Spectra

14
Chapter 4 Turbulence spectra 4.1 Correlations and spectra The autocorrelation of a single variable u(t) at two times t 1 and t 2 is defined as R(t 1 ,t 2 )= u(t 1 )u(t 2 ). (4.1) In the case of a statistically stationary process, the autocorrelation does only depend on the difference τ = t 2 t 1 but not on the actual instants of time, t 1 and t 2 . Instead of (4.1), we could for stationary processes thus also define the autocorrelation as R(τ )= u(t)u(t + τ ). (4.2) With this, we define the normalised autocorrelation function as r(τ )= u(t)u(t + τ )u 2 . (4.3) For any function u(t), it can be shown that u(t 1 )u(t 2 )〉≤〈u 2 (t 1 )1/2 u 2 (t 2 )1/2 , (4.4) (which is the Schwartz inequality) and thus for a stationary process we obtain thus r 1. Figure 4.1 demonstrates how the autocorrelation of a statistically stationary time series can be calculated. A typical autocorrelation plot is shown in figure 4.2. For the time lag τ = 0, the autocorrelation is trivially unity, and it decays with increasing time lag. For the limit of very large time lags the autocorrelation function converges to zero, since turbulence is a random phenomenon. This allows for the definition of an integral time scale T of turbulence: T = 0 r(τ ) dτ, (4.5) see figure 4.2 for a graphical demonstration. T represents the time over which the process u(t) is highly correlated to itself. Let now ˜ u i (t) be the velocity fluctuation vector in a point. Then, more general correlations are given by R ij (τ )= ˜ u i (tu j (t + τ )(4.6) which is the two-point crosscorrelation between u i and u j . For τ = 0 we obtain the one-point cross- correlation which is symmetric and identical to the Reynolds stress R ij (0) = ˜ u i ˜ u j defined in section 3.2. 38

Transcript of Spectra

  • Chapter 4

    Turbulence spectra

    4.1 Correlations and spectra

    The autocorrelation of a single variable u(t) at two times t1 and t2 is defined as

    R(t1, t2) = u(t1)u(t2). (4.1)

    In the case of a statistically stationary process, the autocorrelation does only depend on the difference = t2 t1 but not on the actual instants of time, t1 and t2. Instead of (4.1), we could for stationaryprocesses thus also define the autocorrelation as

    R() = u(t)u(t+ ). (4.2)

    With this, we define the normalised autocorrelation function as

    r() =u(t)u(t+ )

    u2. (4.3)

    For any function u(t), it can be shown that

    u(t1)u(t2) u2(t1)

    1/2u2(t2)1/2, (4.4)

    (which is the Schwartz inequality) and thus for a stationary process we obtain thus r 1. Figure 4.1demonstrates how the autocorrelation of a statistically stationary time series can be calculated. A typicalautocorrelation plot is shown in figure 4.2. For the time lag = 0, the autocorrelation is trivially unity,and it decays with increasing time lag. For the limit of very large time lags the autocorrelation functionconverges to zero, since turbulence is a random phenomenon. This allows for the definition of an integraltime scale T of turbulence:

    T =

    0

    r() d, (4.5)

    see figure 4.2 for a graphical demonstration. T represents the time over which the process u(t) is highlycorrelated to itself.Let now ui(t) be the velocity fluctuation vector in a point. Then, more general correlations are given by

    Rij() = ui(t)uj(t+ ) (4.6)

    which is the two-point crosscorrelation between ui and uj . For = 0 we obtain the one-point cross-correlation which is symmetric and identical to the Reynolds stress Rij(0) = uiuj defined in section3.2.

    38

  • Figure 4.1: Method of calculating the autocorrelation R() = u(t)u(t+ ). This figure has been takenfrom Kundu and Cohen [2002].

    Figure 4.2: Autocorrelation function r() and the integral time scale T . This figure has been taken fromKundu and Cohen [2002].

