You are on page 1of 19

Large-Scale Turbulence Phenomena in Compound Rectangular Channels

L. Meyer K. Rehme Kernforschungszentrum Karlsruhe GmbH, Institut fiir Neutronenphysik und Reaktortechnik, D-76021 Karlsruhe Federal Republic of German3,
In this article we investigate turbulent flow of air through compound rectangular channels to experimentally investigate the turbulence phenomena in compound channels. Detailed experimental data of axial mean velocity, wall shear stresses, five of six Reynolds stresses, auto- and crossspectral densities, and two-point space correlations were measured by hot-wire anemometry in 18 geometrical configurations. The symmetry of the present flow appears to be better than that of previous measurements and the range of measurements is more extensive. The most interesting result is the existence of a quasi-periodic large-scale turbulence structure in most of the geometries investigated. This structure is stationary and independent of the axial position in the channel. It exists in any longitudinal slot or groove in a wall or a connecting gap between two flow channels, provided its depth is more than approximately twice its width. The frequency of this flow oscillation is determined by the geometry of the slot and is linearly dependent on the bulk velocity. Keywords: Compound channels, gap, slot, turbulent flow, turbulent mixing, hot-wire anemometry, flow pulsation, large-scale cortex, Reynolds stresses, spectral density function, probability density function

INTRODUCTION Unusual distributions of the Reynolds stresses have been reported in the literature for turbulent flow in compound channels. Lyall [1] measured the flow distribution and secondary flow in ducts composed of two square interconnected subchannels (Fig. l a , b). He found a strong momentum transfer from the larger into the smaller subchannel which he attributed to secondary flow. Lyall's experiment aimed at a better understanding of how the flow is distributed between two subchannels with different mean velocities. A strong interest in turbulent flow through compound channels exists for river engineers since such flows occur in overbank flows. Most experiments were performed in open channels. Knight and H a m e d [2] measured the distributions of mean velocity and wall shear stresses in different open channels (Fig. l c ) in which a rectangular channel is b o u n d e d by two rectangular flood plains. They found a strong influence of m o m e n t u m transfer between the subareas on the vertical and lateral distribution of longitudinal velocity. Experimental investigations on the distribution of the velocity and wall shear stress were performed by Knight and Lai [3] in a duct of a fully

symmetric rectangular compound cross section (Fig. ld). A n important feature was found "at the flood p l a i n / m a i n channel interface where the isovels bulge outwards in a manner consistent with secondary flow effects at re-entrant corners." In recent years a considerable number of experimental studies have been undertaken in a straight open channel with a prismatic cross section (Fig. l e ) of the Science and Engineering Research Council Flood Channel Facility ( S E R C - F C F , Shiono and Knight [4-7]). The authors report results on the axial mean velocities and on the Reynolds stresses measured by L a s e r - D o p p l e r A n e m o m e try (LDA). The wall shear stresses have also been measured by Knight et al. [8]. The distributions of the shear stresses in the cross section of the compound channel are found to be highly nonlinear in the regions of the flood p l a i n - m a i n channel interface. The turbulence intensities also show unusual distributions in these regions. In general, the unusual Reynolds stresses and the strong momentum transfer between the compound channels are attributed to secondary flow. Shiono and Knight [4] state that the secondary flow cells contribute some 15 to 30% to the boundary shear stress values. They also measured the secondary flow vectors by L D A and stated that care needs

Address correspondence to Dr. Klaus Rehme, Institut fiir Neutronenphysik und Reaktortechnik, D-76021 Karlsruhe, Federal Republic of Germany.

Experimental Thermaland Fluid Science 1994; 8:286 304


1994 by Elsevier Science Inc., 655 Avenue of the Americas, New York, NY 10010 286 0894-1777/94/$7.00

Large-Scale Turbulence Phenomena to be taken when turbulence models are applied to flow through compound channels since secondary flow affects the dimensionless eddy viscosity [4, 5, 7]. A very interesting result is mentioned by Knight and Shiono [6, 9], namely, that the time records of the velocity and turbulence data show a periodic nature of the longitudinal primary velocity and the transverse velocity. Moreover, the shear stress at the main channel-flood plain interface is dominated by periodic oscillations which could be associated with vertical vortices. The periodic nature of the longitudinal velocity in compound open channels seems to be very similar to the quasi-periodic flow pulsations through the gaps between the rods in axial flow through rod bundles. The high mixing rates between subchannels of rod bundles have also been explained by the effects of secondary flow for a long time, although Rowe [10] in 1973 had noticed that a nongradient macroscopic flow process, possibly flow pulsation, affects the mixing between subchannels. Detailed turbulence measurements of the axial flow through closely spaced rod bundles have confirmed the observations of Rowe and have shown that an energetic and almost periodic flow pulsation exists though the gaps between the rods and between rods and channels walls, respectively ([11, 12]; see our Fig. l f). It was demonstrated that these flow pulsations are the reason for the high mixing rates between the subchannels of rod bundles [13, 14] and that secondary flows in subchannels do not contribute significantly to the mixing rates. The distributions of the turbulence intensities in rod bundles are unusual and different from those in tubes and parallel plates [15-17]. The highest turbulence intensities were observed at positions of greatest distance from the walls in the gap region. In this article we demonstrate that the phenomena mentioned are typical not only for rod bundles but also exist in other compound channels. TEST SECTION AND INSTRUMENTATION Flow Configuration

287

The flow configuration used for this work was a vertical straight duct of rectangular cross section. One short wall was adjustable to change the geometry of the cross section. The channel was constructed from Perspex and had a length of L = 7000 mm. The channel was subdivided by inserts of Perspex into two rectangular subchannels that were connected by a gap (Fig. 2). The dimensions of the overall 18 different geometries are listed in Table 1. The first nine geometries consisted of two almost identical rectangular channels connected by gaps of different sizes. Both the width g and the depth d of the gaps were varied. In case of channel 10 the two rectangular channels had half the cross sections of channel 9. Channel 11 was asymmetric: the cross section of one subchannel was twice the cross section of the other subchannel. The geometries of channels 12 to 18 consisted of only one rectangular channel connected to gaps of the same width g but different depths of the gap, further on called slot. The fluid is air at atmospheric pressure and room temperature. The air is driven by a centrifugal blower. Before entering the working section it passes through a filter to remove particles greater than 1 /zm. A special effort was made to obtain a symmetrical flow distribution at the entrance of the channel. Detailed measurements of the velocity distribution at the entrance were performed before starting the measurement program. Finally, the flow was symmetric to within 1% by using a settling chamber, a honeycomb, different perforated plates, and a number of fine-grid screens.

Measurements

I
I
c

' |
I

Most measurements were performed at a position 30 mm upstream of the open outlet at a Reynolds number of Re = 2.5 105. The length-to-hydraulic diameter ratio of the measuring plane was at least L / D h = 40. The hydraulic diameter is defined for the channel neglecting the

r--q
I
\ /

bI

b1 f

d
J

"

b2

liJ
d

@@@
f g

Z qW ~ V X
Figure 2. Cross section of the flow channel.

Figure 1. Cross sections of experimental investigations in

compound channels; ( a ) + (b) Lyall [1]; (c) Knight and Hamed [2]; (d) Knight and Lai [3]; (e) Shiono and Knight [4]; ( f ) Hooper and Rehme [11]; (g) Wu and Trupp [17].

