Fact-checked by Grok 2 weeks ago

Power number

The power number, denoted as N_p, is a dimensionless quantity in and that characterizes the power consumption of rotating s in mixing and processes within stirred tanks. It is defined by the relation N_p = \frac{P}{\rho N^3 D^5}, where P is the power input to the , \rho is the , N is the rotational speed, and D is the diameter. This enables the analysis of energy transfer in systems by normalizing power requirements against properties and operating conditions. In practice, the power number correlates strongly with the impeller (Re = \frac{\rho N D^2}{\mu}, where \mu is the fluid viscosity), allowing prediction of flow regimes from laminar to turbulent. In the turbulent regime, typically at Re > 10^4, N_p becomes constant for a specific impeller geometry, facilitating scale-up of mixing operations from pilot to production scales without proportional changes in power draw. Variations in N_p arise from impeller design factors, such as blade shape and number; for instance, radial-flow impellers like Rushton turbines exhibit higher N_p values (around 5–6) compared to axial-flow propellers (around 0.3–1.5), reflecting differences in and circulation efficiency. Baffling in the tank and impeller-to-tank diameter ratios further influence N_p, with unbaffled systems showing oscillatory behavior at transitional Reynolds numbers. The power number's utility extends to optimizing energy use in industries like pharmaceuticals, , and , where it informs selection to balance mixing intensity against power costs. Experimental determination often involves measurements or in controlled setups, with correlations compiled for common geometries to guide design.

Definition and Formulation

Definition

The power number, denoted as N_p, is a dimensionless that characterizes dissipation rates in systems, particularly in and mixing processes where mechanical impellers impart to . It quantifies the relationship between delivered to the fluid and the characteristic scales of the system, enabling consistent comparisons across varying equipment sizes and operating conditions. The power number is defined as the ratio of the power input P to the product of the fluid density \rho, the cube of the rotational speed N^3, and the fifth power of the D^5: N_p = \frac{P}{\rho N^3 D^5} where P is the power consumption in watts, \rho is the fluid density in kg/m³, N is the rotational speed in revolutions per second, and D is the in meters. The power number was introduced in by J. H. Rushton, E. W. Costich, and H. J. Everett, providing a standardized approach to measuring and correlating power requirements in mixing operations and facilitating advancements in design. The term originates from its explicit connection to the power transferred via mechanical agitation, emphasizing the energetic aspect of fluid motion induced by impellers. In turbulent flow regimes, where inertial forces dominate over viscous effects, the power number remains constant for a specific impeller type and geometry, serving as a reliable indicator of energy input independent of fluid viscosity. This property makes it essential for analyzing high-Reynolds-number conditions in industrial mixers. It complements the Reynolds number in delineating flow regimes during agitation.

Mathematical Formulation

The power number, denoted as N_p, arises from dimensional analysis applied to the power consumption in fluid agitation systems, utilizing the Buckingham π theorem to identify relevant dimensionless groups. Consider the power P delivered to an impeller as a function of the fluid density \rho, viscosity \mu, gravitational acceleration g, impeller rotational speed N, and impeller diameter D. These six variables involve three fundamental dimensions: mass (M), length (L), and time (T). According to the Buckingham π theorem, the number of dimensionless π groups is $6 - 3 = 3. To derive the groups, select repeating variables that span the dimensions: \rho ([M L^{-3}]), N ([T^{-1}]), and D ([L]). The first π group incorporates the dependent variable P ([M L^2 T^{-3}])): \pi_1 = P \rho^a N^b D^c Balancing dimensions yields a = -1, b = -3, c = -5, so \pi_1 = N_p = \frac{P}{\rho N^3 D^5}. The second group uses \mu ([M L^{-1} T^{-1}])): \pi_2 = \mu \rho^d N^e D^f This gives d = -1, e = -1, f = -2, resulting in \pi_2 = \frac{\mu}{\rho N D^2} = \frac{1}{\text{Re}}, where \text{Re} = \frac{\rho N D^2}{\mu} is the impeller Reynolds number. The third group incorporates g ([L T^{-2}])): \pi_3 = g \rho^h N^i D^j Balancing yields h = 0, i = -2, j = -1, so \pi_3 = \frac{g}{N^2 D} = \frac{1}{\text{Fr}}, where \text{Fr} = \frac{N^2 D}{g} is the . Thus, the general relationship is N_p = f(\text{Re}, \text{Fr}). In high flows, typical of fully turbulent regimes (\text{Re} > 10^4), inertial forces dominate over viscous forces, rendering N_p independent of \text{Re} and often approximately constant, provided free-surface effects (captured by \text{Fr}) are minimized, such as in baffled tanks. This simplifies to P = \rho N^3 D^5 N_p, where N_p depends primarily on impeller geometry.

