Academic literature on the topic 'Reynolds Range'

Create a spot-on reference in APA, MLA, Chicago, Harvard, and other styles

Select a source type:

Consult the lists of relevant articles, books, theses, conference reports, and other scholarly sources on the topic 'Reynolds Range.'

Next to every source in the list of references, there is an 'Add to bibliography' button. Press on it, and we will generate automatically the bibliographic reference to the chosen work in the citation style you need: APA, MLA, Harvard, Chicago, Vancouver, etc.

You can also download the full text of the academic publication as pdf and read online its abstract whenever available in the metadata.

Journal articles on the topic "Reynolds Range"

1

Horiguchi, Hironori, Daisuke Yumiba, Yoshinobu Tsujimoto, Masaaki Sakagami, and Shigeo Tanaka. "Reynolds Number Effect on Regenerative Pump Performance in Low Reynolds Number Range." International Journal of Fluid Machinery and Systems 1, no. 1 (August 1, 2008): 101–8. http://dx.doi.org/10.5293/ijfms.2008.1.1.101.

Full text
APA, Harvard, Vancouver, ISO, and other styles
2

Castaing, B., and Y. Gagne. "Inertial and dissipative range intermittency at high Reynolds numbers." Physica Scripta T49A (January 1, 1993): 74–76. http://dx.doi.org/10.1088/0031-8949/1993/t49a/011.

Full text
APA, Harvard, Vancouver, ISO, and other styles
3

HORIGUCHI, Hironori, Daisuke YUMIBA, Yoshinobu TSUJIMOTO, Masaaki SAKAGAMI, and Shigeo TANAKA. "Reynolds Number Effect for the Performance of Regenerative Pump in the Range of Low Reynolds Number." Transactions of the Japan Society of Mechanical Engineers Series B 73, no. 735 (2007): 2260–68. http://dx.doi.org/10.1299/kikaib.73.2260.

Full text
APA, Harvard, Vancouver, ISO, and other styles
4

FARAZMAND, M. M., N. K. R. KEVLAHAN, and B. PROTAS. "Controlling the dual cascade of two-dimensional turbulence." Journal of Fluid Mechanics 668 (November 30, 2010): 202–22. http://dx.doi.org/10.1017/s0022112010004635.

Full text
Abstract:
The Kraichnan–Leith–Batchelor (KLB) theory of statistically stationary forced homogeneous isotropic two-dimensional turbulence predicts the existence of two inertial ranges: an energy inertial range with an energy spectrum scaling of k−5/3, and an enstrophy inertial range with an energy spectrum scaling of k−3. However, unlike the analogous Kolmogorov theory for three-dimensional turbulence, the scaling of the enstrophy range in the two-dimensional turbulence seems to be Reynolds-number-dependent: numerical simulations have shown that as Reynolds number tends to infinity, the enstrophy range of the energy spectrum converges to the KLB prediction, i.e. E ~ k−3. The present paper uses a novel optimal control approach to find a forcing that does produce the KLB scaling of the energy spectrum in a moderate Reynolds number flow. We show that the time–space structure of the forcing can significantly alter the scaling of the energy spectrum over inertial ranges. A careful analysis of the optimal forcing suggests that it is unlikely to be realized in nature, or by a simple numerical model.
APA, Harvard, Vancouver, ISO, and other styles
5

Qian, J. "Inertial range and the finite Reynolds number effect of turbulence." Physical Review E 55, no. 1 (January 1, 1997): 337–42. http://dx.doi.org/10.1103/physreve.55.337.

Full text
APA, Harvard, Vancouver, ISO, and other styles
6

Hollenberg, J. W. "Reynolds Number Effects on Regenerative Pump Performance." Journal of Engineering for Industry 109, no. 4 (November 1, 1987): 392–95. http://dx.doi.org/10.1115/1.3187144.

Full text
Abstract:
Reynolds number effects on the performance of a conventional design regenerative pump were investigated, using glycerine-water mixtures, between an impeller tip speed Reynolds number, RT, of 5.0×103 (all glycerine) and 1.6×106 (all water). Results show that the maximum efficiency, nm, can be expressed in terms of an output to loss ratio, nm/1−nm, which varies as RT0.203 for 2.0×104 < RT < 1.6×106 and as RT1.156 for RT < 2.0×104. These results are consistent with efficiency behavior reported in similar investigations on other types of turbomachines. Further, the design point flow coefficient increased over the range of Reynolds number investigated, while the design point head coefficient exhibited a maximum within this range. In addition, marked departure from scaling behavior occurred in the lower Reynolds number range. Finally, the correlation among torque coefficient, head coefficient, and flow coefficient previously established by the author was further verified and followed scaling behavior for the higher Reynolds number range.
APA, Harvard, Vancouver, ISO, and other styles
7

Michna, Jan, and Krzysztof Rogowski. "Numerical Study of the Effect of the Reynolds Number and the Turbulence Intensity on the Performance of the NACA 0018 Airfoil at the Low Reynolds Number Regime." Processes 10, no. 5 (May 18, 2022): 1004. http://dx.doi.org/10.3390/pr10051004.