288

L. Meyer and K. Rehme vided sample and hold digitization with 12-bit resolution. The total number of samples taken in a continuous stream were 96,000 per channel, with a measuring time of 48 s. The raw data were loaded into the extended memory of the computer by direct memory access (DMA). The evaluation of all correlations, including skewness and flatness of the turbulent velocity fluctuations, takes approximately 1 minute. A t each measuring point the probe is turned into two positions to measure the velocity components normal and parallel to the walls. Measurements were taken in the full rectangular channel at 493 to 783 points depending on the size of the cross section of the channel. The measurements of the turbulence intensities and Reynolds stresses were performed with the low-pass filter set at 10 kHz. In addition to the measurement of the Reynolds stresses, power spectral densities of the fluctuating velocities u and w and the cross spectra u w were measured at a number of measuring positions in the gap and in the channel in the vicinity of the gap. Autocorrelation functions were measured for u and w and cross-correlation functions for u w at the same positions as for the spectra. The joint probability distributions between u and v, as well as u and w, respectively, were also determined. F o r these measurements the sampling rate was 1.3 kHz, and the low-pass filters were set to 500 Hz in o r d e r to avoid aliasing. All relevant frequencies were contained in the spectra below 500 Hz. Spatial cross-correlations of axial and transverse velocity components using two x-probes simultaneously were also measured. The measurements were performed at five different flow rates, with a total variation of the velocity by a factor of 2. The spectra and correlations were determined on-line by fast Fourier transform (FFT) with a DT7010 Floating Point Array Processor ( D A T A T R A N S L A T I O N ) using 128 blocks with 1024 data each. With the stationary flow the statistical error is less than 10%. AXIAL MEAN VELOCITY WALL SHEAR S T R E S S E S A N D R E Y N O L D S S T R E S S E S IN TWO CHANNELS W e discuss only some typical results of this investigation in this article. The full data sets of all experiments are available on request. First of all, we will discuss some results for an almost square channel (No. 1) comparable to the results by Melling and Whitelaw [24]. The gap in one wall connecting it to a parallel channel of the same dimensions was only 1.8 mm wide. Next, experimental data are presented for a typical case of two parallel channels connected by a 10-mm gap (No. 9). The discussion of the results will then concentrate on a rectangular channel with a longitudinal slot at one wall with different depths. Finally, we will present some generalized results for all geometries investigated so far. R e c t a n g u l a r C h a n n e l (No. 1) with a n A s p e c t Ratio o f 1.059
L/D

Table 1. Geometries Investigated (all dimensions in mm)


No. a bI
b2

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18

180.0 180.0 180.1 180.1 180.1 180.1 180.1 180.0 180.0 180.0 180.0 180.0 180.0 180.0 180.0 180.0 180.0 180.0

170.0 170.0 170.0 165.9 164.6 154.9 136.5 154.5 136.4 67.75 67.75 --------

169.7 169.7 169.7 164.4 164.7 154.1 136.1 154.5 136.2 67.75 136.4 136.2 136.2 136.2 136.2 136.2 136.2 136.2

9.88 9.88 9.88 20.29 20.30 40.63 76.96 40.60 76.96 76.96 76.96 76.96 68.00 57.80 48.60 36.90 29.00 16.60

1.80 4.95 9.81 2.45 5.05 5.20 5.09 10.20 10.00 10.00 10.01 10.00 10.00 10.00 10.00 10.00 10.00 10.00

gap. The time-mean values of the axial velocity and wall shear stresses were measured by Pitot and Preston tubes (O.D. D = 0.6 mm), respectively. The wall shear stress was evaluated with the correlations by Patel [18]. The turbulent normal and shear stresses were measured by hot-wire anemometry using an x-wire probe. The x-wire probe was a modified Dantec probe with a wire length of 1.2 mm, a diameter of 5 /zm, and a spacing of 0.35 mm. The probes were fabricated in our laboratory. The calibration and evaluation methods use look-up tables as described by Lueptow et al. [19] and are discussed in detail by Meyer [20]. There are several errors typical for x-wire probes, such as the error due to the velocity component normal to the x plane and errors due to the finite length of the wires and the distance between them. These errors also d e p e n d on the velocity gradient [20-22]. Maximum absolute uncertainties for the present results, including the drift of the anemometer, were estimated to be + 3 % for the instantaneous value of U + u and + 0.2 ms-~ for c and w. Since the fluctuating components u, ~', and w are evaluated by the difference between the instantaneous values and the time integrated average U, the uncertainties for the turbulence intensities u' = ~ u 2, e ' = ~/r 2, and w' = V/~-7 are only in the order of + 1%. The error for the turbulent shear stresses uu and u w can go up to + 5 % close to walls. The uncertainty of the velocity measurements was estimated to + 0 . 7 % , and the uncertainty of the wall shear stress data was estimated to + 1.0% by using the procedure of Moffat [23]. The performance of the calibration and of the measurements is fully automated. The mass flow rate and the traversing of the measuring probes are controlled by a microcomputer HP-RS25C. The x-wire probe is run by two CTA-bridges from the Dantec 55M system. A signal conditioner with four channels was used, with filter, offset, and amplification. The signals were digitized at sample rates of 2 kHz per channel by a DT2828-card that pro-

Contours of the axial mean velocity at a position of h = 40.0 are presented in Fig. 3. It is known that the flow is not fully developed at this L / D h ratio [25]. All results are presented for one half-channel only because of

Large-Scale Turbulence Phenomena symmetry. The measured velocities are normalized by the bulk velocity, which is the area weighted average of the velocity measurements, Ub = 20.0 ms -1. The contours were calculated from the data by a computer program without smoothing. The data were measured on a 44-by-47 grid. The symmetry with respect to the symmetry axis through the gap is very good. The symmetry of the mean velocity is better than +0.9% except for the 5% of the cross section close to the walls and better than + 2 % except for the 1% of the cross section close to the walls. The contours show the typical bulging caused by secondary flows except near the wall with the gap. Here, the mean velocities are higher than near the other three walls for the same relative distance from the walls. This effect must be attributed to the flow in a n d / o r through the small gap, respectively. The measured wall shear stresses zw are presented in Fig. 4, normalized by the average wall shear stress Zav along the four walls neglecting the gap. Along the three walls without the gap the distribution of the wall shear stress shows two relative maxima on both sides of the middle of the walls. This distribution is typical for rectangular channels, and the dip in the middle of the walls is caused by secondary flow [26]. The wall shear stress distribution is quite different along the wall with the gap. Close to the gap the wall shear stress has a maximum. The turbulence data were measured by hot-wire anemometry on a 27-by-27 grid. The transverse turbulence intensity contours in the y direction (v') and z direction (w') are shown in Figs. 5 and 6, respectively. The transverse intensities are normalized by the average friction velocity u* = vC-~av/P that was evaluated from the measured wall shear stresses r w and the density of air p to u* = 0.832 ms -1. The experimental data are slightly dif1.4

289

/bottoms,de
I I I I I I I I

.6 1.4

top
.6
I I I

sid'~~e
I I

~
I

~
I

.2

.4

-z/b~

.6

.8

1.4

.6 1.4

I /I_____
gap
.6 J I i

side
i i i
I I I

.2

.4

y/a

.6

.8

Figure 4. Distributions of the normalized wall shear stresses Tw/Ta, , at L / D h = 40 in channel 1.