Applications in Mixing and Agitation

Power Consumption in Stirred Tanks

In stirred , the power consumption is analyzed through the dimensionless power number N_p, plotted against the impeller Re on a log-log scale, revealing three flow regimes that dictate energy requirements. In the laminar regime (typically Re < 10), viscous forces dominate, and N_p \propto 1/Re, resulting in higher relative power draw at low speeds. The transitional regime ($10 < Re < 10^4) features a sloping curve where N_p decreases gradually as inertial forces increase. In the turbulent regime (Re > 10^4), N_p stabilizes as a constant value, independent of Re, reflecting fully developed . For common configurations in the turbulent regime, N_p \approx 5 for a standard six-bladed Rushton turbine in a baffled tank, providing a benchmark for power prediction in industrial mixing. Several factors influence this power draw, including the presence of baffles, which disrupt circumferential flow to suppress surface vortex formation and promote axial and radial circulation, thereby increasing effective power input compared to unbaffled setups. Additionally, the tank-to-impeller diameter ratio, typically J = D_T / D \approx 3 (or impeller diameter D about one-third of tank diameter D_T), optimizes flow patterns and power efficiency in standard designs. The baseline equation for ungassed power consumption is P = N_p \rho N^3 D^5, where \rho is fluid , N is impeller rotational speed, and D is impeller diameter; this formulation scales power requirements directly with operating conditions. In gassed systems, gas dispersion reduces effective density and alters hydrodynamics, yielding lower power draw expressed as P_g = N_{pg} \rho N^3 D^5, with the gassed power number N_{pg} typically less than N_p, often correlated empirically as the ratio P_g / P versus the gas flow number Q_g / (N D^3) to account for aeration effects.

Impeller-Specific Correlations

Empirical correlations for the power number vary significantly with , reflecting differences in patterns and . In the fully turbulent regime (Re > 10^4), the power number reaches a constant value characteristic of each type, enabling straightforward comparisons of across designs. These values are derived from experimental in standard baffled stirred , where the diameter to (D/T) is typically 0.3–0.5. Representative turbulent power numbers for common impellers are summarized below, highlighting the high power draw of radial-flow impellers like the compared to low-power axial-flow options such as . Note that the , a close-clearance for viscous fluids, operates primarily in the , where its power number is notably lower than turbulent radial impellers but follows a different .
Impeller TypeRegimePower Number (N_p)
Turbulent5.0
Pitched-blade turbineTurbulent1.3
Turbulent0.3
Laminar≈ 300 / Re
These values assume standard geometries and baffled conditions for turbulent cases; deviations can occur with custom designs. In the transitional regime (approximately 10^2 < Re < 10^4), where viscous and inertial forces compete, power numbers decrease nonlinearly with increasing Reynolds number. Empirical correlations bridge the laminar and turbulent limits, often taking forms that include inverse dependence on Re. For specific impeller types, such as pitched-blade or propeller designs, one widely used equation is N_p = \frac{A}{\text{Re}} + B \left(1 + \frac{C}{\text{Re}^{0.5}}\right) where A, B, and C are geometry-dependent constants (e.g., A ≈ 100–300, B ≈ 1–5 for common axial impellers), determined from torque measurements across Re ranges. These allow prediction of power draw without full experimentation, though constants must be fitted to the exact setup. The power number is sensitive to positional geometry within the tank. Impeller clearance from the tank bottom affects hydrodynamic loading; reducing clearance below 0.1T (where T is tank diameter) can increase N_p by 10–20% due to enhanced wall friction and altered discharge flow. Similarly, off-center positioning, even by 5–10% of D, introduces asymmetry in circulation, raising N_p by 10–20% through increased drag and vortex shedding. Such variations underscore the need for centered, optimally cleared installations to maintain correlation accuracy. Accurate correlations generally require baffled tanks to minimize free-surface vortices. In high-solid suspension applications, particle interactions modify the effective power number beyond clear-liquid correlations. For slurries with solids concentrations exceeding 20–30% by volume, N_p increases due to added mass loading, particle-impeller collisions, and hindered flow, often by 20–50% or more depending on particle size and density. A case study with a pitched-blade impeller in a 40% solids kaolin suspension demonstrated N_p rising from 1.3 in water to approximately 1.8–2.0, attributed to enhanced form drag from particle clusters; this necessitates scaling factors like (ρ_slurry / ρ_liquid)^{0.5} in design equations to account for the interactions.

