Airfoils and Wings in Compressible Flow

E.L. Houghton , ... Daniel T. Valentine , in Aerodynamics for Engineering Students (Seventh Edition), 2017

Wings of Finite Span

When the component of gratuitous-stream velocity perpendicular to the leading edge is greater than the local speed of sound the wing is said to accept a supersonic leading edge. In this case, as illustrated in Fig. eight.23, at that place is 2-dimensional supersonic flow over much of the fly, which tin can be calculated using supersonic airfoil theory. For the rectangular fly shown in Fig. eight.23, the presence of a wingtip can be communicated only within the Mach cone apex, which is located at the wingtip. The same consideration applies to any inboard three-dimensional furnishings, such as the "kink" at the centerline of a swept-back wing.

Figure 8.23

Figure 8.23. Typical wing with a supersonic leading edge.

In the opposite case, the component of free-stream velocity perpendicular to the leading edge is less than the local speed of audio, and the term subsonic leading edge is used. A typical example is the swept-back wing shown in Fig. 8.24. In this instance, the Mach cone generated by the leading edge of the heart section subtends the whole wing. This implies that the leading edge of the outboard portions of the fly influences the oncoming flow just every bit it does for subsonic flow. Wings having finite thickness and incidence actually generate a daze cone rather than a Mach cone, equally shown in Fig. 8.25. Boosted shocks are generated by other points on the leading border, and the associated shock angles tend to increase because each successive shock moving ridge leads to a reduction in Mach number. These shock waves progressively decelerate the catamenia and then that, at some section such every bit AA′, the flow approaching the leading edge is subsonic. Thus subsonic fly sections are used over most of the wing.

Figure 8.24

Figure eight.24. Example of a wing with a subsonic leading edge.

Figure 8.25

Figure eight.25. A real wing with nonzero thickness creates a more complex shock construction than the zippo-thickness fly in Fig. eight.24 creates.

Wings with subsonic leading edges accept lower moving ridge drag than those with supersonic edges. Consequently, highly swept wings (eastward.grand., slender deltas) are the preferred configuration at supersonic speeds. On the other manus, swept wings with supersonic leading edges tend to have a greater wave drag than practice straight wings.

Read full affiliate

URL:

https://www.sciencedirect.com/scientific discipline/commodity/pii/B9780081001943000080

Mixed convection menstruum in porous medium

John H. Merkin , ... Teodor Grosan , in Similarity Solutions for the Boundary Layer Menstruum and Heat Transfer of Viscous Fluids, Nanofluids, Porous Media, and Micropolar Fluids, 2022

half-dozen.4.2 Basic equations

It is causeless that the free stream velocity and the ambient temperature (far flow from the plate) are U and T , respectively. Information technology is also assumed that the temperature of the plate is T w , where T west > T corresponds to a heating plate (assisting flow) and T w < T corresponds to a cooling plate (opposing flows). Fig. 6.ane shows the physical model and coordinate system with assisting and opposing flows.

Following the mathematical nanofluid model proposed past Tiwari and Das (2007) along with the Boussinesq and boundary layer approximations, it is easy to show that the steady boundary layer equations of the nowadays problem are, run into also (Nield & Bejan, 2017),

(vi.33) u x + 5 y = 0

(6.34) μ n f μ f u = μ n f μ f U + g Thousand [ ϕ ( ρ β ) southward + ( ane ϕ ) ( ρ β ) f ] μ f ( T T )

(6.35) u T x + five T y = k north f ( ρ C p ) n f ii T y 2

field of study to the boundary conditions

(6.36) 5 = 0 , T = T westward on y = 0 u U , T T as y

The physical meaning of the other quantities is mentioned in the Nomenclature . The thermophysical properties of the nanofluid are given in Table 1.two.

We look for a similarity solution of Eqs. (half dozen.33) to (six.35) of the following form:

(half dozen.37) ψ = α f ( 2 P e ten ) 1 2 f ( η ) , θ ( η ) = ( T T ) / ( T due west T ) , η = P east x 1 2 y / ( ten ii )

Substituting variables (half dozen.37) into Eqs. (6.34) and (6.35), information technology is easy to show that nosotros obtain the following ordinary (similarity) differential equation:

(6.38) k n f k f one ϕ + ϕ ( ρ C p ) southward ( ρ C p ) f f + f f = 0

along with the purlieus conditions

(vi.39) f ( 0 ) = 0 , i ( 1 ϕ ) ii.five f ( 0 ) = 1 ( ane ϕ ) 2.5 + [ i ϕ + ϕ ρ s ρ f β south β f ] f ( η ) one as η

Information technology should be noticed that the trouble described past Eq. (6.38) subject to the purlieus conditions (6.39) reduces to that of a regular fluid–porous medium ( φ = 0 ), first studied by Merkin (1980). Then, we betoken out that when λ  =   0 (forced convection flow), the purlieus value problem (6.38, 6.39) has the solution f ( η ) = 0 , then that f ( 0 ) = 0 .