U/Ub
Figure 3. Contours of the normalized axial velocity U / U b at L / D h = 40 in channel 1.

ferent from the results of Melling and Whitelaw [24] who reported difficulties in measuring these quantities. The contours of u' and w' should be identical if one component is rotated by 90. It is obvious that this is the case, apart from the slight effect on the results near the gap. The data of Melling and Whitelaw show qualitative differences from this behavior. The contours of Figs. 5 and 6, also show a very good symmetry with respect to the symmetry axis through the gap. The contours of the kinetic energy of turbulence are shown in Fig. 7. The kinetic energy k = (1/2)(u 2 + v 2 + 14,'2) is normalized by the square of the average friction velocity. The contours show a good symmetry for the symmetry axis through the gap. The contours are very similar normal to three walls but different at the wall with the gap. The kinetic energy of turbulence is very high close to the gap, with a peak value of k / u *z = 6.1. Except for the region close to the gap, the contours of the kinetic energy are very similar to the

290

L. Meyer and K. Rehme

I
Figure 5. Contours of the normalized transverse intensity u ' / u * a t L / D h = 40 in channel 1.
results of Melling and Whitelaw both in the shape and in the absolute values if the results are c o m p a r e d using the same normalization. Melling and Whitelaw mention that the difference of the squared transverse intensities (w 2 - uZ)/u *~''provides a very sensitive test of symmetry." The contours of this quantity are plotted in Fig. 8. The contours show an excellent symmetry that gives confidence in the measure-

, ,o.

,b,

Figure 7. Contours of the normalized turbulent kinetic energy k / u .2 at L / D h = 40 in channel 1. ment technique and the results of this investigation. The contours of (w 2 - u 2 ) / u .2 should be zero for the diagonal. It can be estimated from the contours that this condition is met with good accuracy. Two R e c t a n g u l a r C h a n n e l s C o n n e c t e d by a 10-mm G a p This channel (No. 9), which is typical for the configurations investigated, consists of two rectangular channels of

Figure 6. Contours of the normalized transverse intensity w ' / u * at L / D h = 40 in channel 1.

Figure 8. Contours of the quantity (w = 40 in channel 1.

U2)//A.2 at L / D h

Large-Scale Turbulence Phenomena 136 180 mm that were connected by a gap 10.0 mm wide and 77 mm deep. The measurements were performed at L / D h = 45.1 on a 23-by-29 grid. Figure 9 presents the contours of the axial velocity normalized by the bulk average that was integrated to U b = 21.5 ms 1. There is excellent symmetry of the contours with respect to the symmetry axis through the gap. The maximum asymmetry is less than _+ 1% except for the 2% of the cross section close to the walls. The isovels in the gap region are quite different from the isovels at the opposite wall. It is surprising that the highest velocities along a parallel to the wall with the gap are not found at the symmetry line through the gap but at a certain distance on both sides of the symmetry line. The bulging of the contour lines toward the corners is much more pronounced than for the almost square channel (Fig. 3). The wall shear stress distribution is displayed in Fig. 10, again, the local data being normalized by the average along the four walls neglecting the shear stress in the gap. Except for the wall with the gap, the distributions are typical for rectangular channels with a dip of the wall shear stress in the middle of the wall. However, the distribution of the wall shear stress along the wall with the gap is significantly different. There is a maximum of Zw on both sides of the gap at about 15 mm from the gap corner; this position is coincident with the strongest gradients of the axial velocity normal to the walls. It is interesting that there is an almost constant gradient of the wall shear stress from the maximum ~'w toward the corners of the channel. The contours of the axial turbulence intensity shown in Fig. l l a display the typical behavior as known from rectangular channels except in the vicinity of the gap. The axial turbulence intensities are normalized by the average 1.4

291

.6 1.4

__ I
top side
I I I I I I I

.6 0

.2

.4 - Z / b ~ 6

.8

1.4

opposite
.6 1.4
I I I

gap
I I

f
.6 0
I I

gap
I

side
I I

.2

.4

y/a

.6

.8

.6 gap
.2 , I o i i I

.5

.4

.3

z/d

.2

0.1

Figure 10. Distributions of the normalized wall shear stresses ~'w/%v at L / D h = 45.1 in channel 9. Origin of z coordinate at the edge of the gap. friction velocity, which was evaluated from the measured wall shear stresses to u* -- 0.924 ms -]. The symmetry of the contours with respect to the symmetry line through the gap is good. Toward the corners, the contours show the typical bulging of low-turbulence intensity fluid transported from the channel center by secondary flow. The

U/U b Figure 9. Contours of the normalized axial velocity U / U b at


L / D a = 45.1 in channel 9.

292

L. Meyer and K. Rehme friction velocity are displayed as contour lines in Fig. 12a. The contours are fairly symmetric with respect to the symmetry line through the gap. The contours are similar to those of a rectangular channel except along the wall with the gap. As was observed for the axial intensity, the transverse intensity in the y direction is also very high in

contours are completely different in the vicinity of the gap. Very high axial intensities are observed close to the edges of the gap with two peaks as high as u ' / u * = 3.4. The two peaks are clearly visible in the three-dimensional (3D) representation (Fig. llb). The turbulence intensities in the y-direction, v' (parallel to the wall with the gap), normalized by the average

))/ d: ,
U~/U *
a

/!
Vt/U *
a

V'/LI* Lll/U
b

Figure 11. (a) Contours of the normalized axial intensity u'/u* at L / D h = 45.1 in channel 9; (b) 3D plot of the normalized axial intensity u'/u* at L / D h = 45.1 in channel 9.

Figure 12. (a) Contours of the normalized transverse intensity v'/u* at L / D h = 45.1 in channel 9; (b) 3D plot of the normalized transverse intensity v'/u* at L / D h = 45.1 in channel 9.

Large-Scale Turbulence Phenomena the vicinity of the gap. The intensity v' has two peaks as high as v ' / u * . ~ 2.0, whereas the maximum value is v ' / u * = 1.4 at the opposite wall. The 3D representation (Fig. 12b) clearly shows the two peaks close to the gap. The transverse intensity in z direction, w' (parallel to the gap), is shown in Fig. 13a as contours. The intensity is normalized by the average friction velocity. The transverse intensity w' also has very high values close to the gap. There is a peak as high as w ' / u * = 2.0 in contrast to a