Full text
Abstract:
In recent years, there has been an increased interest in the old NACA four-digit series when designing wind turbines or small aircraft. One of the airfoils frequently used for this purpose is the NACA 0018 profile. However, since 1933, for over 70 years, almost no new experimental studies of this profile have been carried out to investigate its performance in the regime of small and medium Reynolds numbers as well as for various turbulence parameters. This paper discusses the effect of the Reynolds number and the turbulence intensity on the lift and drag coefficients of the NACA 0018 airfoil under the low Reynolds number regime. The research was carried out for the range of Reynolds numbers from 50,000 to 200,000 and for the range of turbulence intensity on the airfoil from 0.01% to 0.5%. Moreover, the tests were carried out for the range of angles of attack from 0 to 10 degrees. The uncalibrated γ−Reθ transition turbulence model was used for the analysis. Our research has shown that airfoil performance is largely dependent on the Reynolds number and less on the turbulence intensity. For this range of Reynolds numbers, the characteristic of the lift coefficient is not linear and cannot be analyzed using a single aerodynamic derivative as for large Reynolds numbers. The largest differences in both aerodynamic coefficients are observed for the Reynolds number of 50,000.
APA, Harvard, Vancouver, ISO, and other styles
8

Squire, L. C. "The accuracy of flat plate, turbulent skin friction at supersonic speeds." Aeronautical Journal 104, no. 1036 (June 2000): 257–63. http://dx.doi.org/10.1017/s0001924000091570.

Full text
Abstract:
Abstract This paper presents a comparison of turbulent skin-friction coefficients measured by floating element balances in zero pressure gradient conditions for a wide range of Mach numbers and Reynolds numbers. From these comparisons it is clear that although a large number of measurements have been made it is impossible to use these measurements to find the skin-friction coefficient for a given Mach number and Reynolds number to an accuracy of better than ±2 to 3%. This estimate of accuracy applies only to the range of Mach numbers and Reynolds within which there are a reasonable number of measurements to compare. Outside this range, say for Reynolds numbers based on momentum thickness greater than 40,000 and Mach numbers greater than 2 the uncertainty is greater.
APA, Harvard, Vancouver, ISO, and other styles
9

Ramarajan, V., and S. Soundranayagam. "Scale Effects in a Mixed Flow Pump: Part 1." Proceedings of the Institution of Mechanical Engineers, Part A: Power and Process Engineering 200, no. 3 (August 1986): 173–79. http://dx.doi.org/10.1243/pime_proc_1986_200_024_02.

Full text
Abstract:
The variation of efficiency and losses over a range of Reynolds numbers has been measured for a mixed flow pump of specific speed 118 r/min for a number of points covering its operating range. The losses have been separated into those of the impeller and volute. The efficiency is seen to show a steady rise throughout the experimental range in comparison with published results for centrifugal pumps which flatten out at higher Reynolds numbers. A distinct hump is seen in many of the efficiency variation curves as well as in the variation of head with Reynolds number. The hump in the efficiency curves seems to be connected with transition to a roughness dominated regime while that in the head variation appears to be connected to changes in circulation. They both occur at different Reynolds numbers and are unconnected with each other. The frictional component of the losses in the volute is small and the losses there seem to be largely independent of Reynolds number.
APA, Harvard, Vancouver, ISO, and other styles
10

Che Sidik, Nor Azwadi, and Siti Aisyah Razali. "Stability Condition for Single-Relaxation Time Isothermal Lattice Boltzmann Formulation." Applied Mechanics and Materials 695 (November 2014): 667–70. http://dx.doi.org/10.4028/www.scientific.net/amm.695.667.

Full text
Abstract:
In this present research, the Lattice Boltzmann method has been used to determine the stability condition of the single relaxation time. The range of Reynolds number is 100,400 and 1000. Meanwhile, the range of mesh size is varying between 31 to 251. The results show that the increase in both mesh size and Reynolds number give an effect on deviation percentages. The deviation percentages for all mesh and Reynolds number also presented.
APA, Harvard, Vancouver, ISO, and other styles

Dissertations / Theses on the topic "Reynolds Range"

1

Symes, Joseph Alexander. "Dry inclined galloping of smooth circular cables in the critical reynolds number range." Thesis, University of Bristol, 2011. http://ethos.bl.uk/OrderDetails.do?uin=uk.bl.ethos.546204.

Full text
APA, Harvard, Vancouver, ISO, and other styles
2

Sutkowy, Mark Louis Jr. "Relationship between Rotor Wake Structures and Performance Characteristics over a Range of Low-Reynolds Number Conditions." The Ohio State University, 2018. http://rave.ohiolink.edu/etdc/view?acc_num=osu1534768619864476.

Full text
APA, Harvard, Vancouver, ISO, and other styles
3

Bouratsis, Polydefkis. "Scour at the Base of Hydraulic Structures: Monitoring Instrumentation and Physical Investigations Over a Wide Range of Reynolds Numbers." Diss., Virginia Tech, 2015. http://hdl.handle.net/10919/71880.