The quantity of applied interest is the skin friction coefficient C f , which tin be easily shown that is given by

(6.xl) ( ii P due east x ) i two C f = 1 ( one ϕ ) ii.five f ( 0 )

Read full affiliate

URL:

https://www.sciencedirect.com/scientific discipline/article/pii/B9780128211885000035

Mixed finite element—smoothed particle methods for nonlinear fluid–solid interactions

Jing Tang Xing , in Fluid-Solid Interaction Dynamics, 2019

Fluid move

The fluid of mass density ρ and the free stream velocity U are assumed to be viscous with their chosen Reynolds number Re = ρ U c / μ = 1100 , k , of which the period is predominantly laminar (Kinsey and Dumas, 2008). The grid organization of the fluid domain is given by Fig. 12.forty, for which Fig. 12.50 gives more details. As sketched in Fig. 12.50B, the aerofoil is placed at a altitude of fourc from the inlet and 12c from outlet, and the width of fluid domain is set to 10c. The dimensions of active meshfree zone around the aerofoil are fix to 1.35c×i.6c. The velocity of the flow particle on the aerofoil surface is required to equal the velocity of aerofoil motion, and the pressure value at the boundary is obtained by solving the pressure Poisson equation based on the Neumann boundary condition, as discussed in Eqs. (12.162a) and (12.162b). A total of 300 nodes are placed on the aerofoil surface, and the computational domain comprises a total of 25,880 meshfree and 65,354 Cartesian nodes. The movement of mesh is accomplished by displacing the meshfree zone co-ordinate to the prescribed pitching motion. The time stride is chosen every bit 0.001   2nd.

Effigy 12.50. Hybrid grid around aerofoil: (A) arrangement of Cartesian grid and meshfree particles, (B) full domain, (C) leading edge, (D) trailing border (Javed, 2015).

Read full affiliate

URL:

https://www.sciencedirect.com/science/article/pii/B9780128193525000124

Vibration Induced by Pressure Waves in Piping

In Flow-induced Vibrations (Second Edition), 2014

five.5.2.1 Cavity tone

Figure 5.48 depicts a two-dimensional flow, with free stream velocity U, over a cavity having length l and depth h. A sound pulse generated by the shear layer impinging at the trailing edge of the crenel propagates upstream and controls the flow separation from the leading edge. The resulting periodic vortex in the shear layer causes a pure tone, which is too called a cavity tone or cavity noise. The pure tone frequency f is expressed by the modified Rossiter equation [157,158]:

Figure 5.48. Flow over a ii-dimensional cavity.

(five.33) f = U l north α M 1 + ( κ ane ) Thousand 2 / 2 + U U c

where north is an integer associated with the number of vortices inside the crenel (mode number or stage number), α an experimental constant, M the free stream Mach number, κ the specific oestrus ratio of the fluid, and U c the convective speed of the vortices. As discussed in Chapter ii, the non-dimensional frequency, known as the Strouhal number, is defined equally:

(5.34) Southward t = f Fifty U

where U scales the variation of velocity in the length scale L.

Since the convective speed of the disturbances in the cavity U c has a significant role, the Strouhal number based on the free stream velocity U and the longitudinal crenel length l is influenced not only by the geometry of the crenel but also by the characteristics of the boundary layer, such as momentum and displacement thicknesses, Reynolds number and Mach number. Figure 5.49 shows the variation of the Strouhal number St=fl/U with the normalized longitudinal crenel length fifty/h in the case of turbulent flow at the separation point and depression Mach number [159]. The Strouhal number increases with increasing cavity length and gradually approaches to a constant value. In Ref. [160] it is reported that for M>0.two, predictions of Eq. (5.33) are consequent with the experimental data when U c/U=0.6 and α=0.25.

Figure 5.49. Variation of organized-oscillation Strouhal number with cavity length for fluid-dynamic oscillations in a two-dimensional cavity [159].

For applied science applications the crenel tone may be generated in three-dimensional cavity geometries every bit shown in Fig. 5.l [158]. In the presence of resonators the pressure level oscillations in the cavity increase in magnitude equally the resonance condition is approached. For a typical geometry the acoustic mode natural frequencies f r are given past:

Figure five.l. Matrix categorization of fluid-dynamic, fluid-resonant, and fluid-elastic types of crenel oscillations [158].

(five.35) f r = ( c / 2 π ) A / Five H for Helmholtz resonator = i c / 4 h ( i = 1 , 3 , 5 , ) for deep cavity }

where c is the speed of sound, A the cross-sectional expanse of the opening, 5 the book within the resonator, H the depth of the opening plus a correction gene equal to approximately 0.8 times the diameter of the opening [161], and h the cavity depth.