293

value of w ' / u * = 0.91 near the opposite wall. There is one peak directly at the edge of the gap on the symmetry line, not two peaks, as for the other intensities u' and v'. The 3D plot of the transverse intensity w' shows this definitely (Fig. 13b). Moreover, Fig. 13b also shows that the transverse intensity w' is very high along the whole depth of the gap. The symmetry of the contours with respect to the symmetry line through the gap is good. In the three regions of the walls without a gap the contours are similar to those found in rectangular ducts. The transverse intensities v' and w' should be images of each other when reflected about a diagonal. Figures 12a and 13a show that this is rather true if one takes into account that the lengths of the channel walls are different. Of course the kinetic energy of turbulence k = (1/2) (u 2 + v 2 + w 2) displays the features found for the intensities. Figure 14 is a contour plot of the kinetic energy normalized by the square of the average friction velocity. The symmetry of the contours with respect to the symmetry line through the gap is good since the symmetry of the three intensities was good. In the region close to the gap, the kinetic energy of turbulence is very high and shows two peaks at the corners of the gap. The peak value is k / u .2 = 9.2, which is more than double the highest values at the other walls without a gap. In the regions near the walls without a gap the contours are very similar to each other and to those found in rectangular channels. The contours of the kinetic energy bulge toward the corners of the channel, which means relatively low energy in the corners. The Reynolds shear stresses uv and uw are presented in Figs. 15a and 16a as contour lines. The shear stresses are normalized by the square of the average friction velocity. Except for the region near the gap, the contours are

Ww/U*
a

WP/U *
Figure 13. (a) Contours of the normalized transverse intensity w'/u* at L / D h = 45.1 in channel 9; (b) 3D plot of the normalized transverse intensity w'/u* at L / D h = 45.1 in channel 9. k/u*2 Figure 14. Contours of the normalized turbulent kinetic energy k / u .2 at L / D h = 45.1 in channel 9.

294

L. Meyer and K. Rehme

UV/U*
a

2
a 5

UW/U

.2

uw/u*e
b

uv/u .2
Figure l_ft. (a) Contours of the normalized turbulent shear stress u v / u .2 at L / D h = 45.1 in channel 9; (b) 3D plot of the normalized turbulent shear stress ~ / u .2 at L / D h = 45.1 in channel 9. similar to the contours comput_ed by H u r et al. [27]. In the gap region, the shear stress uv has two peaks, a positive and a negativ__2e one, with ~ / u * 2 ~ +_2.4, whereas, the shear stress uw shows only one peak with ~ w / u .2 -~ 4.4.

Figure 16. (a) Contours of the normalized turbulent shear stress ~ w / u .2 at L / D h = 45.1 in channel 9; (b) 3D plot of the normalized turbulent shear stress -~w/u .2 at L / D h = 45.1 in channel 9.

This shows that the shear stresses close to the gap are much higher than anywhere else in the channel. The 3D plots (Figs. 15b and 16b) clearly demonstrate the unusually high shear stresses close to the gap. The perspective in Fig. 15b has been changed__to better demonstrate the behavior of the shear stress uu near the gap. The shear stress uw parallel to the gap is zero at the symmetry line between the two channels at the center of the gap. The

Large-Scale Turbulence Phenomena Reynolds stresses measured in the open water channel by Shiono and Knight [7] show a very similar distribution. The quantity (w 2 - u2)/u .2 is shown in Fig. 17a for completeness. The contours indicate a reasonable symmetry with respect to the axis through the gap. The values of (w 2 - u2)/u .2 are rather high in the neighborhood of the gap. In the regions of the other three walls the contours are similar in shape and magnitude to those of the almost square channel (Fig. 8). In the gap the values of (w e - v 2 ) / u .2 are on a high level, which is clearly displayed by the 3D plot (Fig. 17b). F R E Q U E N C Y ANALYSIS Channel with a Gap

295

A detailed analysis of the frequency spectrum and of the correlations of the velocity components was performed at the symmetry line within the gap and in the vicinity of the gap in the flow channel at five different flow velocities.
Auto- and Cross-Power Spectral Density Function Figure 18a shows the autospectral power density function ~w of the w-component at four positions in the gap 10.0 mm wide (g) and 77 mm long (d) connecting two rectangular channels (No. 9) at a bulk velocity of Ub = 21.5 m s - . Curves 2, 3, and 4 were shifted by one order of magnitude each. All curves have a pronounced peak at 68 Hz. Except for the position 1 at the center of the gap there is a smaller peak approximately at twice that frequency at 136 Hz. The main peak is highest at the center and decreases toward the edge of the gap. But even 10 mm outside of the gap (position 4) a pronounced peak is still visible. The power spectrum ~u of the axial velocity component u in Fig. 18b shows peaks at the same frequencies. However, at the center of the gap the main peak at 68 Hz is missing. Also the cross-spectral power density ~ w shows the main peak and the smaller one at the same frequencies (Fig. 18c). The absolute height of the cross-power spectrum is very small at position 1 in the center of the gap, which leads to relatively strong variations in the curve. These spectra give an explanation for the high values of the turbulent intensities and shear stress values in and near the gap that we saw in the preceding figures. It is a turbulent motion with frequencies at certain low values rather than a generally higher level of turbulence. Auto- and Cross-Correlation These characteristics of a large-scale turbulent fluctuation also can be seen in the autocorrelation and cross-correlation functions in Fig. 19. The autocorrelation functions of w and u clearly show the dominant frequency of 68 Hz, taking the inverse of the times ~- at which the correlation functions have their second maximum. The least damping of the correlation function for the transverse velocity component w occurs at position 1 while the correlation of the axial component is close to zero at those times ~'. The cross-correlation functions Ruw shows a delay between the signals, but the period of the functions is of the same length, independent of the position and equal to that of the autocorrelation function. There is a position-dependent phase shift between u and w.

b Figure 17. (a) Contours of the quantity (w 2 U2)//U.2 at = 45.1 in channel 9; (b) 3D plot of the quantity (w 2 - c 2 ) / u *z a t L I D h = 45.1 in channel 9.
-

LID h

Probability Density Function For a better characterization of the periodic large-scale fluctuations we determined the single and joint probability density functions of u and w. In Fig. 20, the probability density function of the transverse velocity in the center of the gap is symmetric with maxima at + 2.2 ms -1, which is characteristic for a sine wave plus random noise. Moving toward the edge of the gap (position 3), the maximum for the positive velocity component (directed into the gap) shifts to smaller velocities, while the negative components become less concentrated around a specific velocity.

296

L. Meyer and K. Rehme

' '4 . . . . . I _ ^'

' '~)~

'l

.....

10-1

10 -I

0
10 -2

0
1

0 -2

10 -a

10 -3

10 -4
68 10 2 4 68
100

10 -4 2 4 6

I Illi

I I Ill

4 68 10

4 68 I00

4 6

frequency

(Hz)

frequency
b

(Hz)

'

10 -I

e
1 0 -~

10 -3

10 -4
468 10

4613
100

frequency

(Hz)

Figure 18. Auto- and cross-power spectral density functions measured at four positions on the symmetry line through the gap No. 9; (a) ~w of velocity component w parallel to the gap, (b) qb of axial velocity u, (c) cross-spectral density qbuw of u and w; curve 2, 3, and 4 are shifted up by one order of magnitude each. pos. 1, z/d = 0; pos. 2, z/d = -0.25, pos. 3, z/d = - 0 . 5 , pos. 4, z/d = -0.63. Flow conditions were as listed in Table 2. In Table 2 some data measured at the four positions are listed. Consistent with the functions in Fig. 20 the skewness of w is close to zero at the center (position 1) and has its maximum at position 3, the edge of the gap. The skewness of the axial component is largest at position 2 and with opposite sign at position 4, outside the gap. The flatness varies in a large range, being small for w and large for u at the center and vice versa at the edge of the gap. The joint probability density distribution in Fig. 21 shows a shape very unfamiliar in turbulence research at the center of the gap (1), but it is consistent with all

Large-Scale Turbulence Phenomena .4


i\
Rw~

297

I //

--

.3 o ._,.2
v

.1 -1
a

Sr ~

~-"

0
I I I I I I

-6
I I I I I I

-4

-2

'

'

u (m/s)
I I I I I /

REu

.5

---

1 2 a

.4
--I I I I I I I I

---

3
4

/~
iI A

0.006

0.016 b
"I .'~

0.024

0.032
O( .2

/ I /

T - - T - - T - -

.I
Ruw _s dP

'/A
_..~s
II

0 -

J/

-6

' -'4

' -'2

'

w (m/s)
b -0.016 -0.008 0 0.006 0.016 Figure 20. Probability density function of (a) the axial velocity component u, and (b) the transverse velocity component w at four positions in the gap No. 9. For captions see Fig. 18.

