Experimental study of laminar and turbulent boundary layer separation [PDF]

Dec 20, 2016 - Keywords: boundary layer control, separation, Mako shark skin, bristling of scales, adverse pressure ... over the shark's body. Ultimately, man-made rib- lets showed a turbulent skin friction drag reduction of. 9.9% over that of a smooth surface (Bechert et al 1997). ...... University of Alabama TUSCALOOSA.

0 downloads 10 Views 7MB Size

Recommend Stories


Turbulent Boundary-Layer Interaction
Stop acting so small. You are the universe in ecstatic motion. Rumi

Boundary layer separation
Be like the sun for grace and mercy. Be like the night to cover others' faults. Be like running water

supersonic laminar boundary layer separation near a compression corner
The beauty of a living thing is not the atoms that go into it, but the way those atoms are put together.

6.1 Laminar and Turbulent Flow
Before you speak, let your words pass through three gates: Is it true? Is it necessary? Is it kind?

Controlling the Separation of Laminar Boundary Layers in Water
You have survived, EVERY SINGLE bad day so far. Anonymous

Experimental study of relative, turbulent diffusion
I tried to make sense of the Four Books, until love arrived, and it all became a single syllable. Yunus

flow separation characterisation of a forward facing step immersed in a turbulent boundary layer
No amount of guilt can solve the past, and no amount of anxiety can change the future. Anonymous

Experimental Investigation of the Boundary Layer Transition on a Laminar Airfoil Using Infrared
Don’t grieve. Anything you lose comes round in another form. Rumi

Separation Criterion for Turbulent Boundary Layers Via Similarity Analysis
Don't fear change. The surprise is the only way to new discoveries. Be playful! Gordana Biernat

Aero-Optical Measurements in a Heated, Subsonic, Turbulent Boundary Layer
No amount of guilt can solve the past, and no amount of anxiety can change the future. Anonymous

Idea Transcript


Home

Search

Collections

Journals

About

Contact us

My IOPscience

Experimental study of laminar and turbulent boundary layer separation control of shark skin

This content has been downloaded from IOPscience. Please scroll down to see the full text. 2017 Bioinspir. Biomim. 12 016009 (http://iopscience.iop.org/1748-3190/12/1/016009) View the table of contents for this issue, or go to the journal homepage for more Download details: IP Address: 131.247.215.9 This content was downloaded on 07/07/2017 at 18:05 Please note that terms and conditions apply.

You may also be interested in: Use of a rotating cylinder to induce laminar and turbulent separation over a flat plate F Afroz, A Lang and E Jones Movable shark scales act as a passive dynamic micro-roughness to control flow separation Amy W Lang, Michael T Bradshaw, Jonathon A Smith et al. Bristled shark skin: a microgeometry for boundary layer control? A W Lang, P Motta, P Hidalgo et al. Separation control over a grooved surface inspired by dolphin skin Amy W Lang, Emily M Jones and Farhana Afroz Hydrodynamic function of biomimetic shark skin: effect of denticle pattern and spacing Li Wen, James C Weaver, Patrick J M Thornycroft et al. A bio-inspired device for drag reduction on a three-dimensional model vehicle Dongri Kim, Hoon Lee, Wook Yi et al. Surface micro-grooves for near-wall exergy and flow control G F Naterer, P S Glockner, D Thiele et al. Streamwise vortices originating from synthetic jet–turbulent boundary layer interaction D Lasagna, M Orazi and G Iuso Study on wake structure characteristics of a slotted micro-ramp with large-eddy simulation Xiangrui Dong, Yaohui Chen, Gang Dong et al.

Bioinspir. Biomim. 12 (2017) 016009

doi:10.1088/1748-3190/12/1/016009

PAPER

RECEIVED

7 May 2016 REVISED

21 September 2016 ACCEPTED FOR PUBLICATION

7 November 2016 PUBLISHED

20 December 2016

Experimental study of laminar and turbulent boundary layer separation control of shark skin Farhana Afroz1, Amy Lang1, Maria Laura Habegger2, Philip Motta2 and Robert Hueter3 1 2 3

Department of Aerospace Engineering and Mechanics, The University of Alabama, USA Department of Integrative Biology, University of South Florida, USA National Center for Shark Research, Mote Marine Laboratory, Sarasota, FL, USA

E-mail: [email protected] Keywords: boundary layer control, separation, Mako shark skin, bristling of scales, adverse pressure