Full text
Abstract:
Hydraulically induced scour of the streambed at the base of bridge piers is the leading cause of bridge failures. Despite the significant scientific efforts towards the solution of this challenging engineering problem, there are still no reliable tools for the prediction and mitigation of bridge scour. This shortcoming is attributed to the lack of understanding of the physics behind this phenomenon. The experimental studies that attempted the physical investigation of bridge scour in the past have faced two main limitations: i) The characterization of the dynamic interaction between the flow and the evolving bed that is known to drive scour, was not possible due to the limitations in the available instrumentation and the significant experimental difficulties; ii) Most of the existing literature studies are based on the findings of laboratory experiments whose scale is orders of magnitudes smaller compared to bridges in the field, while the scale effects on the scour depth have never been quantified. The objective of this research was to enhance the existing understanding of the phenomenon by tackling the aforementioned experimental challenges. To accomplish this, the first part of this work involved the development of a new underwater photogrammetric technique for the monitoring of evolving sediment beds. This technique is able to obtain very high resolution measurements of evolving beds, thus allowing the characterization of their dynamic properties (i.e. evolving topography and scour rates) and overcoming existing experimental limitations. Secondly, the underwater photogrammetric technique was applied on a bridge scour experiment, of simple geometry, and the dynamic morphological characteristics of the phenomenon were measured. The detailed measurements along with reasonable comparisons with descriptions of the flow, from past studies, were used to provide insight on the interaction between the flow and the bed and describe quantitatively the mechanisms of scour. Finally, the scale effects on scour were studied via the performance of two experiments under near-prototype conditions. In these experiments the effects of the Reynolds number on the flow and the scour were quantified and implications concerning existing small-scale studies were discussed.
Ph. D.
APA, Harvard, Vancouver, ISO, and other styles
4

Howlett, D. P. "Geochronological constraints on Yambah and Chewings-aged deformation at Mt Boothby in the south eastern Reynolds Range, Central Australia." Thesis, 2012. http://hdl.handle.net/2440/92257.

Full text
Abstract:
This item is only available electronically.
Zircon and monazite U–Pb isotope geochronology combined with structural mapping in the Mt Boothby region in the central Aileron Province in Central Australia has constrained the timing of two tectonically distinct phases of high-grade deformation and metamorphism. The first event (D1/M1) occurred at around 1790 Ma and was associated with the emplacement of a bimodal magmatic suite that underwent high-grade deformation prior to the emplacement of voluminous granite also at around 1790 Ma. The timing of D1/M1 coincides with the early stages of the Yambah Event, which is widely recognised in the southern Aileron Province, but has not previously been unequivocally shown to be associated with deformation . Subsequent pervasive reworking occurred over the interval 1600-1570 Ma, and was associated with long-lived granulite-grade metamorphism. The timing of this event coincides with the Chewings Orogeny which largely shaped the tectonic geology further west in the Reynolds and Anmatjira Ranges. During the Chewings Orogeny the c.1790 Ma D1 structures were transposed into a composite S1/S2 fabric. Map scale F2 folding is interpreted to have a shallow plunge suggesting that the S1 fabric may have originally been shallow dipping, raising the possibility that deformation was extensional in nature, and coeval with deposition of the nearby Reynolds Range Group which is constrained to the interval 1806-1785 Ma. Although inferred here to be Yambah aged, the timing constraints for D1 /M1 also overlap with the c. 1800 Ma Stafford Event which was associated with voluminous felsic magmatism, mafic magmatism and extreme geothermal gradient magmatism. This suggests that an extended period of extension, sedimentation, magmatism and deformation may have occurred at around 1800 Ma in the central Aileron Province.
Thesis (B.Sc.(Hons)) -- University of Adelaide, School of Earth and Environmental Sciences, 2012
APA, Harvard, Vancouver, ISO, and other styles
5

Then, M. "Constraints on the origin of early high-heat producing (U-Th enriched) granitic magmatism in central Australia." Thesis, 2016. http://hdl.handle.net/2440/121352.

Full text
Abstract:
This item is only available electronically.
The southern margin of central Australia is characterised by anomalous heat production, 3–5 times higher than global averages. Paleoproterozoic voluminous granitoid complexes in the region are important in the study of this anomalous heat flow. Ca.1800 Ma high-heat producing granites in Mt Boothby have A/NCK (molecular Al2O3/(CaO+Na2O+K2O)) ratios > 1, indicating a predominant origin from partial melting of metasedimentary rocks. The Boothby Orthogneiss is characterised by moderately negative Eu anomalies (Eu/Eu*: 0.03–0.43) and strong depletion in Ba, Rb, Nb and Sr. The enrichment of Ba and Rb relative to Sr and high K2O contents also support a metasedimentary source. The heat production values calculated for the Boothby Orthogneiss and the surrounding Lander formation show that the region is enriched in heat producing elements. The U-Pb zircon age data of inherited zircons in these granites are similar to the detrital zircons of the widespread outcropping; Lander formation. Nd values of -3.5 to 1.3 of the granites infer an evolved crustal source coupled with mixing of a newly mantle-derived component through lower crust assimilation. Zircon saturation temperatures calculated suggest that the Boothby intrusive complex was emplaced at 688–845oC, with a maximum temperature of 776oC, implying an arc environment with associated fluid-flux melting in the mantle wedge, ultimately controlled by subduction dynamics.
Thesis (B.Sc.(Hons)) -- University of Adelaide, School of Physical Sciences, 2016
APA, Harvard, Vancouver, ISO, and other styles
6

Bockmann, K. L. "From Greenschist to Granulite: a mineral equilibria approach to melting and melt loss." Thesis, 2015. http://hdl.handle.net/2440/117961.