    39

  • Let Eij() denote the Fourier transform of the correlation function Rij():

    Eij() =1

    2

    eiRij() d (4.7)

    with the frequency . It can be shown that (4.7) is only properly defined if it an be reversed:

    Rij() =

    eiEij() d. (4.8)

    In this case, the equations (4.7) and (4.8) define a Fourier transform pair.For = 0 and i = j, we immediately see from (4.8) that

    u2i =

    Eii() d, (4.9)

    and

    k =

    0

    Eii() d, (4.10)

    with the turbulent kinetic energy from (3.35), i.e. that Eii() d is the turbulent kinetic energy in afrequency band d centered at . Eii() is thus the (three-dimensional) energy spectrum, showing howkinetic energy is distributed as a function of frequency. It is further interesting to calculate Eii(0):

    Eii(0) =1

    2

    Rii() d =u2j

    0

    rii() d =u2j

    T (4.11)

    which shows that the spectrum at zero frequency is proportional to the integral time scale.So far, we have only considered correlations in time. However, correlations in space can be treated inan analogue way. Let ~x0 and ~x0 + ~x be two points in the three-dimensional Cartesian space. Then thespatial cross correlation of a function ui(~x, t) can be formulated as

    Rij(~x) = ui(~x0, t)uj(~x0 + ~x, t). (4.12)

    The energy spectrum tensor Eij( ~K) with the wave number vector ~K can here be formulated in the sameway as in eq. (4.7):

    Eij( ~K) =1

    (2)3

    ei~K~xRij(~x) d~x (4.13)

    where

    Rij(~x) =

    ei~K~xEij( ~K) d ~K. (4.14)

    For i = j and ~x = ~0 we obtain

    Rii(~0) = u2

    i =

    Eii( ~K) d ~K. (4.15)

    For simplicity, the information about the direction of the Fourier modes is often removed by considering

    the length of the wave number vector, K =(K2i

    )1/2instead of the vector itself. This is obtained by

    integrating over the surfaces of spheres S(K) with radius K, such that we can write instead of (4.15):

    40

  • k =1

    2u2i =

    1

    2Eii( ~K) d ~K =

    0

    1

    2Eii( ~K) dS(K) dK. (4.16)

    With defining

    E(K) =

    1

    2Eii( ~K)dS(K), (4.17)

    we obtain

    k =1

    2u2i =

    0

    E(K) dK. (4.18)

    E(K) is thus the scalar energy distribution function per mass unit for a certain wave number value.When investigating turbulent spectra, often the assumption of isotropy is made. This implies that allgradients of averaged quantities vanish and that the Reynolds stress may be expressed as

    uiuj =2

    3kij . (4.19)

    As a consequence of this, the turbulent kinetic energy equation (3.36) reduces to

    tk = . (4.20)

    For homogeneous turbulence, a dynamic equation for the three-dimensional spectrum can be derivedby means of Fourier transforming the dynamic equation for the two-point cross correlation Rij (see eq.(4.12)), which can be seen as an extention of the Reynolds stress equation (3.29), see Hinze [1975] fordetails:

    tE(K) + KT (K) = P(K) 2K2E(K) (4.21)

    where T (K) is the spectral energy transfer rate, i.e. the transport of energy from the lower wave number tothe higher wave numbers (the so-called spectral energy cascade). P(K) is the production of turbulenceat wave number K, which vanishes for isotropic turbulence dominating the smaller scales, such thatturbulence production is dominant on the larger scales. The last term in (4.21) is the spectral formulationof the turbulent dissipation rate.Thus, for isotropic turbulence, we obtain P(K) = 0 and since T (0) = T () = 0, integration of (4.21)over the whole spectrum leads to

    0

    tE(K) dK = t

    0

    E(K) dK = tk = . (4.22)

    With (4.21), we obtain a second formulation for the dissipation rate for the turbulent kinetic energy:

    = 2

    0

    K2E(K) dK. (4.23)

    It was shown in section 3.3 that the pseudo-dissipation is equal to the dissipation rate for homogeneousturbulence, see equation (3.38): = (j ui)

    2. Since on the spatial scales at which the dissipationis most important local isotropy can be assumed (operators are invariant to rotation and reflection),only two types of gradients contribute to the dissipation rate, the longitudinal gradients, e.g. xu and thetransversal gradients, e.g. zu. Since the autocorrelations of these two types are statistically proportional,i.e.