Abstract The Shortfin Mako shark (Isurus oxyrinchus) is a fast swimmer and has incredible turning agility, and has flexible scales known to bristle up to 50° in the flank regions. It is purported that this bristling capability of the scales may result in a unique pass flow control method to control flow separation and reduce drag. It appears that the scales have evolved to be only actuated when the flow over the body is reversed; thereby inducing a method of inhibiting flow reversal close to the surface. In addition, bristled scales form cavities which could induce boundary layer mixing and further assist in delaying flow separation. To substantiate the hypothesis, samples of skin from the flank region of the mako have been tested in a water tunnel facility under various strengths of adverse pressure gradient (APG). Laminar and turbulent separation over the skin was studied experimentally using time-resolved digital particle image velocimetry, where the APG was generated and varied using a rotating cylinder. Shark skin results were compared with that of a smooth plate data for a given amount of APG. Both the instantaneous and time-averaged results reveal that shark skin is capable of controlling laminar as well as turbulent separation. Under laminar conditions, the shark skin also induces an early transition to turbulence and reduces the degree of laminar separation. For turbulent separation, the presence of the shark skin reduces the amount of backflow and size of the separation region. Under both flow conditions, the shark skin also delayed the point of separation as compared to a smooth wall.

1. Introduction Drag reduction inspired by nature has the potential to come up with practical applications for flow control such as aircraft, submarines, ships etc. For example, surfaces inspired by shark skin have been considered for the past three decades in the development of dragreducing riblets. The shark skin is covered by minute placoid scales, known as dermal denticles, which primarily serve the purpose of a protecting armor (Raschi and Tabit 1992). The base of the tough toothlike scales embedded in the external collagenous layer of the skin (dermis) allows the crown of each scale to be exposed to the water. Fast swimming sharks are particularly known for having streamwise ridges along each crown with a measured degree of flexibility on some body locations. It has been suggested and shown that these unique features of the shark skin can lead to drag reduction by passive flow control mechanisms © 2016 IOP Publishing Ltd

(Lang et al 2008, 2011, 2012, Motta et al 2012, Oeffner and Lauder 2012). Early research focused on the hydrodynamic effects of the ridges on the crown of each scale—also known as riblets—to reduce the turbulent skin friction drag over the shark’s body. Ultimately, man-made riblets showed a turbulent skin friction drag reduction of 9.9% over that of a smooth surface (Bechert et al 1997). In prior work, the model scales were made to lie flat on the surface, and the bristling of the scales was not investigated. Later studies built a shark skin replica consisting of an array of shark scales with compliant anchoring and then tested the model in a fully turbulent oil channel for measuring the reduction in skin friction drag (Bechert et al 2000a, 2000b). This experiment showed that when the scales had bristling capability, but were laid flat, the drag reduction was found to be only about 3%. However as the shark scales were bristled up to 12°, an increase in skin

Bioinspir. Biomim. 12 (2017) 016009

F Afroz et al

Figure 1. Outline of measured scale bristling angles at various locations on the Mako Shark. Modified with permission from Motta et al (2012). Copyright 2012 Wiley.

friction drag resulted. Other research also showed shark scale riblet models to be capable of reducing skin friction drag, and thus bristling capability was not found to act as a means to reduce skin-friction drag (Zhao et al 2012, Bixler and Bhushan 2013). However, the recent research study focuses on flexibility of individual scales and bristling capability of those scales actuate passive flow separation. When flow separates, it results in pressure drag. The bristling capability of flexible scales may control the flow separation and thus reduce pressure drag that acts opposite direction of body motion. It has been hypothesized in some earlier research work that bristled shark scales could act as vortex generators to control the flow separation (Bechert et al 1985, 2000a, 2000b). Vortex generators provide a passive mechanism for controlling flow separation (Lin 2002). These are small, vertical v-shaped structures at the surface that protrude into the boundary layer. As the flow encounters the vortex generators, stream wise vortices are created downstream and mix high momentum turbulent energy in the boundary layer. Scales over shark skin are hypothesized to bristle passively in the reversed flow region near or downstream of the point of separation (Lang et al 2011). Thus bristling of shark scale offers new and different mechanism for controlling flow separation from that of traditional vortex generators. As the flexible denticles bristles forms two-dimensional transverse grooves between scales. The idea that two-dimensional transverse grooves could be a possible means of drag reduction was proposed by Bushnell (1985). As flow passes over a single groove, a partialslip condition results from embedded vortices of trapped fluid in each groove. The embedded vortices in the cavities imposes a shear stress at the bottom of cavity which results a small thrust adding to the net reduction in drag for the surface. This was termed a ‘microair bearing effect’ (Bushnell 1985). However, the mixing induced into and out of the cavities increased the overall skin friction drag. Howard and Goodman 2