Relation to Reynolds and Froude Numbers

The power number N_p functions as an analog to the drag coefficient for impellers, quantifying the dimensionless power dissipation due to fluid dynamic drag forces acting on the rotating blades. This interpretation arises because N_p = P / (\rho N^3 D^5) normalizes power input by fluid density \rho, rotational speed N, and impeller diameter D, akin to how drag coefficients scale forces with dynamic pressure and area. The Reynolds number \mathrm{Re} = \rho N D^2 / \mu governs the balance between viscous and inertial forces in the flow around the impeller, dictating the flow regime and thus the variation of N_p. In laminar regimes (\mathrm{Re} < 10), N_p decreases inversely with \mathrm{Re} (slope of -1 on a log-log plot), reflecting viscous-dominated drag; in turbulent regimes (\mathrm{Re} > 10^4), N_p remains constant as inertial forces dominate. The transitional regime ($10 < \mathrm{Re} < 10^4) shows variable dependence, with correlations like N_p = c \mathrm{Re}^x where x ranges from -1 to 0. The Froude number \mathrm{Fr} = N^2 D / g captures the ratio of inertial to gravitational forces, becoming significant in unbaffled tanks where free-surface effects lead to vortex formation and liquid elevation against gravity. In baffled systems, vortex suppression minimizes gravitational influence, rendering N_p independent of \mathrm{Fr}; however, unbaffled configurations require \mathrm{Fr} corrections to account for additional power losses from surface deformation. In unbaffled systems, the combined effects of \mathrm{[Re](/page/Re)} and \mathrm{Fr} manifest through vortex dynamics, where increasing \mathrm{Fr} (higher speeds) deepens the vortex and alters power draw. Correlations adjust N_p as N_{p,\mathrm{cor}} = N_p \cdot \mathrm{Fr}^m, with exponent m varying by impeller type and \mathrm{[Re](/page/Re)}, typically small and positive. Rushton et al. (1950) established the general form N_p = c (\mathrm{[Re](/page/Re)})^x (\mathrm{Fr})^y for such interdependence, with y > 0 prominent at low \mathrm{[Re](/page/Re)} where free-surface flows amplify gravitational impacts. Power curves are conventionally plotted as N_p versus \mathrm{Re}, with \mathrm{Fr} influencing the shape primarily in unbaffled low-\mathrm{Re} regimes, where vortexing reduces effective mixing efficiency and power transfer. For instance, unbaffled turbines exhibit steeper N_p declines with \mathrm{Re} compared to baffled ones due to \mathrm{Fr}-driven solid-body rotation. Beyond mixing tanks, N_p-like groups appear in , where screw propellers operate with N_p \approx 0.5 under open-water conditions, balancing inertial and gravitational forces akin to unbaffled impellers. In axial fans, the dimensionless C_p = P / (\rho N^3 D^5) parallels N_p, correlating with \mathrm{Re} for varying geometries and flow rates.