The sound pressures when the vortex shedding frequency coincides with the acoustic natural frequency are examined for a Helmholtz resonator and a rectangular deep cavity in Refs. [162,163]. Airtight side branches are treated equally a variation of a deep crenel. The rounding off effect of rectangular side-branch entrances on acoustic resonance is investigated in Ref. [164]. Rounding off the upstream corner of the side branch increases the menstruum velocity required to generate peak acoustic resonance while having a relatively small effect on the magnitude of the peak. The Strouhal numbers, corresponding to the quarter wave resonance frequency based on l+r u (r u =upstream entrance radius) are, however, independent of r u . On the other paw, increasing the downstream radius substantially reduces the maximum acoustic pressure. The frequency at the peak acoustic force per unit area is relatively unaffected by the change in downstream radius. In order to apply the experimental information obtained for rectangular side branches to circular side branches, the rectangular side-branch width l is replaced past the effective diameter d due east=ivl[165].

For pipage systems involving two branches in close proximity, as shown in Fig. five.51, the pressure pulsation amplitudes at resonance increase with a subtract of spacing fifty or the angle θ between the branches [166]. Amplitudes of acoustic pulsations at the resonance of single, tandem, and coaxial arrangements increase in this order. As the diameter ratio of the side branch relative to the chief pipe is increased, the maximum amplitude at resonance increases and the resonance range shifts to college Strouhal numbers [167]. For a side branch installed downstream of an elbow, the critical Strouhal number based on the boilerplate velocity is college when the branch is at the outer side of the elbow as compared to the inner side of the elbow. This is because the local velocity at the outer side of the elbow is higher than boilerplate [167].

Figure five.51. Configurations of closed side branches: (a) unmarried branch, (b) tandem branches, and (c) coaxial branches.

Read total chapter

URL:

https://www.sciencedirect.com/scientific discipline/article/pii/B9780080983479000059

Introduction to convective mass transfer

C. Balaji , ... Sateesh Gedupudi , in Heat Transfer Technology, 2021

12.four The velocity, thermal, and concentration boundary layers

Just as a velocity purlieus layer develops if the surface velocity and the free stream velocity are different and a thermal boundary layer develops if the surface temperature and the gratuitous stream temperature are unlike, a concentration boundary layer develops if the surface concentration and the gratuitous stream concentration of the species are unlike. Fig. 12.ii A–C show the evolution of the velocity, thermal, and concentration boundary layers, respectively, for flow over a flat plate. The species molar concentration at the surface, C A,s , is greater than the gratuitous stream concentration, C A,∞. The concentration boundary layer thickness δ c is normally divers equally the value of y, which satisfies

Figure 12.two. (A) Velocity purlieus layer, (B) thermal boundary layer, and (C) concentration boundary layer.

(12.14) ( C A , due south C A ) ( C A , s C A , ) = 0.99

For a stationary surface, the mass transfer at the surface will be by only improvidence and is given by

(12.15) J A , s = D A B C A y | y = 0

Using Eqs. (12.12) and (12.15), nosotros go

(12.16) h m = D A B C A y | y = 0 ( C A , s C A , )

On mass basis, the equations are

(12.17) j A , s = D A B ρ A y | y = 0

and

(12.xviii) h m = D A B ρ A y | y = 0 ( ρ A , s ρ A , )

Note that the convective mass transfer is due to the combined effect of mass improvidence and advection (bulk fluid motion). If at that place is no flow, then the mass transfer is due to diffusion alone.

Read full chapter

URL:

https://world wide web.sciencedirect.com/scientific discipline/article/pii/B9780128185032000125

Boundary layer equations in fluid dynamics

Hafeez Y. Hafeez , Chifu East. Ndikilar , in Applications of Heat, Mass and Fluid Boundary Layers, 2020

four.12.i Momentum thickness θ and momentum integral

Momentum thickness is the distance that, when multiplied by the square of the gratuitous stream velocity, equals the integral of the momentum defect. Alternatively, the total loss of momentum flux is equivalent to the removal of momentum through a altitude θ. It is a theoretical length scale to quantify the furnishings of fluid viscosity virtually a physical boundary.

A shear layer of unknown thickness grows along the precipitous flat plate in Fig. iv.9. The no-slip wall condition retards the flow, making it into a rounded profile u ( y ) , which merges into the external velocity U which is constant at a "thickness" y = δ ( 10 ) . By utilizing the control volume, nosotros discover (without making any assumptions about laminar versus turbulent period) that the drag strength on the plate is given by the following momentum integral across the go out plane:

Figure 4.9

Effigy 4.ix. Growth of a boundary layer on a flat plate.

(4.99) D ( 10 ) = ρ b 0 δ ( x ) u ( U u ) d y ,

where b is the plate width into the paper and the integration is carried out along a vertical plane 10 = constant .

Eq. (four.99) was derived by Kármán, who wrote it in the user-friendly grade of the momentum thickness θ as

(four.100) D ( x ) = ρ b U 2 θ = ρ b 0 δ ( x ) u ( U u ) d y ,

(4.101) θ = 0 δ u U ( 1 u U ) d y ,

where Eq. (four.101) is called the momentum thickness θ.