Full text
Abstract:
This item is only available electronically.
Melt loss during regional high-grade metamorphism has important consequences for interpreting the metamorphic evolution of the lower crust and for understanding processes leading to the chemical differentiation of the crust. However, melt loss typically modifies the protolith; making it difficult to reconstruct the conditions of prograde metamorphism and the extent to which melt loss modified the rock composition. The Reynolds Range in central Australia preserves a rare example where a single melt-prone stratigraphic unit can be traced from greenschist to granulite grade conditions. Using this as a natural laboratory, P–T mineral equilibria forward models have been calculated to explore melt loss and melt reintegration where both the protolith and the residuum compositions are preserved. Incremental melt loss modelling from the protolith composition along an isobaric heating path at 5 kbar shows that the residual granulite facies rock composition is consistent with around 18% melt loss from the protolith. Large-scale, one-step melt loss from a closed rock system that had built up 18% melt resulted in a similar residual composition to incremental melt loss. The fertility of the open (incremental) system and the closed system showed the closed system produced 5.4% more melt along a heating path from 700–800 °C. Determination of the concentrations of K–U–Th with increasing metamorphic grade shows that K and U concentrations decreased with increasing metamorphic grade. Conversely, Th concentrations increased, resulting in a slight overall increase in heat production from the protolith to the residuum, despite around 18% volume loss associated with melt extraction. An implication for this is that for melt prone rocks such as metapelites, melt loss during granulite facies metamorphism does not deplete the concentration of heat producing elements in the lower crust as is typically assumed.
Thesis (B.Sc.(Hons)) -- University of Adelaide, School of Physical Sciences, 2015
APA, Harvard, Vancouver, ISO, and other styles
7

Weiss, S. "Constraints on the origin of the ca 1780 Ma high heat producing Napperby Gneiss, Aileron Province, Central Australia." Thesis, 2016. http://hdl.handle.net/2440/121355.

Full text
Abstract:
This item is only available electronically.
The Arunta Region of Central Australia contains Paleoproterozoic granites extremely enriched in high heat producing elements, in comparison to a global upper crustal average of 1.69 μWm-3. This study uses geochemistry, geochronology, and zircon saturation thermometry to investigate the source and tectonic environment of emplacement of the ca. 1780 Ma Napperby Gneiss. The Napperby Gneiss is peraluminous, suggesting a metasedimentary source. Samples have negative Eu anomalies ranging from 0.10 to 0.57, and show further evidence of fractionation in negative correlations of Ba and Sr with increasing SiO2. Initial εNd values are similar to surrounding exposed metasedimentary rocks and suggest a strong influence of an evolved crustal source but indicate a necessary juvenile component. Matches of inherited xenocrystic zircons from the gneiss with detrital patterns from the regional metasedimentary Lander Formation indicate that sediments similar to the Lander Formation are the source of the protolith granite. Zircon saturation temperatures suggest the granites were emplaced at 790°C – 872°C. Heat production is less than the slightly older ca 1800 ma suites of the Aileron province, and zircon saturation temperatures are higher. The Napperby was produced by dehydration melting rather than fluid flux melting, possibly in a back arc extensional environment with heat provided by upwelling mantle.
Thesis (B.Sc.(Hons)) -- University of Adelaide, School of Physical Sciences, 2016
APA, Harvard, Vancouver, ISO, and other styles

Books on the topic "Reynolds Range"

1

Pfenninger, Werner. Optimization of natural laminar flow airfoils for high section lift-to-drag ratios in the lower Reynolds number range. Washington, D. C: AIAA, 1989.

Find full text
APA, Harvard, Vancouver, ISO, and other styles
2

Parsons, Chuck. Texas Ranger N.O. Reynolds, the intrepid. Honolulu, HI: Talei Publishers, 2005.

Find full text
APA, Harvard, Vancouver, ISO, and other styles
3

author, Brice Donaly E., ed. Texas Ranger N.O. Reynolds, the Intrepid. Denton, Texas: University of North Texas Press, 2014.

Find full text
APA, Harvard, Vancouver, ISO, and other styles
4

J, Boyle R., and Lewis Research Center, eds. Aerodynamics of a transitioning turbine stator over a range of Reynolds numbers. [Cleveland, Ohio]: National Aeronautics and Space Administration, Lewis Research Center, 1998.