Usage in Scale-Up Procedures

In scale-up procedures for mixing and processes, the N_p serves as a critical dimensionless for predicting and maintaining consumption across scales, assuming geometric similarity between vessels and operation within the same regime. Specifically, in the turbulent regime where N_p remains constant, scaling at constant per unit volume (P/V) is achieved through the relation P/V \propto \rho N^3 D^2, where \rho is , N is impeller rotational speed, and D is diameter; this allows adjustment of N and D to ensure consistent relative to volume as systems transition from to sizes. Challenges in scale-up often stem from shifts between flow regimes, such as from laminar to turbulent, where N_p is no longer constant but varies inversely with the , potentially leading to inconsistent mixing if not accounted for. In multiphase systems, incorporating effects via N_p-Re-Fr criteria is essential to address and interfacial phenomena that can alter power draw and hydrodynamics during enlargement. Practical guidelines for applying N_p in scale-up vary by process goal; for blending, maintaining constant impeller tip speed (N D) preserves shear and circulation patterns, while for mass transfer-limited operations like gas dispersion or reactions, constant P/V via N_p ensures comparable transfer coefficients and reaction uniformity. Selection of impeller type during scaling must consider N_p values specific to the geometry to align with process demands. The adoption of the in the during the , building on foundational work like Rushton et al.'s correlations for power characteristics, enabled more predictive design and significantly reduced reliance on trial-and-error approaches in scaling mixing operations.

Experimental and Theoretical Considerations

Measurement Techniques

The primary experimental method for determining the power number N_p involves direct measurement of torque on the using dynamometers, which capture the rotational force transmitted to the fluid. These devices, often mounted on the or couplings, provide high-precision for calculating input P via the relation P = 2\pi N T, where N is the impeller rotational speed in revolutions per second and T is the measured . The power number is then computed as N_p = P / (\rho N^3 D^5), with \rho as fluid and D as ; this approach yields accuracies within 1-2% in setups for turbulent regimes. In setups with geared drives or motors, measurements require corrections for mechanical losses, including friction in bearings and gearboxes, which can account for 10-20% of total input power in typical configurations. These adjustments are made by subtracting no-load power draw (measured at zero ) and applying manufacturer-specified efficiencies, ensuring the net power to the fluid is accurately isolated for N_p . An indirect alternative, the calorimetric method, determines power consumption through heat balance in the , particularly useful for non-intrusive measurements in opaque or during active bioprocesses like fermentations. Here, the power input equals the rate of heat accumulation, given by P = m C_p \Delta T / t, where m is the mass, C_p is , \Delta T is the rise, and t is time; minimizes external losses, achieving precisions of 2-5% above 300 rpm. This technique complements torque methods in validating N_p under operational conditions. Modern advancements since the early 2000s incorporate laser Doppler velocimetry (LDV) to measure local velocity fields in stirred tanks, enabling indirect validation of N_p through comparison with computational fluid dynamics (CFD) simulations of power dissipation. LDV provides non-intrusive, point-wise velocity data that informs turbulence models, confirming simulated N_p values within 5-10% of direct measurements for complex impeller geometries and flows.