(s)
Figure 19. Auto- and cross-correlation functions measured at four positions on the symmetry line through the gap No. 9. (a) Rww of velocity component w parallel to the gap; (b) Ru, of axial velocity u; (c) cross-correlation R,w. For captions see Fig. 18. presented data. Outside the gap (4) it resembles a distorted distribution close to a wall with a positive value of
UU.

was taken with two x-probes simultaneously at a slightly higher velocity than the measurement we described. Here the low-pass filter was set to 200 Hz. We can easily recognize the strong negative skewness in w at the edge (3) and the symmetry at the center (1). Except for the axial velocity at the center, all oscillations are in phase. There is also an indication in w 3 and u 3 for the second peak at twice the main frequency.
Two-Point Cross-Correlations Using two x-probes simultaneously, we can determine spatial cross-correlations of any velocity components. First, we look at correlations of data taken at different transverse positions (z) but at the same axial position (x).

In Fig. 22 a short section of a time series of the fluctuating velocity components is shown for two positions, namely, position 1 and 3, at the center and at the edge of the gap, respectively. The measurement
Velocity-Time Traces

Table 2. Flow Data Pertaining to the Figs. 18-20


Position No. U (m/s) u---u(m/s) 2 w---w(m/s) 2 uw (m/s) 2 Uskew Uflat Wskew Wfla`

1 2 3 4

11.54 12.40 14.73 16.77

0.976 1.895 4.938 4.749

3.596 3.712 3.061 1.066

--0.172 1.731 3.095 1.338

0.188 3.056 0.386 2.785 --0.003 2.189 --0.377 2.643

--0.007 --0.395 --0.839 --0.498

1.671 1.863 2.834 3.803

298

L. Meyer and K. Rehme

1
W

~2 2
in steps of 40 m m upstream at the same transverse position. The measurements shown were taken at the edge of the gap, but the results regarding the dependence on axial distance are independent of this transverse position. Figure 24 shows clearly the linear relation between the phase shift and the axial distance between the two measurement positions. At 200 mm the phase shift is zero, (by chance) the length of one period. The cross-correlations of the axial velocities have similar behavior and are not shown. A n o t h e r way to determine the length of one wavelength is to plot the phase between two signals versus the frequency (Fig. 25). For this two-point measurement the autospectral density function had its main peak at f = 75

Figure 21. Qualitative joint probability density functions of the axial (u) and the transverse (w) velocity components at four positions in the gap No. 9. For captions see Fig. 18.

Figure 23 shows the correlations of data taken with one probe at the fixed position 0 at the edge of the gap and the second probe moving through the whole length of the gap. The phase shift of the transverse velocity component w is small. F r o m phase plots (not shown) we get a maximum phase shift of approximately 30 at the center; this phase shift is independent of the frequency. The correlations of the axial velocity fluctuations, however, show little phase shift for the velocities at the same side of the symmetry line only but a shift by 180 for velocities taken at different sides of the symmetry line. Finally, we will look at correlations of data taken at different axial positions (x) but at the same transverse position (z). The correlations shown in Fig. 24 are of data taken with one probe (1) fixed at a position at the usual axial measurement plane and the second probe (2) moved

(m/s) w3 2 0 -2

....

, ....

i ....

i ....

i ....

r ....

i ....

i ....

i ....

i ....
R w l

w2

18 u3 16
14

- l

. . . .

. . . .

. . . .

. . . . .

. . . .

. . . .

.... 0.02

-0.02 1

-0.01

0.01
i i

2
wt 0

Rut

u~

-2 16 14 12
0 0.05 .1 .15 .2 .25

LI 1

,
-0.01

[
. . . .

4 J . . . . L . . . . i . . . .

-1 -0.02

0.01

0.02

time

(s) Figure 23. Spatial cross-correlation functions of (a) transverse velocity and (b) of axial velocity components, measured at different transverse positions in the gap at the same axial position; probe 2 was at the fixed position 0 at the edge of the gap; probe 1 was at the positions 1 through 4. Pos. 1, z/d = -0.25; pos. 2, z/d = 0; pos. 3, z/d = 0.25; pos. 4, z/d = 0.5.

Figure 22. Time traces of the axial (u) and the transverse velocity (w) at two positions in the gap No. 9, measured simultaneously. Low-pass filter setting was 200 Hz. Position 1 is at the center of the gap, z/d = 0; pos. 3 is at the edge of the gap, z/d = - 0 . 5 . The full line marks the mean velocity, the dashed line is the transport velocity of the fluctuation Uc.

Large-Scale Turbulence Phenomena

299

Rwl

~2

80 180

mm

120 mm 200
_[ -0.02 .... i .... i .... -0.01 i .... .... i .... 0.01 mm mm i .... J .... 0.02

(S)
Figure 24. Spatial cross-correlation functions of the transverse velocity w, measured at different axial positions at the edge of the gap ( z / d = -0.5). Probe 2 was moved upstream relative to probe 1 in steps of 40 mm.

From the correlations we know that a velocity component directed into the gap is linked to an above average component of the axial velocity (positive u) and vice versa. For an exact picture of the oscillation or vortex in the gap, multipoint measurements or flow visualization have to be performed. However, we can propose a flow model that is suited for all measured data: two vortices driven by the higher velocities outside the gap and rotating in opposite directions are transported with the velocity Uc axially within the gap. The centers of the vortices are on both sides of the centerline, and the diameters are roughly one gap of depth d (Fig. 26). The axial distance of the vortex centers is A/2. Channel with a Slot All types of measurements in the connected channels were also performed in a channel with one axial slot in one of the walls. Seven different depths of the slot were tested (No. 12-18 in Table 1). The flow in this channel was changed by the slot in a similar manner as that in the two channels connected by the axial gap in the wall; therefore, those isoline plots will not be shown. The large-scale fluctuation in and near the slot was also found there. Autospectral Power Density Function Figure 27 shows the autospectral power density functions of w measured at the edge of six different slots with the same bulk velocity of 21.8 ms ~. The frequency of the peak is low with deep slots and high with shallow slots, while the height of the peak decreases with decreasing slot depth. Not shown here is the curve of the shallowest slot with a depth of 1.7 times its width (16.6 mm), because the peak can hardly be detected. The curves pertaining to a slot depth of 77 and 58 mm deviate somewhat of that regularity. The curve of the 58-ram slot (No. 14) is similar to that of the 68-mm slot (No. 13), the peak frequency being rather lower instead of being higher. This puzzling result, which was checked by repeated measurements, might hint at effects of vortex sizes being related to the slot depth and needs to be further investigated. The power spectra of the axial velocity component u display peaks at the same frequencies (fig. 27b); the power density at frequencies lower than the peak is higher by an order of magnitude than that of the transverse velocity component. Velocity-Time Traces In Fig. 28 a sample of velocity-time traces is shown measured simultaneously at the edge and 40 mm into the slot, at the centerline of a Table 3. Characteristic Data in Gap No. 9, d = 77 mm for Five Different Velocities

