow reactors, also known as tubular reactors, consist of a cylindrical pipe with openings on each end for reactants and products to flow through.Missing: definition | Show results with:definition
  2. [2]
    Plug flow reactor (PFR) - Vapourtec
    The plug flow reactor model (PFR, sometimes called continuous tubular reactor, CTR) is normally the name given to a model used in chemical engineering.
  3. [3]
    Reactor - processdesign
    Feb 21, 2016 · Plug Flow Reactors are tubular reactors that assume perfect radial mixing and approximate zero axial dispersion (Towler, 2012). A PFR is used ...
  4. [4]
    [PDF] When Chemical Reactors Were Admitted And Earlier Roots of ...
    Early chemical innovations included the mid-18th century chamber process for sulfuric acid and the early 19th century Leblanc process for sodium carbonate.Missing: historical | Show results with:historical
  5. [5]
    Plug Flow - an overview | ScienceDirect Topics
    Plug flow is defined as a flow regime where plugs of liquid alternate with pockets of gas, moving along the horizontal line in a pipe, typically occurring ...
  6. [6]
    [PDF] Basic reactor models and evaluation of rate expressions from ...
    4. PLUG FLOW REACTOR (PFR):. The main assumption of the ideal plug flow reactor is that the fluid is perfectly mixed in the direction perpendicular to main flow ...
  7. [7]
  8. [8]
  9. [9]
    None
    ### Key Assumptions for Plug Flow Reactor Model (Chapters 12-14)
  10. [10]
  11. [11]
    [PDF] Lecture 16 Navier-Stokes equation, Reynolds number
    Navier-Stokes Equation (1-D):. ∂vx. ∂t. = ν. ∂2vx. ∂y2 −. 1 ρ. ∂P. ∂x. + ... What if Flow is Turbulent? “Plug Flow”: PπR2. ⇒ Cannot solved explicitly for ...
  12. [12]
    [PDF] Notes on Thermodynamics, Fluid Mechanics, and Gas Dynamics
    Sep 6, 2023 · This condition is quite common. For example, a plug flow where the velocity profile is constant (ux = constant) is. C. Wassgren. 50. 2023- ...
  13. [13]
    [PDF] A First Course on Kinetics and Reaction Engineering Unit 11 ...
    Thus the ideal plug flow reactor is assumed to be totally unmixed in the axial direction and, at the same time, to be perfectly mixed in the radial direction.
  14. [14]
    [PDF] Pipe Flows - Purdue Engineering
    Dec 15, 2021 · The derivation in this section was previously covered in Chapter 8 and is repeated here, in a slightly condensed form, for convenience.<|control11|><|separator|>
  15. [15]
    [PDF] 7. Basics of Turbulent Flow - MIT
    Friction [Shear] Velocity, U*:​​ This velocity scale, u*, is called the shear velocity, or the friction velocity, and it characterizes the shear at the boundary. ...
  16. [16]
    None
    ### Summary of Plug Flow in Turbulent Pipe Flow
  17. [17]
    plug flow - Energy Glossary - SLB
    A multiphase flow regime in pipes in which most of the gas moves as large bubbles dispersed within a continuous liquid. The bubbles may span much of the pipe.
  18. [18]
  19. [19]
    Flow Chemistry: Recent Developments in the Synthesis of ...
    Nov 26, 2015 · In this review we focused our attention only on very recent advances in the continuous flow multistep synthesis of organic molecules which found application as ...
  20. [20]
    Polymerization Processes, 2. Modeling of Processes and Reactors
    Oct 15, 2011 · ... Polymer coupling. Batch and plug flow reactors provide equivalent results, so they can conveniently be considered together. There are thus ...<|control11|><|separator|>
  21. [21]
    A mathematical model for a tubular reactor performing ethylene ...
    A mathematical model for a tubular reactor performing ethylene oxidation to ethylene oxide by a catalyst deposited on metallic strips ... tubular plug flow ...
  22. [22]
    Poiseuille Flow - Thermopedia
    This is the Hagen-Poiseuille Equation, also known as Poiseuille's Law. ... Axial velocity profile. Figure 2. Axial velocity profile. Poiseuille flow is a ...
  23. [23]
    Hagen-Poiseuille Equation - an overview | ScienceDirect Topics
    The Hagen-Poiseuille equation is given as: Δ P = − 8 μL R 2 U or Δ P = − 8 μL π R 4 Q .
  24. [24]
    Experimental study of the viscous sublayer in turbulent pipe flow
    The results show that, adjacent to the wall, there is a layer of very small thickness y+ = 1.6 ± 0.4 in which a linear velocity gradient occurs virtually at all ...
  25. [25]
    Blasius Correlation - an overview | ScienceDirect Topics
    The Blasius correlation refers to a mathematical relationship used to predict the frictional resistance coefficient for fluid flow in circular tubes, ...
  26. [26]
    Derivation of Hagen-Poiseuille equation for pipe flows with friction
    Apr 4, 2020 · This is also called plug flow. Furthermore, turbulent flows can occur at narrowings and high flow velocities, which would then also lead to an ...<|control11|><|separator|>
  27. [27]
    Reactors - processdesign
    Feb 22, 2016 · Plug Flow Reactor (PFR). A PFR with tubular geometry has perfect radial mixing but no axial mixing. All materials hav the same residence time ...Missing: differences | Show results with:differences
  28. [28]
    [PDF] Chemical Reaction Engineering
    Plug Flow Reactor Design Equations c T.L. Rodgers 2013. 31. Page 44. CHAPTER 3. PLUG FLOW REACTORS (PFR). [? ]. 3.6 References. 32 c T.L. Rodgers 2013. Page 45 ...
  29. [29]
    [PDF] PFR vs. CSTR: Size and Selectivity - MIT OpenCourseWare
    Levenspiel plots comparing CSTR and PFR volumes for changing kinetics. Left: The CSTR has the smaller volume. Right: The PFR eventually has the smaller volume.
  30. [30]
    Measurement of a flow-velocity profile using a laser Doppler ...
    Jun 26, 2020 · This paper reports on the measurement of a flow-velocity profile using a laser Doppler velocimetry (LDV) system having a focus tunable lens (FTL).
  31. [31]
    Characterising flow with continuous aeration in an oscillatory baffle ...
    Aug 30, 2023 · ... plug flow reactor type behaviour. The level of plug flow is ... Particle Image Velocimetry, Chem. Eng. Res. Des., 2003, 81(8), 842 ...<|control11|><|separator|>
  32. [32]
    Measurement of residence time distribution in microfluidic systems
    A method uses a tracer injector to release a tracer pulse, and its spreading is monitored with fluorescence microscopy to measure residence time distribution.
  33. [33]
    Residence time distributions in microchannels with assistant flow ...
    Aug 22, 2023 · The average residence time for particles is 5 ms and the shortest residence time of particles is about 2.5 ms, which is half of the mean ...INTRODUCTION · Residence time distribution · Verification · IV. CONCLUSION
  34. [34]
    A note on flow behavior in axially-dispersed plug flow reactors with ...
    An axially-dispersed plug flow reactor (AD-PFR) model applies in these situations of non-ideal, low-Re flow reactors. To better understand the chemical kinetics ...Missing: origin | Show results with:origin
  35. [35]
    Determination of Bodenstein number and axial dispersion of a ...
    In this study, the dispersion characteristic and the flow pattern of a specialized triangular EL-ALR have been studied using RTD analysis.
  36. [36]
    Heat Transfer and Residence Time Distribution in Plug Flow ...
    Jul 9, 2021 · Residence time distribution measurements show a near-plug flow with high Peclet numbers on the order of 100–1000 s, although there was ...