(1985) observed the effect of transverse embedded grooves on the drag characteristics of axisymmetric bluff bodies where it was predicted that the grooves would help to control flow separation due to mixing enhancement in the boundary layer, and thus reduce pressure drag. A drag decrease and corresponding reduction in the size of the separation region was found in this study. This and other studies have indicated that embedded cavity or d-type roughness surfaces can lead to passive turbulence augmentation to control flow separation (Gad-el-Hak 1989). A recent study conducted on a flapping robotic foil determined the self-propelled swimming speed as a measure of locomotive performance of both mako shark skin covered foils versus sanded skin surface (Oeffner and Lauder 2012). Shark skin with the intact surface denticles showed a mean 12.3% increase in swimming speed compared with the same sanded skin foils. This study also suggested that shark skin scales may alter vortex location which could increase thrust. But questions still remained as to the mechanism by which the shark skin scales modify the boundary layer flow. The present study focuses on investigating flow over skin samples from the Shortfin Mako (Isurus oxyrinchus) as it is one of the fastest swimming marine creatures (Stevens 2008). These scales are 0.2 mm in average size and are uniformly oriented from nose-totail except around pit organs and lateral lines of shark body. These tiny scales are anchored to the shark’s skin in such a way that they are able to pivot about their base only when flow is reversed in the streamwise direction (Motta et al 2012). Under superambient subcutaneous pressure, randomly selected scales in different regions of two dead shortfin mako sharks skin were softly operated with a fine acupuncture needle to the maximum resting erection angle and the maximum bristling angles of scales in different regions is presented in figure 1 (Motta et al 2012). Among the bristling angle for 16 different regions it can be found that the highest scale flexibility occurs on the flank and

Bioinspir. Biomim. 12 (2017) 016009

F Afroz et al

Figure 2. Scales from flank region, (a) SEM of non-bristled Mako scales. (b) SEM of manually bristled scales and the formation of cavities between the scales.

trailing edges of the pectoral fin, with bristling angles up to approximately 50° on the flank. It is very interesting that the regions where maximum bristling occurs correspond to the regions that one would predict the presence of an adverse pressure gradient(APG) due to the shark’s body shape. Moreover, scanning electron microscope images of individual scales showed that the geometry of the structural base differs by body location. More firmly anchored scales (lower bristling capability) have a more symmetrical base while flank region scales have bases that are wider than deep, facilitating pivoting and thus scale bristling (Motta et al 2012). It is hypothesized that for a flow experiencing an APG, it is the region closest to wall with lowest momentum where reverse flow is initiated. This reversed flow actuates the scales thereby disrupting the process leading to flow separation; this is considered to be the primary mechanism leading to separation control. Several research (Lang et al 2011, 2012, 2014) also suggested that these scales, which are loosely embedded in the shark’s skin, are able to erect in the presence of reversing flow, thereby trapping the reversing flow while also forming cavities within the surface (as observed in figure 2) that can lead to mixing enhancement and increased momentum in the near-wall region of the boundary layer as a secondary mechanism to control flow separation. The formations of cavities (about 200 μm in size) between the scales of real shark skin in a bristled orientation are shown in figure 2(b). The scales are oriented in staggered rows with overlap between the tips of the scales on one row and the bases of the scales on the following row. Previous experiments conducted by Lang et al (2008) over a bristled shark skin model revealed the formation of embedded vortices between replicas of the scales. As the scales bristle, the formation of embedded vortices would allow the flow to pass over the skin with a partial slip condition, thereby leading to higher momentum adjacent to the surface, while also increasing mixing and thus momentum exchange within the boundary layer. Thus the bristled shark skin may have more than one technique of 3

controlling the boundary layer to decrease overall pressure drag. The present study provides insights to the method by which the scales can control the boundary layer flow with a focus on determining the capability of shark skin—from the flank region—to control flow separation over a flat, non-moving surface. This will lead to further validation that a passive, separation control mechanism is in fact present on the skin and that further research is necessary to elucidate this mechanism that will ultimately result in man-made bio-inspired surfaces for engineering applications where flow separation decreases the performance. Some of the examples of engineering applications include submarines, wind turbine blades, helicopter rotor blades, and aircraft control surfaces.