Find full text
APA, Harvard, Vancouver, ISO, and other styles
5

D, Moore Royce, United States. Army Aviation Research and Technology Activity., and United States. National Aeronautics and Space Administration., eds. Performance of two 10-lb/sec centrifugal compressors with different blade and shroud thicknesses operating over a range of Reynolds numbers. [Washington, DC]: National Aeronautics and Space Administration, 1987.

Find full text
APA, Harvard, Vancouver, ISO, and other styles
6

D, Moore Royce, United States. Army Aviation Research and Technology Activity., and United States. National Aeronautics and Space Administration., eds. Performance of two 10-lb/sec centrifugal compressors with different blade and shroud thicknesses operating over a range of Reynolds numbers. [Washington, DC]: National Aeronautics and Space Administration, 1987.

Find full text
APA, Harvard, Vancouver, ISO, and other styles
7

Performance of two 10-lb/sec centrifugal compressors with different blade and shroud thicknesses operating over a range of Reynolds numbers. [Washington, DC]: National Aeronautics and Space Administration, 1987.

Find full text
APA, Harvard, Vancouver, ISO, and other styles
8

R, Whetstone James, and National Institute of Standards and Technology (U.S.), eds. Measurements of coefficients of discharge for concentric flange-tapped square-edged orifice meters in water over the Reynolds number range 600 to 2,700,000. Gaithersburg, MD: U.S. Dept. of Commerce, National Institute of Standards and Technology, 1989.

Find full text
APA, Harvard, Vancouver, ISO, and other styles
9

Measurements of coefficients of discharge for concentric flange-tapped square-edged orifice meters in natural gas over the Reynolds number range 25,000 to 16,000,000. Gaithersburg, MD: U.S. Dept. of Commerce, National Institute of Standards and Technology, 1989.

Find full text
APA, Harvard, Vancouver, ISO, and other styles
10

Biewener, Andrew A., and Shelia N. Patek, eds. Movement in Water. Oxford University Press, 2018. http://dx.doi.org/10.1093/oso/9780198743156.003.0005.

Full text
Abstract:
This chapter examines how the physical properties of water influence and explain the great diversity of swimming performance and mechanisms - from the scale of spermatozoa on up to whales. The key parameters of inertia, viscosity and their manifestation in the critically important Reynolds number are explained and placed in the context of a range of swimming mechanisms, including undulatory movement and fin-based, jet-based, flagellar and ciliary propulsion. The air-water interface also presents an intriguing mechanical challenge for the many organisms that move on top of the water’s surface. The chapter concludes with a brief overview of the burgeoning field of biorobotic swimmers.
APA, Harvard, Vancouver, ISO, and other styles

Book chapters on the topic "Reynolds Range"

1

Danjkov, B. N., E. S. Kornienko, and V. V. Kudrjavtsev. "Supersonic Separation Zone Pressure Fluctuations for Wide Range of Reynolds Number." In Separated Flows and Jets, 237–43. Berlin, Heidelberg: Springer Berlin Heidelberg, 1991. http://dx.doi.org/10.1007/978-3-642-84447-8_34.

Full text
APA, Harvard, Vancouver, ISO, and other styles
2

Čorbo, Tarik, Almin Halač, and Muris Torlak. "Computation of the Fluid Flow Around Octahedral Bodies for a Wide Reynolds Number Range." In Lecture Notes in Networks and Systems, 645–50. Cham: Springer International Publishing, 2021. http://dx.doi.org/10.1007/978-3-030-90055-7_51.

Full text
APA, Harvard, Vancouver, ISO, and other styles
3

König, M., and H. Eckelmann. "An Experimental Study of the Three-Dimensional Structure of the Wake of Circular Cylinders in the Laminar and Transitional Reynolds Number Range." In Bluff-Body Wakes, Dynamics and Instabilities, 341–44. Berlin, Heidelberg: Springer Berlin Heidelberg, 1993. http://dx.doi.org/10.1007/978-3-662-00414-2_73.

Full text
APA, Harvard, Vancouver, ISO, and other styles
4

Bragg, Don C., and Michael G. Shelton. "The Value of Old Forests: Lessons from the Reynolds Research Natural Area." In USDA Forest Service Experimental Forests and Ranges, 61–84. New York, NY: Springer New York, 2014. http://dx.doi.org/10.1007/978-1-4614-1818-4_3.

Full text
APA, Harvard, Vancouver, ISO, and other styles
5

Ishida, Takahiro, Takahiro Tsukahara, and Yasuo Kawaguchi. "DNS of Rotating Turbulent Plane Poiseuille Flow in Low Reynolds- and Rotation-Number Ranges." In Progress in Turbulence V, 177–82. Cham: Springer International Publishing, 2014. http://dx.doi.org/10.1007/978-3-319-01860-7_28.

Full text
APA, Harvard, Vancouver, ISO, and other styles
6

Speziale, Charles G. "Modeling Of Turbulent Transport Equations." In Simulation and Modeling of Turbulent Flows. Oxford University Press, 1996. http://dx.doi.org/10.1093/oso/9780195106435.003.0009.