Limitations and Assumptions

The power number is predicated on several key assumptions regarding behavior and conditions, including that the is Newtonian, the is incompressible, and is isotropic in the fully developed turbulent regime. These assumptions hold primarily for single-phase liquid systems in standard baffled tanks at high Reynolds numbers, where viscous effects are negligible and the remains steady. However, the power number becomes invalid or inaccurate under conditions that violate these assumptions, such as or , where vapor formation or gas introduction disrupts the and alters energy transfer mechanisms. In cavitating flows within agitators, the formation of vapor bubbles leads to deviations from predicted power consumption, often resulting in overprediction of the power number due to reduced effective drag on the . Similarly, aeration introduces gas bubbles that lower the overall power draw by modifying the density and flow resistance, rendering standard power number correlations unreliable without adjustments. Limitations arise in non-standard geometries, where deviations from ideal configurations like centered s and full baffling affect the power number. For instance, increasing the off-bottom clearance beyond one-third of the (D/3) elevates the power number by enhancing recirculation and reducing pumping , with studies showing power consumption rising as clearance varies from 0.17T to 0.41T (T being ). In unbaffled tanks, surface vortexing introduces free-surface effects that necessitate a correction to the power number correlation, as gravitational influences become significant and reduce power input compared to baffled systems. In multiphase systems involving solids or gases, the power number requires modification because the effective density (ρ) of the mixture differs from that of the pure liquid, impacting the dimensionless scaling. For solid-liquid suspensions, particle loading alters the and , leading to errors in power predictions unless an effective ρ is used in the formulation. In gas-liquid systems, such as aerated reactors, the gassed power number (N_{p_g}) typically drops by 20-50% compared to the ungassed value due to and bubble-induced drag reduction, often approximated as N_{p_g} = 0.5 N_p for sparged conditions. Post-2010 (CFD) studies have highlighted further limitations, revealing that the power number primarily captures macro-scale hydrodynamics but is insensitive to micro-scale mixing phenomena, such as and local gradients critical for reactive processes. To address this, the power number is often supplemented by the average energy dissipation rate per unit mass, defined as \epsilon = \frac{P}{\rho V}, where P is input, ρ is , and V is the fluid volume, providing better insight into local mixing efficiency at smaller scales.