Hz. At this frequency there is a phase shift of 27r between measurements 200 mm axially apart. We now have all the information to characterize the large-scale oscillation through the gap. First of all, it moves with a certain velocity irrespective of the position in the gap. This transport velocity can be determined in various ways (from Figs. 24 and 25) and relates to Uc = Aft~eak , with A being the length of one period. ~able 3 gives data of five measurements taken at different velocities. The velocity Uc lies between the two velocities at the center of the gap Umin and at the edge Ue. Considering the scatter and the uncertainty of the measurements, Uc turns out to be the average of those two velocities with this geometry of the gap. The length A of a large-scale fluctuation is constant, independent of the flow velocity and only a function of geometry.

2~

12

AX ( r a m )

/ ~ " / / /

/t /

3 r~/z

--

200 160

(D

lZO / /" -80 / / - , o / , / /


f.P

/ .
/f ~/

t~ a=

r/'~
0 0 20

/r.,

Ue (m/s)

Umin (m/s)
8.4 10.7 13.1 15.3 18.0

Uc (m/s)
9.8 12.2 15.0 17.0 20.0

A (mm)
200 200 201 200 200

f (Hz)
49 62 75 86 100

40

60

80

100

frequency

(Hz)

11.3 14.2 17.3 20.1 23.6

Figure 25. Phase plot of the transverse velocities w, measured at different axial positions at the edge of the gap. Probe 2 was moved upstream relative to probe 1 in steps of 40 mm.

*U~ is the velocity at the edge of the gap, Umi. is the minimum velocity at the center of the gap, U~ is the transport velocity of the large-scale fluctuation, A is the wavelength of the fluctuation, and f is the frequency of the fluctuation.

31111 L. Meyer and K. Rehme L

g=77

12

14

lfi

12

14

13

I i i iii

,.-~

is

\
~#2 =

1 0 -1

i0_2

i0_3

I 0-'

,iti

lllL

68

68

1 O0
i i i i

10 frequency
a
i ' ' i i [ i 13 i '

100 (Hz)

' ' ' I

12 14

l 0 -1

16

17

1 0 -2

i0-3

li!!~.-13

15

17

'~

10

-4

,,

,,,,I

,,

, ,,,I

, , , 4 6

68 10

68
100

frequency
b

(Hz)

Figure 27. Autospectral power density. (a) @w for the transverse, and (b) @. for the axial velocity component, measured at the edge of slots with different depth d. 77-mm deep slot. The low-pass filter was set to 200 Hz in order to identify the large-scale oscillations more easily. A striking feature of the w-traces is their asymmetry. The change from negative values (directed out of the slot) to positive values (into the slot) is slow, while the change from positive to negative values occurs abruptly. This feature can also be detected in the time records of the velocity traces by Knight and Shiono [6]. Corresponding with the latter the axial velocity drops from its high value to a low one, since a flow out of the slot carries a low-velocity mass and vice versa. The amplitudes of the fluctuations of the velocity components are space dependent. The highest amplitudes of the transverse component was found in all slots at about 10 to 20% slot depth from the edge into the slot. Close to the bottom of the slot this component decreases, being at 8 mm from the bottom wall one-third of the maximum. Also 10 mm outside the slot the amplitude of the w-component decreases to a similar value. The opposite is the case for the amplitudes of the axial velocity fluctuation. The maximum values are found close to the bottom and

4-

Figure 26. Flow model for the gap region. Top: side view; bottom: top view.

Large-Scale Turbulence Phenomena


( m / s ) | i , i

301

. i

i AIO (mm)

we

2 0 -2

20
30

19"wl 1,2
0

40
50

18 16
Ue

14 12 10
8

-1 -0.04

. . . .

. . . .

i -0.02

. . . .

. . . .

.... 0

i ....

t .... 0.02

i .... 0.04

Rul

t~

w40

0
-2

S
. . . . ' . . . . t . . . . i . . . . . 0 . . . i . . . . i 0.02 . . . . i . . . . 0.0 -0.02

12

U4o I 0 8
-1 -0.04

.I

.2 time

.3 (s)

.4

.5

(s)
b

Figure 28. Time traces of the axial (u) and the transverse (w) velocity at two positions in the slot No. 12 with d = 77 mm. Index e, measured at the edge, index 40, measured 40 mm into the slot. The full line marks the mean velocity, the dashed line is the transport velocity of the fluctuation U~; positive w is directed into the slot. outside the slot, but the variation is less than a factor of 2 between minimum and maximum amplitudes. Two-Point Cross-Correlations The velocity-time traces and, even better, the spatial cross-correlations in Fig. 29 indicate that there is almost no phase shift between transverse velocity components measured at different positions into the slot. The axial velocity fluctuations, however, display an increasing phase shift going into the slot. Consequently, a similar phase shift exists between locally measured u and w values. The phase shift between u and w turned out to be similar for different slot depths, being - 3 8 near the slot edge and increasing to - 9 0 near the bottom of the slot. These data are valid for frequencies close to the main frequency of the fluctuation; at higher frequencies the phase shift is lower. Spatial cross-correlations in both directions, transverse

Figure 29. Spatial cross-correlation functions of (a) transverse velocity, and (b) of axial velocity components, measured at different transverse positions in the slot No. 12 at the same axial position. Probe 2 was at the fixed position + at the edge of the gap; probe 1 was moved in steps of 10 mm into the slot.

and axial, were measured for three slot depths, namely, 37, 58, and 77 m m (No. 16, 14, 12), each at five different velocities. The results are too numerous to be shown here. F o r the cross-correlation of w and u at different axial positions similar results as those shown in Figs. 24 and 25 for the channels connected by a gap were obtained. Table 4 lists some characteristic data for three slot depths at the minimum and maximum measured velocity. The period length A varied somewhat with different velocities by less then 10%. The ratio of the velocity at the slot edge Ue and the transport velocity of the oscillation Uc varied little with velocity, but they vary with the slot depth. The flow model of two vortices that fitted to the flow in the gap does not fit to the flow in the slot. But this is not surprising because the gap is closed on one side. The frequency is lower and the wavelength is larger than in

Table 4. Characteristic Data of Measurements at Three Slot Geometries

Slot Depth (mm)

Ue (m/s)

U~ (m/s)

Ue / UC

f (Hz)

fd / Ue

A (ram)

35 57 77

9.0-18.5 10.3-20.9 10.0-20.9

6.5-14.0 5.5-12.3 6.1-12.1

1.38-1.32 1.87-1.70 1.73-1.75

33-70 21-43 18-37

0.128-0.132 0.122-0.117 0.139-0.133

195-200 300-320 300-325

*Shown are the data at the minimum and the maximum velocity measured. Nomenclature as in Table 3.