Momentum thickness is thus a measure of total plate drag. Drag also equals the integrated wall shear stress along the plate

D ( 10 ) = b 0 x τ west ( x ) d x ,

(4.102) D ( ten ) d ten = b τ w ( x ) .

Meanwhile, the derivative of Eq. (4.100), with U = constant , is

(4.103) D ( 10 ) d ten = ρ b U ii d θ d 10 .

Past comparison this with Eq. (4.102), we arrive at the momentum-integral relation for a apartment-plate boundary-layer flow:

(4.104) τ w ( ten ) = ρ U 2 d θ d x .

It is valid for either laminar or turbulent flat-plate flow.

To get a numerical result for laminar flow, von Kármán assumed that the velocity profiles had an approximately parabolic shape

(4.105) u ( x , y ) U ( 2 y δ y 2 δ ii ) , 0 y δ ( ten ) ,

which makes it possible to estimate both momentum thickness and wall shear

θ = 0 δ ( 2 y δ y 2 δ 2 ) ( i 2 y δ + y two δ 2 ) d y ii 15 δ ,

(4.106) τ w ( x ) = μ u y | y = 0 2 μ U δ .

Past substituting (four.106) into (4.104) and rearranging, we obtain

(4.107) δ d δ 15 ν U d x ,

where ν = μ ρ . We tin integrate from 0 to x, assuming that δ = 0 at x = 0 , the leading edge, to get

1 2 δ 2 = 15 ν x U ,

or

(four.108) δ 10 5.5 ( ν U x ) 1 ii = 5.5 R east 10 1 2 .

This is the desired thickness approximate. It is all approximate, of course, part of von Kármán's momentum-integral theory, merely it is accurate, being only x% college than the known verbal solution for a laminar flat-plate flow, for which we gave δ 10 five.0 R eastward x ane two .

Past combining Eqs. (iv.108) and (four.106), nosotros as well obtain a shear-stress judge along the plate as

(4.109) c f = ii τ w ρ U 2 ( 8 15 R eastward 10 ) 1 two = 0.73 R e x 1 2 .

Again this judge, in spite of the crudeness of the profile assumption (4.105), is only 10% higher than the known exact laminar-plate-menstruation solution c f = 0.664 / R e ten 1 / ii . The dimensionless quantity c f , called the skin-friction coefficient, is coordinating to the friction factor f in ducts.

Read full chapter

URL:

https://www.sciencedirect.com/scientific discipline/article/pii/B9780128179499000128

Grid Turbulence

David S-Yard. Ting , in Basics of Engineering Turbulence, 2016

7.7 Spectral space

At that place are some advantages to recasting the equations of turbulence from real space (defined as space r and time t) into Fourier space (defined every bit moving ridge-number space 1000 and time t). Fourier transform acts like a filter, sorting out or differentiating the different scales present inside a fluctuating indicate.

Turbulence measurements are typically for Eulerian spectra, where fluctuations of randomly oriented eddies are quantified every bit they are swept by a probe at a velocity U. Consequently, simply a i-dimensional slice of the three-dimensional spectrum is measured. The spectral density with frequency n is E 1(n) for the velocity component u. The fraction of u 2 ¯ between n and n +dn is East 1(n)dn and

(7.48) u 2 ¯ = 0 Eastward 1 due north d due north

One problem that arises with expressing spectra in terms of frequency is the fake increase in frequency when the mean velocity U is increased. In other words, Fourier transforming a higher velocity gives a college frequency, but there may or may not be a corresponding increase in the turbulent fluctuation frequencies. Every bit we are interested in frequencies which are directly related to the eddy sizes and non the artifact of changing frequencies associated with variations in sweeping velocity, it is better to limited frequency in terms of the moving ridge-number m 1, where

(7.49) k 1 ii π n / U = two π / wavelength

where subscript "1" is used to distinguish the one-dimensional wave-number from the iii-dimensional wave-number 1000.

We see that

(vii.50) East 1 ( k 1 ) d k 1 = E 1 ( n ) d north

Simply dk 1 = iiπdn/U, and hence

(7.51) E 1 ( k ane ) = Eastward one ( n ) U / ( 2 π )

Therefore

(vii.52) u 2 ¯ = 0 E one k 1 d k 1

Figure 7.8 is a schematic of a spectral space. The dashed line indicates the effects of Reynolds number on the frozen turbulence assumption; whereas for the smallest eddies, some corporeality of dissipation is expected, specially at higher Re.

Figure 7.8. The result of Re in spectral space.

(Created by A. Vasel-Be-Hagh).