Full text
Abstract:
The high-Reynolds-number turbulent flows of technological importance contain such a wide range of excited length and time scales that the application of direct or large-eddy simulations is all but impossible for the foreseeable future. Reynolds stress models remain the only viable means for the solution of these complex turbulent flows. It is widely believed that Reynolds stress models are completely ad hoc, having no formal connection with solutions of the full Navier-Stokes equations for turbulent flows. While this belief is largely warranted for the older eddy viscosity models of turbulence, it constitutes a far too pessimistic assessment of the current generation of Reynolds stress closures. It will be shown how secondorder closure models and two-equation models with an anisotropic eddy viscosity can be systematically derived from the Navier-Stokes equations when one overriding assumption is made: the turbulence is locally homogeneous and in equilibrium. A brief review of zero equation models and one equation models based on the Boussinesq eddy viscosity hypothesis will first be provided in order to gain a perspective on the earlier approaches to Reynolds stress modeling. It will, however, be argued that since turbulent flows contain length and time scales that change dramatically from one flow configuration to the next, two-equation models constitute the minimum level of closure that is physically acceptable. Typically, modeled transport equations are solved for the turbulent kinetic energy and dissipation rate from which the turbulent length and time scales are built up; this obviates the need to specify these scales in an ad hoc fashion. While two-equation models represent the minimum acceptable closure, second-order closure models constitute the most complex level of closure that is currently feasible from a computational standpoint. It will be shown how the former models follow from the latter in the equilibrium limit of homogeneous turbulence. However, the two-equation models that are formally consistent with second-order closures have an anisotropic eddy viscosity with strain-dependent coefficients - a feature that most of the commonly used models do not possess.
APA, Harvard, Vancouver, ISO, and other styles
7

B. Dehankar, Prashant. "Assessment of Augmentation Techniques to Intensify Heat Transmission Power." In Heat Exchangers. IntechOpen, 2022. http://dx.doi.org/10.5772/intechopen.101670.

Full text
Abstract:
The heat exchanger detects heat between two processes of liquids in the chemical, petrochemical, food, beverage, and hot metals, and so on. Although the required heat transfer calculations and pressure reductions are achieved with a two-pipeline temperature switch (DPHE), the optimization of the heat transfer parameter is used to measure laboratory test settings. This will allow you to build a DPHE model with twisted tapes and mimic the ASPEN PLUS and work it out by trying to scale the lab that has already been produced and standardized for DPHE. Parameter values for this study range from 0.02 kg / sec −0.033 kg / sec as suspension, pressure reduction, and Reynolds numbers. Also to study the mechanism of increased heat transfer by the use of twisted tape with Y1 = 4.3 and Y2 = 7.7 deviations. They are trying to compare the results of a mathematical model with simulation. This mode of inactivity has the effect of equilibrium heat transfer, pressure drop, the collision factor, and the number Reynolds. We tested the modeling and simulation effects and tried to measure the 4 input parameters of the two output parameters: cold flow rate, hot flow rate, cold and cold temperatures. DPHE, therefore, confirmed the flow rates of weight between 0.02–0.07 kg/s with experiments and simulations performed by Aspen Plus.
APA, Harvard, Vancouver, ISO, and other styles
8

DeHart, Jason D. "Situating Cultural Awareness Through Comics." In Advances in Early Childhood and K-12 Education, 1–14. IGI Global, 2022. http://dx.doi.org/10.4018/978-1-6684-4215-9.ch001.

Full text
Abstract:
This chapter explores the affordances of the comics/graphic novels medium for engaging in critical reading and awareness of relevant cultural issues and a range of experiences. The author situates comics as medium before highlighting the work of Faith Erin Hicks, noting both promise and limitation in the graphic novel series The Nameless City as a way of beginning conversations about relevant topics. The attention then shifts to include the work of Nate Powell, framed as artivism, and the adaptation process of moving the verse novel work of Jason Reynolds to graphic novel form. In this section, the author invites the collaboration of the artist who crafted the adaptation. Finally, attention is given to additional titles that teachers and researchers can consider.
APA, Harvard, Vancouver, ISO, and other styles
9

Chen, Lin. "Principles, Experiments, and Numerical Studies of Supercritical Fluid Natural Circulation System." In Advanced Applications of Supercritical Fluids in Energy Systems, 136–87. IGI Global, 2017. http://dx.doi.org/10.4018/978-1-5225-2047-4.ch005.

Full text
Abstract:
Due to the unique thermal and transport properties, Supercritical natural circulation loop (NCL, or thermosyphon) has been proposed in many energy conversion systems. This chapter presents the principals of supercritical natural circulation loop system and its application challenges. A specially designed experimental prototype system is introduced and compared with numerical findings. The system is operated in wide range of high pressures in the critical region. It is found that in a supercritical NCL system, very high Reynolds number natural convection flow can be achieved only by simple heating and cooling. Thermal performance analysis and parameter effects are carried out along with the experimental development. The heat transfer dependency on operation and its mechanisms are also explained and summarized in this chapter. The comparison of experimental and numerical results contributes to better understanding of NCL stability phenomena and applications in energy systems.
APA, Harvard, Vancouver, ISO, and other styles
10

Chen, Lin. "Principles, Experiments, and Numerical Studies of Supercritical Fluid Natural Circulation System." In Handbook of Research on Advancements in Supercritical Fluids Applications for Sustainable Energy Systems, 219–69. IGI Global, 2021. http://dx.doi.org/10.4018/978-1-7998-5796-9.ch007.