302

L. Meyer and K. Rehme


300

No. g
X

d
vJ ~ ~ "

.. j

No. d ( m m )

'

/ / ~

4 5
6

2 20 ,5 2 0
5 40

Y~/ 100 A/ 90 80 70 r~
N

: 18 ,7
16 15 14 O13 [] 12

//.

200

II

577

[] o

a lo 40 9,0 77y

/ /

37 48

1olo V

60 50

77

100

9o ao

~-~

40

~:

7o
t~
60 50

30

20

"Re x

2~ 10 5

2 Re x 105

Figure 30. Frequency of the large-scale oscillations for different gap geometries versus the bulk Reynolds number. comparable gap geometries. A single small vortex within one period could be responsible for the abrupt change from positive to negative w-values. Effects of Gap / Slot Geometry Strouhal Number The spectral power densities were measured at five velocities for each geometry. The frequencies at which a peak in the spectra was clearly detectable are plotted in Fig. 30 versus the Reynolds number for all channels connected by a gap. The Reynolds number z, is defined with the bulk average axial velocity in the channel Uu and the hydraulic diameter D h of the channel neglecting the gap. From Fig. 30 it is obvious that the frequency of the peak in the power spectrum increases with decreasing length of the gap (g/d = 5/77, 5/40, 5/20 or 10/77, 10/40). The frequency also increases with decreasing width of the gap (g/d = 5/20, 2.5/20 or 10/40, 5/40) with the exception of the geometries with a length of the gap of 77 mm. The frequencies are almost the same for the two different gap widths (No. 7 and 9). We tried different length scales to obtain a constant Strouhal number for all geometries with a gap. The result is to define a Strouhal number with the frequency f, the velocity Ue at the edge of the gap at the symmetry line through the gap, and a length scale as the square root of width times depth of the gap ReUbO h

Figure 31. Frequency of the large-scale oscillations for different slot geometries versus the bulk Reynolds number. shown in Fig. 31 as a function of the Reynolds number. Again, the frequencies increase linearly with the Reynolds number. The frequencies increase with decreasing length of the slot with one exception (58 and 68 mm, respectively). If the Strouhal numbers are defined with the depth of the slot d as length scale Str s = fd/U~, the Strouhal numbers are within & _ 20% independent of the slot depth. The mean Strouhal number is 0.135. Velocity Gradient In order to furnish data for modeling the flow in slots, the centerline velocities are presented in Fig. 32. The velocities, normalized with the velocity at the edge of the slot, are plotted versus the relative slot depth up to 0.8 slot depth into the main channel. The relative velocities are invariant with the absolute velocity and vary little, although systematically with the slot depth, with the exception of the 68-mm relative to the 77-ram results.
1.6

No. d ( m m )
1.4 ----- ---

12 13 14
1,5

77 68
58 48

1.2

-----

lfl 17 18

37 29 17

f
i I l l

S: ....................

.8

f(gd) 1/2
Str o The Strouhal numbers for all channels with a gap are constant within _+16%. The mean Strouhal number is 0.115. The frequencies at which the peak in the spectra was observed in the channel with slots of different depths are

.6

.4 .8 .6 .4 .2 0

-.2

-.4

-.6

-.8

z/d

32. Relative velocity at the centerline of the slots versus relative slot length. Origin of z-coordinate is at the edge of the slot.
Figure

Large-Scale Turbulence Phenomena 303


1.4 , , , No. , , d(mm) 77 68 , , , , , , , , , , ,

1.2

- ----

12 13

/:~,
,//~'~

--

,4
15

58
48

//',~
///,~_ ""~\

--

18

a7

//#-

-\~\

).6

.4

.~'

.8

.6

.4

.z

o z / d

-.2

-.4

-.6

-.8

Figure 33. Relative velocity gradient at the centerline of the slots versus relative slot length. All profiles have a turning point, which is a precondition for an instability. Figure 33 shows the relative velocity gradient, and it is peculiar that the value at the slot edge is similar for all slot depths between 37 and 77 mm. The absolute height of the velocity gradient increases systematically with increasing depth of the slot, again with the exception of the 77-mm results. PRACTICAL S I G N I F I C A N C E / U S E F U L N E S S The results of this investigation shed a new light on the sediment transport phenomena in river engineering and confirm earlier findings relating to interchannel mixing in axial flow through rod bundles. Both transport processes are predominantly not governed by secondary flows but by large eddies with a geometry-dependent length scale and flow velocity-dependent frequency. It was shown that a pulsating flow occurs whenever a small flow cross section connects to a large cross section or when two large cross sections are connected by a small cross section. Within the small cross section, large eddies transport heat or mass effectively in or out of the small cross section or from one large cross section into the other. These effects are not predicted by the current fluid dynamics computer codes employing conventional turbulence models. The simple geometry involved should make it a challenge to test codes and models against the presented data. CONCLUSIONS Detailed measurements of the axial component of the mean velocity and of 5 of the 6 Reynolds stresses have been made in 18 different geometries. The compound flow channels consisted either of two rectangular channels connected by a gap or of one rectangular channel with a slot in one wall. The results for the smallest gap confirm the qualitative features of turbulent flow through rectangular ducts, such as those by Melling and Whitelaw [24]. The data, however, seem to be quantitatively more precise than those of previous investigations. The gap between the channels and the slot connected to one channel, respectively, strongly affect the turbulent flow features in the vicinity of the gap or slot. The turbulence intensities are very high close to the gap or

slot, respectively. The Reynolds shear stresses in the vicinity of the gap or slot are also much higher than anywhere else in the channel. A strong large-scale quasi-periodic flow oscillation was observed in most of the channels investigated. Such large-scale flow oscillations have been reported for axial flow through rod bundles and for flow in rectangular open channels bounded by flood banks. There is a strong peak in the power density spectra of the axial and transverse velocity components and of the shear stress. The frequency at which the peak in the spectrum occurs varies linearly with the Reynolds number. The Strouhal number of the frequency peak seems to depend only on the gap/slot geometry. It seems that we have a new type of instability for a fluid without any unsteadiness in the velocity distribution. This instability is quasi-periodic and can be measured at any axial position along the channel. It exists in any longitudinal slot or groove in a wall or a connecting gap between two flow channels, provided its depth is more than approximately twice its width. RECOMMENDATIONS AND FUTURE RESEARCH NEEDS More parameter variations of the gap/slot geometry are necessary to find the exact relations between the geometry and the frequency and size of the oscillations. The role of the velocity gradient has to be quantified. To confirm the nature of the oscillations/eddies it is desirable to perform flow visualization studies in various geometries. Experiments should be designed to investigate the flow at lower Reynolds numbers with the objective to find the lower bound for the occurrence of the large-scale eddies. The effect of many parallel slots, separated by a thin wall only, should be studied, possibly finding a relation to the flow over riblets. Theoretical work should be attractive in the field of the analysis of stability of parallel flow, in order to provide insight into the originating mechanism for the instabilities that lead to the periodic oscillations. The application of a large eddy simulation (LES) code should yield similar results and provide a tool to perform parameter variations more easily than experimentally. The authors thank G. WBrner and E. Mensinger, KfK-INR, for the performance of the measurements, V. B~iumer, University of Braunschweig, for writing the program for the frequency analysis, and Dr. W. V~ith, KfK-INR, for the fruitful discussions on aspects of the frequency analysis and their interpretation.