As discussed in Chapter 4, at sufficiently high Reynolds number, there exists an inertial sub-range in the turbulence spectrum, which can be represented by a simple ability function of the course

(7.53) Due east ane ( k 1 ) = A i ɛ a one thousand 1 b

where exponents "a" and "b" tin be found via dimensional assay. Note that 1000 1 ≡ 2πn/U has units of g−1 and from Eq. (vii.51), the spectral density

(7.54) E 1 ( k ane ) [ m iii / s 2 ]

With ɛ ∼ [southward2/mthree], we take for Eq. (vii.53)

(7.55) [ m 3 / s 2 ] = [ k 2 / southward 3 ] a [ 1 / thousand ] b

which yields a = 2/3 and b = −five/iii, or

(seven.56) East i ( k 1 ) = A i ɛ 2 / 3 k 1 five / 3

Hinze (1975) institute that for isotropic turbulence, A 1 ≈ 0.56 for the x component of the turbulence velocity.

Notation that the spectrum of the cross-stream components are not the same as E 1(thou 1) of the streamwise component, even in isotropic turbulence. The differences are caused by "aliasing" of the three-dimensional spectrum E(k) of q 2 ¯ / two by the one-dimensional slice taken past a sensor that sees the wave field swept at speed U. According to Hinze (1975), for isotropic turbulence

(vii.57) E two ( k 1 ) = East iii ( k ane ) = ½ [ E ane ( yard one ) m two E 1 ( k 1 ) / k 1 ) ]

In the inertial sub-range, we have

(seven.58) East 2 ( chiliad 1 ) = E 3 ( k 1 ) = 4 A 1 ɛ two / 3 k 1 five / 3 / 3

for isotropic turbulence.

Figure seven.ix is a spectra plot of the OPP turbulence at 10.8 k/south free-stream velocity. We see that the iii spectra corresponding to 20, 60, and 100 hole diameters downstream of the OPP plummet nicely unto each other, which is not the instance for the SHPP turbulence (see Liu and Ting [2007]). For k 1Λ of less than unity, which corresponds to the largest structures, a slight deviation from the self-preservation state is noted. This is somewhat expected, equally the largest structures are expected to acquit some "genetic biases" from the jetwake interactions immediately behind the OPP. Some very high frequency noise is also noted.

Effigy 7.ix. Normalized streamwise turbulence velocity spectrum E ane/(21000Λ/3) at U =  10.eight m/southward downstream of the OPP.

(Created past R. Liu).

Problems

Problem seven.1. The decay of turbulence

The decay of turbulent kinetic energy is typically expressed in the power law form,

kE/mass =A (tt 0)n .

1.

Prove that the smallest value of n is unity, i.east., n ≥ one.

ii.

How does n alter as the turbulence enters into the "final" period of decay?

Trouble 7.two. The anisotropy of grid turbulence

Carry out a refined assay of grid turbulence that accounts for the nonisotropic nature of the turbulence. Assume that v 2 ¯ = w 2 ¯ 0.75 u ii ¯ and derive the ability-constabulary disuse functions for the Taylor microscale λ thou and the integral Λg. Limited the equations in terms of u 2 ¯ rather than q 2 ¯ . Talk over differences between this and isotropic results.

Trouble 7.3. Integral-Taylor scale ratio in isotropic filigree turbulence

Using the usual isotropic results, derive a human relationship for the ratio of macro to microscale Λggrand as a function of distance ten/M. How is the ratio affected by Re =UM/ν; where Chiliad is the mesh size? Compare this to full general scaling results.

Trouble 7.4. Initial versus terminal turbulence decay

A cubical box of book L 3 filled with fluid is shaken to generate a sufficient amount of turbulence and then the turbulence is left to disuse.

1.

Derive an expression for the disuse of the kinetic energy threeu 2/ii every bit a function of time.

2.

When the turbulence decays to Re (= uL/ν) of less than ten, the inviscid estimate ɛ =u three/50 may be replaced by an estimate of the type ɛ =cνu 2/Fifty 2, because the weak eddies remaining at depression Re lose their energy directly to pasty dissipation. Compute c past requiring that the dissipation rate is continuous at uL/ν = 10.

3.

Derive an expression for the decay of the kinetic energy during the final decay period when uL/ν < 10.

4.

If L = one m, ν = i.5 × x−vii 10002/s and u = 1 g/s at time t = 0, how long does information technology take before the turbulence enters the final period of disuse?