    (xu)

    2(z u)

    2, (4.24)

    41

  • the dissipation rate is also proportional to these correlations. In detail one can show for isotropic turbu-lence:

    1

    15 =

    (xu)

    2

    ,

    2

    15 =

    (z u)

    2

    , (4.25)

    such that with three correlators of longitudinal type, (xu)

    2

    , and six correlators of transverse type,

    (zu)

    2

    , this adds up to (see exercise 5.28 in Pope [2000]).

    In order to show the relation between power and dissipation spectra, we can thus consider the one-dimensional case with homogeneous and isotropic turbulence. Let us define for this case the two-pointcorrelation

    u(x)u(x+ s) = R(s) =

    eiKsE(k)dK, (4.26)

    with the energy spectrum E(K).In the same way, we can define the two point correlation for the gradients of u:

    xu(x)xu(x+ s) = B(s) =

    eiKsD(K)dK. (4.27)

    For this isotropic case, we obtain with (4.24):

    lims0

    B(s), (4.28)

    such that we can interpret D(K) as the dissipation spectrum. We are now interested, how E(K) andD(K) are related to each other. For this, we use the simple equality

    xu(x+ s) = su(x+ s). (4.29)

    Due to homogeneity, we have

    xu(x)xu(x + s) = 0 xu(x)xu(x+ s) = u(x)xxu(x + s). (4.30)

    With this, we obtain

    eiKsD(K)dK = B(s)

    = xu(x)xu(x+ s)

    = u(x)xxu(x + s)

    = u(x)ssu(x+ s)

    = ssu(x)u(x+ s)

    = ssR(s)

    = ss

    eiKsE(K)dK

    =

    eiKsK2E(K)dK,

    (4.31)

    42

  • such that

    D(K) = K2E(K). (4.32)

    The same could have been carried out with zu(z)zu(z+s). Thus, we dissipation spectrum is obtainedfrom the power spectrum by simply multiplying with the square of the wave number.Now, we have a complete picture of the energy cascade process as it is depicted in figure 4.3.In order to calculate Rij(~x), we would need simultaneous observations of u(~x) in a volume of space. Firstof all, this is currently close to impossible in the ocean. It is thus often assumed that the turbulencein a volume of space can be represented by turbulence along a one-dimensional axis x. Also after thissimplification, it is impossible to obtain simultaneous observations with instruments moving vertically orhorizontally through the water (see section 6.1 for observational techniques which do produce simultaneousmeasurements in one or two dimensions). With such instruments, observations of u(x) are approximatedby assuming that the turbulence changes only little during the course of the instrument. Then, the timeseries u(t) may be converted into a simultaneous profile u(x) by setting x = t U0 with U0 denoting theprofiling speed of the instrument. This assumption of frozen turbulence is called Taylors hypothesis.Examples of such profiling systems moving through the water are discussed in section 6.2.

    4.2 Three-dimensional kinetic energy spectra

    Energy spectra for turbulence have been theoretically derived by Kolmogorov [1941]. He assumed that atsufficiently high Reynolds numbers the flow is locally homogeneous and isotropic and to be statisticallyin equilibrium in this range of high wave numbers. Thus, in this range of the spectrum, turbulenceshould uniquely be determined by dissipation rate and viscosity . By means of dimensional analysis,a characteristic length scale for the viscous eddies can then be defined. The involved quantities havethe following units1:

    () = L (4.33)

    () = L2T1 (4.34)