References

  1. [1]
    [PDF] Agitation
    Power number. The power number is defined as. Npo = P/(pN³D³). (9.10) where the power P has units of ft lbs or hp or Js". The power number correlates well ...
  2. [2]
    [PDF] Tackling Difficult Mixing Problems - AIChE
    Under turbulent conditions in a baffled tank, the impel- ler power number is essentially constant, and power is proportional to fluid density. Viscosity ...
  3. [3]
    [PDF] MASTERING MIXING FUNDAMENTALS - Webflow
    Power numbers of various impellers are normally compared in the fully turbulent range which typically starts at NRe=103 to 104. Power Number - Reynolds Number ...
  4. [4]
    [PDF] Understand the Real World of Mixing
    The power number is characteristic of the impeller design. Impellers with high Np (typically radial-flow impellers) begin to operate in the turbulent regime ...<|control11|><|separator|>
  5. [5]
    [PDF] Chemineer - Avoid Blending Mix-ups - Brown & Morrison
    extrapolating the power number shown below for the HE-3 and square pitch marine impellers to other impeller-to-tank-diameter ratios. If the impeller-to-tank-.
  6. [6]
    [PDF] EFFECT OF IMPELLER TYPE AND NUMBER AND LIQUID LEVEL ...
    In Equation 1-21 the power number is adjusted to impeller system power number. Equation 1-21 shows that the dimensionless blend time is dependent on impeller ...
  7. [7]
    Power Number - an overview | ScienceDirect Topics
    The power number is defined as a dimensionless number used to scale up agitation tanks, calculated using the formula \( N_p = \frac{P}{\rho N^3 D^5} \), where ...
  8. [8]
    Mixing with helical ribbon agitators: Part II. Newtonian Fluids
    White, A. M., and E., Brenner, “Studies in Agitation,” Trans. Am. Inst. Chem. Engrs., 30, 585 (1934). Google Scholar. Yap, C. Y., “ Mixing of Viscoelastic ...
  9. [9]
    [PDF] POWER CONSUMPTION
    It is desired to find by dimensional analysis a relationship for the power required to maintain the speed of the impeller in terms of the above variables. The ...
  10. [10]
    [PDF] ChE 512
    The flow is a function of the dimensionless quantities listed below and the observed flow patterns A to D are described below. 3 5 power number c. P. Pg. N. n d.
  11. [11]
    Impeller Reynolds Number - an overview | ScienceDirect Topics
    The power number is constant in turbulent systems which depends on the bioreactor and impeller design, whilst in the laminar regime, the P0 of the impeller ...
  12. [12]
    Fluid Flow and Mixing With Bioreactor Scale-Up
    Jun 12, 2024 · The power number for different impeller designs becomes constant once flow is turbulent (1). Under turbulent flow conditions in a stirred tank, ...<|control11|><|separator|>
  13. [13]
    Effect of the Inclination of Baffles on the Power Consumption and ...
    Jun 12, 2017 · Installation of the baffles effectively destroys the circular fluid flows, inhibiting the vortex formation so that the liquid surface becomes ...
  14. [14]
    Mixing Systems: Understanding Impeller Size and Ratios
    Dec 30, 2024 · The D/T ratio, where D is impeller diameter and T is tank width, is ideally 1/3, but most systems use 1/5 to 3/5. Impeller should be 1/3 the  ...
  15. [15]
    Estimation of Gas Holdup Using the Gassed to Ungassed Power ...
    Nov 2, 2020 · This novel correlation is easy to use, offers reasonable precision, and can serve as a valuable alternative to more complex correlation models.
  16. [16]
    [PDF] Mixing: - Impeller Performance in Stirred Tanks - Framatome BHR
    The effect of impeller diameter is now taken into account and large diameter impellers (D/T. ≈ 0.5) are more efficient than smaller ones (D/T ≈ 0.3), pumping ...
  17. [17]
    Pitch Blade Turbine Mixing Impeller - Fusion Fluid Equipment
    Best for low to medium viscosity applications in turbulent and some laminar flow mixing applications. ... POWER NUMBER (NP) = 1.27; FLOW NUMBER (NQ) = 0.79; Flow ...
  18. [18]
    Hydrofoil (PF3) - Fusion Fluid Equipment
    PF3 Hydrofoil Impeller provides excellent pumping ability with a strong axial flow. ... POWER NUMBER (NP) = 0.3; FLOW NUMBER (NQ) = 0.5; Flow Direction: Axial ...
  19. [19]
    (PDF) Power Consumption of Anchor Impeller over Wide Range of ...
    Aug 7, 2025 · Power consumption for an anchor impeller was measured over a wide range of Reynolds number from laminar to turbulent flow regions.
  20. [20]
    [PDF] Power Consumption and Flow Regime Transition in a Stirred Tank ...
    An empirical correlation has been derived using predicted experimental data, which presents a new relation in estimation of power consumption in stirred tanks.
  21. [21]
    [PDF] Studies on the Power Requirement of Mixing Impellers (IV)
    power number (Np) decreases with the increase in Reynolds number (Re), but the rate of decrease falls gradually until finally a constant value is reached ...
  22. [22]
    Effect of impeller clearance on the power number. - ResearchGate
    The downward movement of the discharge streams reduces the vortex activity behind the impeller blades, leading to weaker form drag and a decrease in the power ...
  23. [23]
    The Effect of Impeller and Tank Geometry on Power Number for a ...
    Previous studies of the Rushton turbine have shown that the power number is sensitive to the details of impeller geometry, and in particular to the blade ...Missing: regime | Show results with:regime