NOMENCLATURE a, b d Dh f g k L p R width of channel (Fig. 2), m depth of gap/slot, m hydraulic diameter of channel, m frequency, Hz width of gap/slot, m kinetic energy of turbulence, m2/s 2 length of flow channel, m probability density function, dimensionless correlation function, dimensionless

304

L. Meyer and K. R e h m e Re Str u R e y n o l d s n u m b e r , dimensionless Strouhal n u m b e r , d i m e n s i o n l e s s fluctuating velocity in axial direction, m / s t u r b u l e n c e intensity in axial direction, m / s
2-.v/p

8.

u,

friction velocity, m / s m e a n axial velocity, m / s transport velocity of vortex, m / s fluctuating velocity in direction n o r m a l to gap/slot, m/s t u r b u l e n c e intensity in transverse direction n o r m a l to gap, m / s t u r b u l e n c e intensity in transverse direction, parallel to gap, m / s axial dimension, m transverse dimension, n o r m a l to gap, m transverse dimension, parallel to gap, m

9.

U u
v' =

10.

11.

12. 13. 14.

x y z )t

G r e e k Symbols length of o n e p e r i o d of large-scale oscillation, m v viscosity of air, m 2 / s t i m e delay in cross-correlation function, s r w wall shear stress, Pa ~'av shear stress a v e r a g e d along all f o u r walls, Pa spectral p o w e r density, m 2 / s

15.

16.

17.

Subscripts
b e G S bulk e d g e of g a p / s l o t gap slot

18.

19.

REFERENCES
1. Lyall, H. G., Measurement of Flow Distribution and Secondary Flow in Ducts Composed of Two Square Interconnected Subchannels, Symp. on Internal Flows, Salford, UK, EI6-E23, 1971. 2. Knight, D. W., and Hamed, M. E., Boundary Shear in Symmetrical Compound Channels, J. Hydraulic Engrg. 110, 1412-1430, 1984. 3. Knight, D. W., and Lai, C. J., Turbulent Flow in Compound Channels and Ducts, Int. Symp. on Refined Flow Modelling and Turbulence Measurements, Iowa, pp. I2-1-I2-10, 1985. 4. Shiono, K., and Knight, D. W., Transverse and Vertical Reynolds Stress measurements in a Shear Layer Region of a Compound Channel, 7th Int. Symp. on Turbulent Shear Flows, August, Stanford, CA, Paper 28-1, 1989. 5. Shiono, K., and Knight, D. W., Mathematical Models of FLow in Two or Multi Stage Straight Channels, Proc. Int. Conf. on River Flood Hydraulics, W. R. White, Ed., Wallingford, UK, Paper G1, 229-238, Wiley, New York, 1990. 6. Knight, D. W., and Shiono, K., Turbulence Measurements in a Shear Layer Region of a Compound Channel, J. Hydraulic Res. IAHR, 28(2), 175-196, 1990. 7. Shiono, K., and Knight, D. W., Turbulent Open-Channel Flows

20. 21.

22. 23. 24. 25.

26. 27.

with Variable Depth across the Channel, J. Fluid Mech. 222, 617-646, 1991. Knight, D. W., Samuels, P. G., and Shiono, K., River Flow Simulation: Research and Developments, J. Institution of Water and Environmental Management 4(2), 163-175, 1990. Shiono, K., and Knight, D. W., Turbulence Measurements by LDA in Complex Open Channel Flows, Proc. 5th Int. Symp. on Applications of Laser Techniques to Fluid Mechanics, July, Lisbon, Portugal, Paper 3.5, 1990. Rowe, D. S., Measurement of Turbulent Velocity, Intensity and Scale in Rod Bundle Flow Channels, BNWL-1736, Battelle Pacific Northwest Laboratories, Richland, WA, 1973. Hooper, J. D., and Rehme, K., Large-Scale Structural Effects in Developed Turbulent Flow Through Closely-Spaced Rod Arrays, Z Fluid Mech. 145, 305-337, 1984. M611er, S. V., On Phenomena of Turbulent Flow through Rod Bundles, Exp. Thermal Fluid Sci. 4, 25-35, 1991. M611er, S. V., Single-phase Turbulent Mixing in Rod Bundles, Exp. Thermal Fluid Sci. 5, 26 33, 1992. Rehme, K., The Structure of Turbulence in Rod Bundles and the Implications on Natural Mixing between the Subchannels, Internat. J. Heat Mass Transfer 35(2), 567-581, 1992. Rehme, K., Experimental Observations of Turbulent Flow through Subchannels of Rod Bundles, Exp. Thermal Fluid Sci. 2, 341-349, 1989. Hejna, J., and Mantlik, F., The Structure of Turbulent Flow in Finite Rod Bundles, Proc. 1st World Conf. Exp. Heat Transfer, Fluid Mech., and Thermodynamics, R. K. Shah, E. N. Ganic, and K. T. Yang, Eds. Elsevier, New York, pp. 1712-1719, 1988. Wu, X., and Trupp, A. C., Experimental Study on the Unusual Turbulence Intensity Distributions in Rod-to-Wall Gap Regions, Exp. Thermal Fluid Sci. 6(4), 360-370, 1993. Patel, V. C., Calibration of the Preston Tube and Limitation on Its Use in Pressure Gradients, J. Fluid Mech. 23(1), 185 208, 1965. Lueptow, R. M., Breuer, K. S., and Haritonidis, J. H., Computer Aided Calibration of x-Probes Using a Look-up Table, Exp. Fluids 6, 115-118, 1988. Meyer, L., Calibration of a Three-Wire Probe for Measurements in Nonisothermal Flow, Exp. Thermal Fluid Sci. 5, 260-267, 1992. Bremhorst, K., The Effect of the Wire Length and Separation on x-Array Hot Wire Anemometer Measurements, IEEE Trans. Instr. Measurement IM-21,244-248, 1972. Vagt, J.-D., Hot Wire Probes in Low Speed Flow, Progr. Aerosp. Sci. 8, 271-323, 1979. Moffat, R. J., Describing the Uncertainties in Experimental Results, Exp. Thermal Fluid Sci. 1, 3-7, 1988. Melling, A., and Whitelaw, J. H., Turbulent Flow in a Rectangular Duct, J. Fluid Mech. 78, 289 315, 1976. Demuren, A. O., and Rodi, W., Calculation of Turbulence-Driven Secondary Motion in Noncircular Ducts, J. Fluid Mech. 40, 189-222, 1984. Knight, D. W., and Patel, H. S., Boundary Shear in Smooth Rectangular Ducts, J. Hydraulic Engrg., 111(1), 29-47, 1985. Hur, N., Thangam, S., and Speziale, G. C., Numerical Study of Turbulent Secondary Flows in Curved Ducts, J. Fluids Engrg. 112, 205-211, 1990.

Received September 13, 1993; revised January 18, 1994

You might also like