Read total chapter

URL:

https://www.sciencedirect.com/science/article/pii/B9780128039700000076

Mixed convection purlieus-layer flow along a vertical flat plate

Ioan Pop , Derek B. Ingham , in Convective Rut Transfer, 2001

Small values of 10 (≪ one)

Virtually the leading border the purlieus-layer is formed mainly by the retardation of the free stream velocity by the effects of viscosity and the issue of the buoyancy strength increases as the boundary-layer develops from the leading edge. This suggests the following transformation:

(2.v) ξ ¯ = g β T * U two ten , η = ( U two v ten ) one two y , f ( ξ ¯ , η ) = ψ ( ii v U 10 ) one two , θ ( ξ ¯ , η ) = T T T *

where T * = T w T . On introducing the transformation (2.five) into Equations (ii.2) and (ii.3) we obtain

(2.vi) three f η 3 + f 2 f η 2 + 2 ξ ¯ ( ± θ + f ξ ¯ 2 f η 2 f η 2 f η ξ ¯ ) = 0

(2.7) one P r 2 θ η two + f θ η + 2 ξ ¯ ( f ξ ¯ θ η f η θ η ξ ¯ ) = 0

and the boundary conditions (ii.4) reduce to

(2.eight) f ( ξ ¯ , 0 ) = 0 , f η ( ξ ¯ , 0 ) = 0 , θ ( ξ ¯ , 0 ) = i for ξ ¯ > 0 f η 1 , θ 0 equally η , ξ ¯ > 0

The method of solution of Equations (2.vi) and (2.7) is to aggrandize the functions f and θ in a serial of small values of ξ ¯ ( one ) of the form

(2.9) f = f 0 ( η ) ± ξ ¯ f 1 ( η ) + ξ ¯ two f ii ( η ) ± θ = θ 0 ( η ) ± ξ ¯ θ i ( η ) + ξ ¯ 2 θ 2 ( η ) ±

where fi and θi (i = 0, i, two) are given by the post-obit sets of ordinary differential equations

(2.10) f 0 ' ' ' + f 0 f 0 ' = 0 , 1 P r θ 0 ' ' + f 0 θ 0 ' = 0 f 0 ( 0 ) = 0 , f 0 ' ( 0 ) = 0 , θ 0 ( 0 ) = ane f 0 ' i , θ 0 0 as η

(2.11) f 1 ' ' ' + f 0 f 1 ' ' 2 f 0 ' f 1 ' + 3 f 0 ' ' f one = 0 , i P r θ 1 ' ' + f 0 θ i ' two f 0 ' ' θ one + 3 f 1 θ 0 ' = 0 f ane ( 0 ) = 0 , f 1 ' ( 0 ) = 0 , θ 1 ( 0 ) = 0 f 1 ' 1 , θ 1 0 equally η

(2.12) f 2 ' ' ' + f 0 f two ' ' 4 f 0 ' f 2 ' + 5 f 0 ' ' f 2 two f 1 ' ii + 2 θ 1 = 0 1 P r θ ii ' ' + f 0 θ 2 ' four f 0 ' ' θ two + three f 1 θ 1 ' 2 f one ' ' θ 1 + 5 θ 0 ' f 2 = 0 f 2 ( 0 ) = 0 , f two ' ( 0 ) = 0 , θ 2 ( 0 ) = 0 f two ' 0 , θ 2 0 as η

These equations were solved numerically by Merkin (1969) for Pr = one and from these results the non-dimensional skin friction and heat transfer on the plate tin can be obtained from the expressions

(2.13) w ( ξ ¯ ) = ( two ξ ¯ ) one 2 two f η 2 ( ξ ¯ , 0 ) , q w ( ξ ¯ ) = ( 2 ξ ¯ ) i 2 θ η ( ξ ¯ , 0 )

Thus we have

(2.fourteen) w ( ξ ¯ ) = ( 2 ξ ¯ ) 1 2 ( 0.4696 ± 1.6216 ξ ¯ 1.2699 ξ ¯ two ± ) q w ( ξ ¯ ) = ( two ξ ¯ ) 1 ii ( 0.4696 ± 0.3834 ξ ¯ 0.6544 ξ ¯ two ± )

for ξ ¯ 1 .

Read full chapter

URL:

https://world wide web.sciencedirect.com/scientific discipline/article/pii/B9780080438788500055

Part i

T. ALBRECHT , ... I. GRANT , in Advances in Wind Engineering, 1988

Two circular cylinders in tandem

Measurements were made of the frequencies of vortex shedding from the two cylinders. At each of the iii upstream cylinder positions the three different cylinders were used in turn, and the measurements were fabricated at iv free-stream velocities ranging from approximately six.7 g s−1 to 21.0 ms−1.

The consequence are tabulated in Tables 1, two and 3, in which

Table 1. Frequencies of vortex shedding with an upstream cylinder of diameter 25.7mm.

ℓ mm Uo ms−i R×10−4 ftwo Hz f1 Hz S
133 6.78 1.02 48.5 48.5 0.184
133 eleven.52 1.73 81.3 81.3 0.181
133 xvi.26 two.44 114.1 114.1 0.180
133 20.98 3.15 144.3, 146.9 144.3, 146.9 0.180, 0.177
184 vi.81 1.02 52.8 52.eight 0.199
184 11.53 i.73 85.8 85.viii 0.191
184 16.28 ii.44 118.9, 119.six 119.two 0.188
184 twenty.99 3.15 153.9 153.9 0.188
254 six.75 i.01 51.6 51.half-dozen 0.196
254 eleven.50 one.73 86.1 86.1 0.193
254 sixteen.25 ii.44 119.two 120.6 0.191
254 21.00 3.fifteen 156.ane 153.9 0.188

Table two. Frequencies of vortex shedding with an upstream cylinder of diameter 24.6mm.