2. Experimental procedures Shark skin samples from the flank region—with most flexible scales—of a Shortfin Mako (Isurus oxyrinchus) were tested under different values of APG. The traditional method of inducing an APG over a plane surface is either changing the contour of the top wall (Sohn et al 1998) or varying the angle of a displacement body in the test section above the flat plate (Gaster 1969). In these methods, after a short region of the favorable pressure gradient, an APG induces the separation bubble. But both these techniques need an additional arrangement to prevent boundary layer separation over the displacement body itself for high curvatures and thus high APG. The variation of APG was performed with the use of a rotating cylinder which was proven to induce both laminar and turbulent boundary layer separation (Afroz et al 2014, 2016). This system is very simple where rotation speed of cylinder induces flow separation over flat surface at downstream location of the cylinder. This rotating cylinder system generates wide variation and controlled range of APG. Moreover, it does not induce any other separation region and thus no need of any other arrangement to prevent separation

Bioinspir. Biomim. 12 (2017) 016009

F Afroz et al

Figure 3. Schematic of the experimental set up showing the use of a rotating cylinder to induce flow separation over a flat plate and shark skin at downstream position of cylinder.

Figure 4. (a) Diagram of flat plate with shark skin mounted in it, (b) affixed real shark skin samples from flank region, frozen.

elsewhere in the system. Time-resolved digital particle image velocimetry (TR-DPIV) was used to compare the laminar and turbulent boundary layer separation over shark skin with that of a smooth plate data for a given amount of APG. It is not possible to achieve the real bursting speed of a shortfin mako shark (~10 m s-1) in water tunnel facility. But the results might specify the threshold reverse flow that requires actuating the flow separation control mechanism. The maximum flow speed capacity of water tunnel was 0.5 m s−1. Therefore, the shark skin was tested for one free stream tunnel speed (0.132 m s−1) at laminar flow condition with a local Reynolds number (Rex) of Rex » 1.3 ´ 105 in the measurement region based on the distance from leading edge of the flat plate. But for the turbulent flow, studies were done for two free stream tunnel speeds (0.25 m s−1) and (0.3 m s−1) which gave Reynolds number values of Rex » 2.5 ´ 105 and Rex » 3 ´ 105 in the measurement region based on the distance from leading edge. 4

2.1. Experimental setup The experimental studies were carried out in a water tunnel facility at the University of Alabama. The water tunnel has a test section size of 38 cm wide by 76 cm tall by 275 cm long. To induce the presence of an APG, a rotating cylinder (diameter of 5.1 cm) was placed at a distance of L=101.6 cm (≈20D) downstream from the leading edge of the flat plate and spanned the total width (45.72 cm) of the plate (figure 3). By adjusting the gap height (G) from the plate as well as the rotation speed, the strength of the APG was varied. This system induces separation over a flat plate at a point just downstream of the cylinder. For these experiments the cylinder center was placed at the beginning edge of shark skin to ensure the formation of separation occurs over the shark skin sample. First, the shark skin model was tested by placing the cylinder at certain gap WD height (G/D) and rotation speed VR = 2U for inducing a known amount of APG (Afroz et al 2014, 2016). Then keeping the cylinder at same

(

)

Bioinspir. Biomim. 12 (2017) 016009

F Afroz et al

Figure 5. Thereotical pressure coefficient gradient based on cylinder gap height, G/D=0.75, and different cylinder rotation speeds (VR), (a) Rex » 1.3 ´ 105, (b) Rex » 2.5 ´ 105 and, (c) Rex » 3 ´ 105 .

gap height (G/D) and rotation speed (VR), the shark skin plate was replaced by a smooth plate model. This procedure insured testing was performed for the shark skin and smooth surfaces under the exact same amount of APG for direct comparison. Experiments were carried out for both laminar and turbulent boundary layer conditions. As shown in figure 3, the flow was tripped to be turbulent by putting a thin plate (10 mm in height and 25 mm in wide) at a distance, LT=50 cm (≈10D) from leading edge for the latter case. The model used in the experiment is a flat plate interchangeable model, made out of black Plexiglas, and was mounted vertically in the test section. Attached to the front of this plate there is a leading plate with elliptical nose and then the flat plate model 5

which is followed by a trailing edge flap. This flap was adjusted based on flow speed and blockage to prevent the formation of a separation bubble at the nose of the flat plate. The shark skin samples from the flank region, three pieces each measuring 14 cm in width and 17.78 m in height with proper orientation of the scales maintained to ensure sensitivity to reversing flow, were glued on another flat plate surface (figure 4). Thus, the shark skin surface measured a height of 53.34 cm (about 87.5% of total plate height). It should be mentioned that special care was taken to cut the exact dimensions of shark skin. The muscles and tissues under the skin were removed with help of a surgical blade as much as possible without damaging the skin structure. However, even with this care there

Bioinspir. Biomim. 12 (2017) 016009

F Afroz et al

Figure 6. Experimental and theoretical velocity profile, (a) laminar flow (Rex » 1.3 ´ 105), (b) turbulent flow (Rex » 2.5 ´ 105).