Full text
Abstract:
Due to the unique thermal and transport properties, supercritical natural circulation loop (NCL, or thermosyphon) has been proposed in many energy systems, such as solar heater, nuclear cooling, waste heat recovery, geothermal, etc. This chapter presents the principals of supercritical natural circulation loop and its application challenges. A specially designed experimental prototype system is introduced and compared with numerical findings. The system is operated in wide range of pressures from around 6.0 MPa to 15.0 MPa in the near-critical region. It is found that in a supercritical natural circulation system, very high Reynolds number natural convection flow can be achieved only by simple heating and cooling. Thermal performance analysis and parameter effects are carried out along with the experimental development. The heat transfer dependency on operation and its mechanisms are also explained and summarized in this chapter. The comparison of experimental and numerical results contributes to better understanding of NCL stability phenomena and applications in energy systems.
APA, Harvard, Vancouver, ISO, and other styles

Conference papers on the topic "Reynolds Range"

1

Albrecht, Mitchell B., David A. Olson, Ahmed M. Naguib, and Manoochehr Koochesfahani. "Reynolds Number and Freestream Shear Effects on a NACA 0012 Airfoil in the Reynolds Number Range 10,000-30,000." In AIAA SCITECH 2022 Forum. Reston, Virginia: American Institute of Aeronautics and Astronautics, 2022. http://dx.doi.org/10.2514/6.2022-1055.

Full text
APA, Harvard, Vancouver, ISO, and other styles
2

Smith, Adam E., Stanislav Gordeyev, Theresa Saxton-Fox, and Beverley J. McKeon. "Subsonic Boundary-Layer Wavefront Spectra for a Range of Reynolds Numbers." In 45th AIAA Plasmadynamics and Lasers Conference. Reston, Virginia: American Institute of Aeronautics and Astronautics, 2014. http://dx.doi.org/10.2514/6.2014-2491.

Full text
APA, Harvard, Vancouver, ISO, and other styles
3

Kogiso, Nozomu, Tatsurou Tsushima, and Yoshisada Murotsu. "Wing planform optimization of human powered aircraft in low Reynolds number range." In 8th Symposium on Multidisciplinary Analysis and Optimization. Reston, Virigina: American Institute of Aeronautics and Astronautics, 2000. http://dx.doi.org/10.2514/6.2000-4739.

Full text
APA, Harvard, Vancouver, ISO, and other styles
4

Boyle, R. J., B. L. Lucci, V. G. Verhoff, W. P. Camperchioli, and H. La. "Aerodynamics of a Transitioning Turbine Stator Over a Range of Reynolds Numbers." In ASME 1998 International Gas Turbine and Aeroengine Congress and Exhibition. American Society of Mechanical Engineers, 1998. http://dx.doi.org/10.1115/98-gt-285.

Full text
Abstract:
Midspan aerodynamic measurements for a three vane-four passage linear turbine vane cascade are given. The vane axial chord was 4.45cm. Surface pressures and loss coefficients were measured at exit Mach numbers of 0.3, 0.7, and 0.9. Reynolds number was varied by a factor of six at the two highest Mach numbers, and by a factor often at the lowest Mach number. Measurements were made with and without a turbulence grid. Inlet turbulence intensities were less than 1% and greater than 10%. Length scales were also measured. Pressurized air fed the test section, and exited to a low pressure exhaust system. Maximum inlet pressure was two atmospheres. The minimum inlet pressure for an exit Mach number of 0.9 was one-third of an atmosphere, and at a Mach number of 0.3, the minimum pressure was half this value. The purpose of the test was to provide data for verification of turbine vane aerodynamic analyses, especially at low Reynolds numbers. Predictions obtained using a Navier-Stokes analysis with an algebraic turbulence model are also given.
APA, Harvard, Vancouver, ISO, and other styles
5

BRASSEUR, J. "Evolution characteristics of vortex rings over a wide range of Reynolds numbers." In 4th Joint Fluid Mechanics, Plasma Dynamics and Lasers Conference. Reston, Virigina: American Institute of Aeronautics and Astronautics, 1986. http://dx.doi.org/10.2514/6.1986-1097.

Full text
APA, Harvard, Vancouver, ISO, and other styles
6

Catrakis, Haris, Roberto Aguirre, and Jennifer Nathman. "Large-Reynolds-Number Turbulent Fluid Interfaces and the Upper Range of Scales." In 42nd AIAA Aerospace Sciences Meeting and Exhibit. Reston, Virigina: American Institute of Aeronautics and Astronautics, 2004. http://dx.doi.org/10.2514/6.2004-1113.