    () = L2T3, (4.35)

    such that we obtain

    =

    (3

    )1/4, (4.36)

    which is called the Kolmogorov microscale at which turbulence dissipates into heat due to the effect ofviscosity.The picture of turbulence is that energy is extracted from the meanflow at the low wave number rangewith wavenumbers of K l1 where l is the integral length scale and K the length of the wave numbervector ~K. The turbulent kinetic energy is then cascading down to higher wave numbers by means of theprocess of vortex stretching due to the non-linear terms in the Navier-Stokes equations. When reachingthe large wave numbers where viscosity dominates, this energy is then dissipated into heat. The range oflength scales much smaller than the integral length scale is called the equilibrium range since there thespectrum is nearly isotropic and generally in equilibrium. This means that the energy in the equilibriumrange is not depending on how much energy is present at the larger scales, but does only depend on theparameters that determine the nature of the small-scale flow, so that we can write

    E = E(K, , ) K l1. (4.37)

    1L represents a length in metres and T a time in seconds.

    43

  • Figure 4.3: Sketch of the energy cascade process. This figure has been taken from the turbulence lecturefrom Schatzmann [1999].

    44

  • The short wave number subrange of the equilibrium range which is not affected by viscosity is called theinertial subrange and can thus be described by

    E = E(K, ) 1 K l1. (4.38)

    Kolmogorov [1941] used dimensional analysis for theoretically deriving the shape of the spectrum in theinertial subrange.The involved quantities have the following units:

    (E) = L3T2 (4.39)

    () = L2T3 (4.40)

    (K) = L1, (4.41)

    the shape of the spectrum must be

    E(K) = K02/3K5/3 1 K l1 (4.42)

    with the Kolmogorov number K0 ranging between 1.4 and 1.8, and depending slightly on the Reynoldsnumber Re. Equation (4.42) is called Kolmogorovs K

    5/3 law.For the viscous subrange dimensional analysis does not provide a solution. The general form for thewhole equilibrium range may be formulated as

    E(K) = K02/3K5/3f(K) K l1 (4.43)

    with the non-dimensional function f . Several formulations have been suggested for f , such as

    f(K) =(1 0.5K0(K)

    4/3)2

    (4.44)

    (see Kovasznay [1948]),

    f(K) =(1 + (1.5K0)

    2(K)4)4/3

    (4.45)

    (see Heisenberg [1948]) or

    f(K) = exp(1.5K0(K)

    4/3)

    (4.46)

    (see Corrsin [1964] and Pao [1965]). These theoretical spectra are shown in figure 4.4. The accordingdissipation spectra are shown in figure 4.5.It should be recalled that the spectra derived here are three-dimensional spectra which are generallydifficult to observe in the ocean. Usually, one-dimensional profiles with stationary remote-sensing instru-ments or actively profiling instruments are observed, see chapter 6. Thus, one-dimensional spectra haveto be considered. This is however only feasable with reasonable effort for the idealised case of isotropy.In isotropic turbulence, no preferred direction exists, such that the off-diagonal terms of the Reynoldsstress tensor vanish, see figure 4.6.

    4.3 One-dimensional kinetic energy spectra

    The argumentation in this section follows mostly the excellent text book by Pope [2000].In practise, oceanic or atmospheric observations of turbulent quantities are mostly along one dimensiononly. It is thus important to understand the properties of the one-dimensional spectra in relation to thoseof fully three-dimensional spectra. This is of course only possible when the turbulence is homogeneousand isotropic on the scales of interest, which in turbulence are generally the small-scales. On the

    45

  • 0.01

    0.1

    1

    10

    100

    1000

    10000

    100000

    0.001 0.01 0.1 1K

    E(K

    )/(

    5)1/4

    Kovasznay [1948] viscous subrange

    Heisenberg [1948] viscous subrange

    Corrsin [1964] & Pao [1965] viscous subrange

    Kolmogorov [1941] five-third law

    Figure 4.4: Theoretical three-dimensional turbulence energy spectra for the equilibrium range.

    46

  • 0.01

    0.1

    1

    0.001 0.01 0.1 1 10K

    (K)2E(K

    )/(

    5)1/4

    Kovasznay [1948] viscous subrange

    Heisenberg [1948] viscous subrange

    Corrsin [1964] & Pao [1965] viscous subrange

    Figure 4.5: Theoretical three-dimensional turbulence dissipation spectra for the equilibrium range.