ℓ mm Uo ms−i R×x−four Fii Hz fane Hz S
132 six.77 1.02 fifty.8 50.half dozen 0.184
132 11.50 ane.73 84.eight 84.8 0.181
132 16.26 2.44 118.7 117.7, 118.7 0.178, 0.18
132 21.01 3.15 151.3 151.three 0.177
184 half-dozen.76 1.02 53.four 53.2 0.194
184 11.52 1.73 89.7 89.ii 0.190
184 xvi.27 2.44 124.3 124.3 0.188
184 21.00 3.15 159.ane 158.2, 159.i 0.185, 0.186
254 six.77 i.02 54.four 54.2 0.197
254 11.53 i.73 ninety.ii, 91.1 ninety.2, 91.4 0.192, 0.195
254 16.29 2.44 125.1, 126.8 125.5 0.190
254 21.02 3.15 161.0, 162.3 160.half-dozen 0.188

Table 3. Frequencies of vortex shedding with an upstream cylinder of diameter 23.6mm.

ℓ mm Uo ms−1 R×10−4 f2 Hz fone Hz S
132 6.78 1.02 53.0 53.0 0.185
132 eleven .48 i.72 88.5 89.0 0.183
132 16.xxx 2.44 123.5 124.3 0.180
132 21.06 three.16 157.6, 160.five 160.5 0.180
181 6.74 ane.01 55.3 55.3 0.194
181) 11.47 1.72 92.two 92.0 0.189
184 16.48 ii.47 130.8 130.8 0.187
184 21.08 3.xvi 166.2 166.2 0.186
254 six.76 1.02 56.vii 56.1, 56.7 0.196, 0.198
254 11.49 1.72 93.0, 93.8 94.0 0.193
254 16.27 ii.44 130.9 131 .2 0.190
254 21.0 3.15 166.eight, 168.0 166.eight, 167.6 0.187, 0.188

U0 is the free-stream velocity;

R, the Reynolds number, is U0d2/5, where ν is the kinematic viscosity of air and d2 is the diameter of the downstream cylnder;

f1 and f2 are the frequencies of vortex shedding from the upstream and downstream cylinders respectively;

Southward, the Strouhal number, is f1di/U0, where d1 is the bore of the upstream cylinder;

ℓ is the separation distance betwixt the centres of the cylinders.

In the tables two values for the frequency are given in those cases where two peaks of approximately the same amplitude were present in the spectrum. The peaks probably arise as a effect of the finite length of signal which was recorded.

It can be seen that, except for the final two results in Table 1, the shedding frequencies are, inside the experimental error, equal, any divergence between the two frequencies existence less than 0.75%. The frequency of vortex shedding from the upstream cylinder, therefore, determines that from the downstream cylinder. Too, it may be noted that the values of S increase as the separation increases and tend to the values reported by Westward and Apelt (ref. 4) for a single cylinder, which are 0.202 for R = 1 × 10iv, decreasing to 0.192 for R = three × 104.

Whilst the results are interesting, in that they show that the 2 shedding frequencies are synchronized for different diameters of the upstream cylinder, they are, nevertheless, express. In particular, they do non ascertain the range of upstream cylinder diameters for which the two frequencies are equal. The range must be limited equally, for instance, different frequencies were institute by Baxendale, Grant and Barnes (ref. 5) when the diameter of the upstream cylinder was half that of the downstream cylinder. The investigation was, therefore, extended. However, rather than using a succession of round cylinders having decreasing diameters it was decided to mimic the effect of unlike upstream cylinders by using a flat plate. The shedding frequency could and then exist inverse by simply altering the angle of attack of the plate rather than by laboriously changing the cylinder. Too, the new downstream cylinder allowed one possible ambiguity to exist avoided. In the experiments described above the hot wire downstream of the downstream cylinder measured the fluctuations in the wake from both cylinders. By measuring, in the second gear up of experiments, the force per unit area fluctuation close to the separation indicate on the downstream cylinder there could exist no dubiousness as to the frequency of vortex shedding from that cylinder.