Full text
APA, Harvard, Vancouver, ISO, and other styles
7

Morse, Daniel R., and James A. Liburdy. "Vortex Detection and Characterization in Low Reynolds Number Separation." In ASME 2007 International Mechanical Engineering Congress and Exposition. ASMEDC, 2007. http://dx.doi.org/10.1115/imece2007-43011.

Full text
Abstract:
This study focuses on the detection and characterization of vortices in low Reynolds number separation flow over the elliptical leading edge of a flat plate airfoil. Velocity fields were obtained using Time Resolved Particle Image Velocimetry (TRPIV). The Reynolds number based on chord length ranged from 14,700 to 66,700. Experiments were performed for velocities of 1.1, 2.0 and 5.0 m/s and angles of attack of 14°, 16°, 18° and 20°. These velocities correspond to chord length Reynolds numbers of 1.47×104, 2.68×104, and 6.70×104, respectively. A local swirl calculation was used to determine regions of high circulation, and the convection of the centers of these regions was used to determine convective velocities of these vortical structures. The streamwise convective velocity normalized by the freestream velocity is observed to range from approximately 0.4 to 0.65 over the range of angles of attack, with slightly increasing values as the angle of attack increases.
APA, Harvard, Vancouver, ISO, and other styles
8

Chauhan, Kapil, Ivan Marusic, and Nick Hutchins. "Organised motions in turbulent boundary layers over a wide range of Reynolds number." In 6th AIAA Theoretical Fluid Mechanics Conference. Reston, Virigina: American Institute of Aeronautics and Astronautics, 2011. http://dx.doi.org/10.2514/6.2011-3932.

Full text
APA, Harvard, Vancouver, ISO, and other styles
9

Mironov, S. G., V. M. Aniskin, and A. A. Maslov. "Flowing of supersonic underexpanded micro-jets in the range of moderate Reynolds numbers." In PROCEEDINGS OF THE XXV CONFERENCE ON HIGH-ENERGY PROCESSES IN CONDENSED MATTER (HEPCM 2017): Dedicated to the 60th anniversary of the Khristianovich Institute of Theoretical and Applied Mechanics SB RAS. Author(s), 2017. http://dx.doi.org/10.1063/1.5007610.

Full text
APA, Harvard, Vancouver, ISO, and other styles
10

Ogata, Satoshi, Keizo Watanabe, and Yoshihisa Osano. "Drag of a Circular Cylinder in Surfactant Solutions at Intermediate Reynolds Number Range." In ASME/JSME 2003 4th Joint Fluids Summer Engineering Conference. ASMEDC, 2003. http://dx.doi.org/10.1115/fedsm2003-45768.

Full text
Abstract:
To clarify the behavior of the drag coefficient of a circular cylinder in the intermediate Reynolds number range, the flow around a circular cylinder in surfactant solutions was investigated experimentally by measurement of the drag in the Reynolds number range of 3 × 102 to 7 × 103. The experiments were performed in a vertical re-circulating water tunnel. The drag coefficient was measured using an apparatus which could measure the drag acting on the circular cylinder directly. Five cylinders of diameter d = 5, 7, 10, 13 and 20 mm were tested, the ratios of length to diameter (l/d) were 12, 24 and 48. The test surfactant solutions were aqueous solutions of Ethoquad O/12 at concentrations of 50, 100 and 200 ppm, and sodium salicylate was added as a counterion. It was clarified that the drag coefficient of the cylinder in surfactant solutions increased comparing that in tap water in the Reynolds number lower approximately 103 < Re < 3 × 103. According to the increase of the Reynolds number, the drag coefficient decreased. When Reynolds number exceeded approximately 103 < Re < 3 × 103, the drag coefficient in surfactant decreased in comparison with that in tap water finally. In other ward, the drag reduction occurred in this Reynolds number range. The maximum drag reduction was about 55% for 200 ppm solution and 20mm diameter at Re ≅ 7 × 103. The value of the drag coefficient in surfactant solutions was dependent on not only (l/d) but also cylinder diameter. The drag coefficient increased with increasing cylinder diameter. The increase in the concentration of surfactant solution emphasized the characteristics of drag reduction and drag increase.
APA, Harvard, Vancouver, ISO, and other styles

Reports on the topic "Reynolds Range"

1

Whetstone, James R. Measurements of coefficients of discharge for concentric flange-tapped square-edged orifice meters in water over the Reynolds number range 600 to 2,700,000. Gaithersburg, MD: National Bureau of Standards, 1989. http://dx.doi.org/10.6028/nist.tn.1264.

Full text
APA, Harvard, Vancouver, ISO, and other styles
2

Whetstone, James R. Measurements of coefficients of discharge for concentric flange-tapped square-edged orifice meters in natural gas over the Reynolds number range 25,000 to 16,000,000. Gaithersburg, MD: National Bureau of Standards, 1989. http://dx.doi.org/10.6028/nist.tn.1270.

Full text
APA, Harvard, Vancouver, ISO, and other styles
We offer discounts on all premium plans for authors whose works are included in thematic literature selections. Contact us to get a unique promo code!

To the bibliography