    Figure 4.6: Isotropic and anisotropic turbulent fields. Each dot represents a uv pair at a certain time.This figure has been taken from Kundu and Cohen [2002].

    47

  • dissipative and the high wavenumber end of the inertial scales, local homogeneity and local isotropy canbe assumed for most geophysically relevant cases.For simplicity the axis along which a one-dimensional profile is considered is the x1 axis. It is thendistinguished between fluctuations of the velocity component paralell to this axis (longitudinal spectra)and those orthogonal to this axis (transverse spectra).A one-dimensional spectrum is defined as:

    Eij(K1) =1

    Rij(~e1x1)eiK1x1dx1. (4.47)

    With this, the longitudinal spectrum is of the following form:

    E11(K1) =1

    R11(~e1x1)eiK1x1dx1 =

    2

    0

    R11(~e1x1) cos(K1x1)dx1, (4.48)

    where the latter identity is due to the fact that R11 is real and symmetric.For the transverse spectrum we define:

    E22(K1) =1

    R22(~e1x1)eiK1x1dx1 =

    2

    0

    R22(~e1x1) cos(K1x1)dx1. (4.49)

    Backtransformation results in

    R11(~e1x1) =

    0

    E11(K1) cos(K1x1)dK1, (4.50)

    and an equivalent result for the transverse correlation, R22(~e1x1). Note that

    R11(~0) = u2

    1 =

    0

    E11(K1)dK1, (4.51)

    and

    R22(~0) = u2

    2 =

    0

    E22(K1)dK1. (4.52)

    In isotropic turbulence, the velocity correlation Rij(~x) can be expressed by the two non-dimensionalsecond order tensors ij and xixj/x

    2 and two scalar multiples, since the only directional information iscontained in the correlation distance ~x:

    Rij(~x) = u2(g(x)ij + (f(x) g(x))

    xixjx

    ), (4.53)

    with x2 = x2k and uiuj = u2ij , such that u

    2

    i = 3u2. For isotropic turbulence, we have u2 =

    u21 = u2

    2 = u2

    3. With setting ~x = ~e1x1, i.e. x2 = x3 = 0 and x1 = x, this can be transformed into

    R11(~e1x1)

    u2= f(x) =

    R11(~e1x1)

    u21

    ,

    R22(~e1x1)

    u2= g(x) =

    R22(~e1x1)

    u22

    ,

    R22(~e1x1) = R33(~e1x1),

    Rij(~e1x1) = 0, for i 6= j.

    (4.54)

    48

  • With the incompressibility condition jRij = 0, we can derive the following relation from (4.53):

    g(x) = f(x) +1

    2xxf(x). (4.55)

    For deriving (4.55), it has been used that

    jx = j(x2

    i )1/2 =

    xix

    (4.56)

    and

    j

    (xixjx2

    )=

    2xix2

    . (4.57)

    With (4.54), the spectra E11(K1) and E22(K1) may be expressed as

    E11(K1) =2

    u2

    1

    0

    f(x1) cos(K1x1) dx1 (4.58)

    and

    E22(K1) =2

    u2

    2

    0

    g(x1) cos(K1x1) dx1. (4.59)

    In the same way as the integral time scale T defined in (4.5), the longitudinal and transverse integrallength scales l11 and l22, respectively, can be defined:

    l11 =

    0

    f(x) dx, (4.60)

    and

    l22 =

    0

    g(x) dx. (4.61)

    The question now is, how the one dimensional spectra are related to the three-dimensional spectrum E(k)from (4.16). For finding this, we need to do some further calculations. Let us first reformulate (4.14) bysetting i = j = 1 and ~x = ~e1x1:

    R11(~e1x1) =

    E11eiK1x1( ~K) dK1 dK2 dK3

    =

    (

    E11( ~K) dK2 dK3

    )eiK1x1 dK1

    =

    0

    E11(K1) cos(K1x1)dK1

    =

    1

    2E11(K1)e

    iK1x1dK1,

    (4.62)

    (where we used (4.50)) such that we can conclude

    E11(K1) = 2

    E11( ~K) dK2 dK3. (4.63)

    In analogy, we conclude

    49

  • E22(K1) = 2

    E22( ~K) dK2 dK3. (4.64)

    We further need to consider the three-dimensional spectrum Eij( ~K) under isotropic conditions. Since for

    that case the spectrum can only depend on the wave number vector ~K itself (see also (4.53)), it must beof the following form:

    Eij( ~K) = A(K)ij +B(K)KiKj . (4.65)

    It can further be generally shown that the incompressibility condition can be spectrally formulated as

    KiEij( ~K) = 0. (4.66)

    Therefore, with (4.65) and (4.66), we obtain

    B(K) = A(K)

    K2, (4.67)

    such that

    Eij( ~K) = A(K)

    (ij

    KiKjK2

    ). (4.68)

    With (4.17), we conclude

    E(K) =

    1

    2Eii( ~K)dS(K)

    =

    1

    2A(K)

    (ii

    KiKiK2

    )dS(K)

    =

    1

    2A(K) (3 1) dS(K)

    = A(K)

    dS(K)

    = A(K)4K2,

    (4.69)

    such that

    A(K) =E(K)

    4K2(4.70)

    and thus

    Eij( ~K) =E(K)

    4K2

    (ij

    KiKjK2

    ). (4.71)

    Therefore, with (4.63), we obtain

    E11(K1) =

    E(K)

    2K2

    (1

    K21K2

    )dK2 dK3. (4.72)

    (4.72) is an integration over the entire K2-K3 plane, for fixed K1. Let us define

    50

  • Kr =K2

    2+K2

    3 K2r = K

    2 K21

    (4.73)

    (with K =K2

    1+K2

    2+K3

    3) as the radial wave number in the K2-K3 plane. Then, since the integrand

    is radially symmetric, (4.72) can be expressed as

    E11(K1) =

    0

    E(K)

    2K2

    (1

    K21K2

    )2KrdKr. (4.74)

    With (4.73), we see that KrdKr = KdK, and K(Kr = 0) = K1 such that (4.74) can be reformulatedinto

    E11(K1) =

    K1

    E(K)

    K

    (1

    K21

    K2

    )dK. (4.75)

    By taking the cosine Fourier transform of (4.55), see (4.58) and (4.59), we obtain for isotropic turbulencewith u21 = u

    22 a relationship between the one-dimensional spectra E11(K1) and E22(K1):

    E22(K1) = E11(K1) +1

    u2

    1

    0

    x1xf(x1) cos(k1x1) dx1. (4.76)

    With general integration by parts with any a and b, we obtain:

    ba

    xxf(x) cos(kx) dx = bf(b) cos(kb) af(a) cos(ka)

    ba

    f(x) cos(kx) dx

    +

    ba

    xkf(x) sin(kx) dx,

    (4.77)

    such that with f(x) 0 for x we obtain from (4.76) the following equation:

    E22(K1) =1

    2E11(K1) +

    1

    u21

    0

    x1K1xf(x1) sin(k1x1) dx1. (4.78)

    On the other hand we directly obtain the following identity from (4.58):

    dE11(K1)

    dK1=

    2

    u2

    1

    0

    x1K1xf(x1) sin(k1x1) dx1, (4.79)

    such that combination of (4.78) with (4.79) leads to

    E22(K1) =1

    2

    (E11(K1)K1

    dE11(K1)

    dK1

    ). (4.80)

    With (4.42) and (4.75), we finally obtain for the longitudinal spectrum

    E11(K1) =18

    55K0

    2/3K5/31

    (4.81)

    and for the transverse spectrum, we obtain from this by means of (4.80):

    E22(K1) =24

    55K0

    2/3K5/31

    . (4.82)

    51