Read full chapter

URL:

https://www.sciencedirect.com/scientific discipline/commodity/pii/B9780444871565500308

The Effect of a Unmarried Roughness Element on a Flat Plate Boundary Layer Transition

1000. Ichimiya , in Engineering Turbulence Modelling and Experiments 4, 1999

three.1 Mean Velocity

Figure 2 testify the wall-normal distributions of spanwise mean velocity components W normalized by the complimentary stream velocity Ue . The meridian from the wally is normalized by the boundaiy layer thickness δ. Within the turbulence wedge, the velocities take large values about the wall and smaller values farther from the wall at both Ten  =   50 and 200. At X  =   50, velocities accept various signs with Z nigh the wall; thus indicating that the spanwise velocity changes its direction oftentimes, although symmetry with respect to the eye of the wedge is adequately well maintained. This suggests that many streamwise vortices exist, and that they are symmetrie with respect to the wedge eye. This will be considered again later with the assist of the vorticity. At 10  =   200 and Z  >   0, most points nearly the wall take positive values, although velocities take negative values at Z  =   -12.5. Likewise, in this streamwise position, symmetry with respect to the middle is well established, but unlike at 10  =   50 the existence of a pair of streamwise vortices is suggested.

Effigy 2. Distribution of the mean velocity component in the streamwise direction

In Fig. 2 distributions of the normal average W, turbulent average WT and not-turbulent average WNorth at the interfaces of the wedge are as well shown. Turbulent and non-turbulent averages are shown in the region of intermittency factors γ ≧ 0.1 and γ ≦ 0.9, respectively. Absolute values of the turbulent average are larger than those of the non-turbulent average except for the region near the wall. From this distribution, the turbulent region tends to overflow the wedge, causing it to expand in a lateral direction.

Effigy 3 show the distributions of wall-normal hateful velocity components V in the spanwise management. At X  =   50 the distribution near the wall at -5≦Z≦5 varies markedly with Z. A comparison of this distribution with the streamwise hateful velocity U / Ueast in the previous paper [two] helps to understand the beliefs of the fluid motion. At the position where U/ Ue is minimal, V / Udue east takes positive values, whereas at the position where U/Ueast is maximal, V/ Ueastward takes negative values, namely, where fluid rises, low-velocity fluid near the wall is elevated and U/Ue is minimal, whereas where fluid falls, loftier-velocity fluid far from the wall is suppressed and U/Ue is maximal. This suggests that in this streamwise position, many streamwise vortices exist, and that they are lifted up or pushed down at their corresponding positions. Outside the wedge, the velocities are virtually zero. Far from the wall, at Y  =   two, the maxima or minima cannot be seen (different distributions near the wall). At X  =   200, like U / Ue and W / Ue , extreme modify are not seen, so information technology may be said that the many streamwise vortices at X  =   50 disappeared. At Z  =   10, values are slightly smaller than the surrounding values, whereas at 16.5 ≦ Z≦18.5 values are greater than the surrounding values. These distributions show that a suppression and meridian of fluid exists at Z  =   10 and 16.5Z≦ 18, respectively. The wall-normal velocity of the suppression and summit is smaller than at 10  =   50 because the differences in values from the surround are smaller. Values outside the wedge are almost nada.

Effigy 3. Distribution of the mean velocity component in the wall-normal direction

As mentioned above, it is suggested that many streamwise vortices exist within the wedge at Ten  =   l, and a pair of streamwise vortices exists at the interfaces between the wedge and outer laminar region at X  =   200. To confirm this, isopleths of the mean streamwise vorticity Ω ten   =   Westward/∂y    V/∂z normalized by the roughness superlative k and the costless stream velocity Ue are shown in Fig. 4. The solid and broken lines in the figure indicate positive and negative values, respectively. The chain lines indicate the boundary layer thickness. The differentials ∂W/∂y and ∂Five/∂z at the middle point betwixt ii points were estimated from slopes betwixt the two points in the distributions of Wand V in y-and z-directions, respectively. At 10  =   fifty, many streamwise vortices be inside the wedge. The sign of the vortices (i.eastward., the direction of rotation of vortices adjacent to each other) is reverse. The sign of the vortices is almost symmetrical with respect to Z  =   0. The extent of isopleth in the wall-normal management varies with Z. At X  =   200, in almost all regions within the wedge, a negative vorticity exists within the wedge, although a highly positive vorticity exists near the wall. In item, at Z  =   xv and around the edge of the boundary layer, the negative vorticity is highest.

Figure iv. Isopleths of the mean vorticity component in the streamwise direction

In this position, which corresponds to the interface between the wedge and the outer laminar region, a streamwise vortex exists which rotates from the outside to the within of the wedge in this position. At X  =   200, as Figs. ii and iii suggest, a positive vorticity exists in the region of Z  <   0, and this vorticity makes a pair with the negative vorticity in Z  >   0. Thus, at Ten  =   200, a pair of streamwise vortices exists at both interfaces, although the vorticity is only about one-tenth of that at X  =   fifty.

Figure 5 show the velocity vectors on the y-z aeroplane. These vectors were obtained from the mean velocities W / Ue and V/Ue in Figs. 2 and three, respectively. At X  =   50, the direction of the vector coincides well with the sense of vortex rotation in Fig. 4. The vectors also confirm the beingness of the many streamwise vortices.

Figure 5. Velocity vectors

Read full chapter

URL:

https://world wide web.sciencedirect.com/science/article/pii/B9780080433288500576