  24. [24]
    Impeller power draw during turbulent operation in solid‐liquid ...
    May 28, 2019 · The power draw data is well-described using linear relations between the impeller power number and the density difference correlating parameter ...
  25. [25]
    "Impeller Power Draw in Liquid-Solid Suspensions" - eCommons
    Apr 5, 2017 · In this study, the power draw was characterized in liquid-solid suspensions for a number of impeller types including axial-flow, mixed-flow, and ...
  26. [26]
  27. [27]
    Effect of Reynolds number power number for the different impeller...
    The relationship between dimensionless Power number, Reynolds number, and Froude numbers were found for each impeller type at transitional flow regimes. The ...
  28. [28]
    Marine Prop (PR3) - Mixing Impellers - Fusion Fluid Equipment
    POWER NUMBER (NP) = 0.34 · FLOW NUMBER (NQ) = 0.5 · Flow Direction: Axial · Diameters: 3” to 13” · Bores: ¼” to 1.25” · NOTE: These Power and Flow numbers reflect ...
  29. [29]
    Air System Fans: Engineering Reference — EnergyPlus 8.6
    Fan Efficiency and Shaft Input Power Model: A dimensionless parameter in the form of an Euler number can be used to simplify the description of fan static ...
  30. [30]
    Rules of Thumb: Scale-up - Features - The Chemical Engineer
    Oct 26, 2023 · Ideally, we try to maintain geometric similarity between the equipment at different scales to remove extra uncertainty during scale-up.
  31. [31]
    [PDF] handbook of industrial mixing - science and practice
    ... mixing scale-up fails to produce the. Page 32. INTRODUCTION xxxv required ... power number, and that some fraction of the total energy is dissipated in ...
  32. [32]
    Scale-up in laminar and transient regimes of a multi-stage stirrer, a ...
    This paper presents a scientific approach of the scale-up at constant power per unit of volume in the case of a process in the laminar or transient regime.
  33. [33]
    (PDF) Correlation of Power Consumption for Several Kinds of Mixing ...
    Aug 6, 2025 · The power number (N p ), a dimensionless parameter that relates impeller power consumption to fluid properties and operating conditions, is ...
  34. [34]
    (PDF) The Application of a Strain Gauge Technique to Measurement ...
    Aug 9, 2025 · The work described in this paper establishes the accuracy with which the power number can be measured within a mixing vessel and presents ...
  35. [35]
    Development of a method for reliable power input measurements in ...
    Using different measurement systems, including strain gauges, shaft‐mounted torque ... power number of the upper impeller would be 0.65. This value is ...
  36. [36]
    [PDF] CHARACTERIZING IMPELLER PERFORMANCE IN STIRRED ...
    Apr 25, 2017 · First term in equation is determined by geometry and flow regime: – Reynolds number, pipe roughness. – Pipe length to diameter ratio. – Fittings ...
  37. [37]
    Measurement of Power Consumption in Stirred Vessels—A Review
    Depending upon the principle of measurement, four main techniques have been reported: electric, calorimetric, reaction torque and strain.
  38. [38]
    Power Input Measurements in Stirred Bioreactors at Laboratory Scale
    May 16, 2018 · It is well known that the power number increases as the number of baffles increases, until a critical reinforcement condition is achieved38.<|separator|>
  39. [39]
    [PDF] Comparative Study on Calorimetric Determination of Power ... - Aidic
    The calorimetric method is only accurate at rotational speeds above 300 RPM due to insufficient heat flow at lower settings.
  40. [40]
    Modeling turbulent flow in stirred tanks with CFD - ScienceDirect.com
    ... power and circulation numbers, has been investigated. The results have been validated by laser doppler velocimetry (LDV) data. The stationary and time ...
  41. [41]
    [PDF] Chapter 6 SOLUTION OF VISCOUS-FLOW PROBLEMS
    1. Make reasonable simplifying assumptions. Almost all of the cases treated here will involve steady incompressible flow of a Newtonian fluid in a single.
  42. [42]
    [PDF] Multiphase solid-liquid and solid-liquid-gas mixing in stirred tanks
    May 31, 1989 · The introduction of gas into a liquid in which a mechanical agitator is rotating produces a mixture whose average density is lower and difficult ...
  43. [43]
    [PDF] Effect of Impeller Clearance and Liquid Level on Critical Impeller ...
    Figs. 10 and 11 shows effect of impeller clearance on power consumption for all impellers.
  44. [44]
    (PDF) Bioprocess Engineering Principles - Academia.edu
    ... gassed power number (Ni,)g is: (Nr,). carbon dioxide on the rheological behavior and oxygen =0.5 (Nr,)0. What maximum stirrer speed is now possible transfer ...
  45. [45]
    [PDF] Influence of mixing and heat transfer in process scale-up - DiVA portal
    The microscale is the scale of the molecule, where flow is no longer convective. Micromixing, in this context, refers to mixing at or near the molecular ...<|separator|>
  46. [46]
    Energy Dissipation Rate and Micromixing in a Two-Step Micro ...
    The specific energy dissipation rate ε (W/kg) was calculated by use of the following equation: ε = N ρ V ,, (2). where ρ is the density of solution, kg/m3; V ...