Figure 7. Comparison of various characteristics value of theoretical and experimental laminar base flow.

was some minor waviness to the skin after being mounted to the flat plate (figure 4). Due to the fact that biological materials deteriorate at room temperature, the shark skin sample was kept frozen when not in use. Thawing and then refreezing causes eventual deterioration, thus this process was only performed twice to insure soundness of the skin sample during testing. Care has also been taken so that shark skin did not come in direct contact with other objects or surfaces to avoid any contamination or damage to the scales. 2.2. Data acquisition and analysis technique To measure the flow field with a DPIV system, the flow was seeded with 14 μm silver-coated hollow glass spheres and then illuminated by a laser sheet generated by a Quantronix Falcon 20 mJ Nd:YLF laser. The maximum power output of this laser is 20 watts and it has a beam wavelength of 532 nm, 7–30 amps current 6

output range, and 0.1–1.0 kHz frequency range. LabVIEW software here was used for image acquisition. Then Insight3G DPIV software was used to analyze the acquired images and obtain the velocity vector field for laminar and turbulent separation. The size of each measurement window for image acquisition was 4.5 cm×2.25 cm and images were captured by a Basler A504K high-speed camera at a rate of 1000 frames s−1 for turbulent flow and 500 frames s−1 for laminar flow. A total of 9600 images (9.6 s) for each rotation speed were captured at 1280×512 pixels and then analyzed to obtain both instantaneous and averaged statistics of the flow field. The images were captured 2.5 cm downstream of the cylinder center as identified in figure 3. DPIV analysis was conducted on the images with various sizing of interrogation windows, the smallest being 16 × 8 pixels for increased data resolution in the nearwall region. A minimum intensity image generator was used to determine the background noise common to all

Bioinspir. Biomim. 12 (2017) 016009

F Afroz et al

Figure 8. Time-averaged velocity contour for laminar flow separation at Rex » 1.3 ´ 105 and at different amount of APG (a) VR=1.208. (b) VR=1.409.

Figure 9. Comparison of separation point, transition point and LSB size between flat plate and shark skin at Re different amount of APG (a) VR=1.208, (b) VR=1.409.

Table 1. Comparison of LSB over flat plate and shark skin under different amount of APG. Smooth plate

Shark skin

VR

XS

XT

h

XS

XT

h

1.208 1.409

35 32

80 82

5.1 8.8

42 34

69 65

1.65 3.2

of the images, and this noise was subtracted from each image to increase the accuracy of the vector calculations. During processing, Insight 3G software utilized a 7

x

» 1.3 ´ 105 and at

Recursive Nyquist Grid, FFT Correlation Engine, and Gaussian Peak Engine while post-processing filtering was performed to eliminate any bad vectors due to localized insufficient seeding. The DPIV processing software gave an estimation of the average percentage of valid vectors in each processing window, and the least measurement of good vectors was about 97%. The error associated with the DPIV technique is primarily due to insufficient data (poor seeding density or poor image quality) (Willert and Gharib 1991, Hart 2000). In this experimental study, adequate seeding density was

Bioinspir. Biomim. 12 (2017) 016009

F Afroz et al

Figure 10. Time-averaged vorticity under same amount of APG, Rex » 1.3 ´ 105 at VR=1.409.

( ) field for laminar flow at Re

Figure 11. Time-averaged normalized Reynolds stress – APG, (a) VR=1.208, (b) VR=1.409.

u¢v¢ U2

maintained as much as was possible during data acquisition. In literature, an acceptable estimate of error is 5% in velocity measurements and 10% in vorticity (Willert and Gharib 1991, Huang et al 1997). The pixel/ cm calibration for these data sets was 37.8 microns/ pixel. Combining the error for tilting of the ruler for calibration process and processing grid error, the total error for each velocity vector was kept below 5%. 8

x

» 1.3 ´ 105 and at different amount of

3. Theoretical model and base flow 3.1. Inviscid model for inducing pressure gradient An inviscid model was used to quantify the strength of the pressure gradient induced along the wall for different cylinder rotation speed, VR, and cylinder location, G/D. This inviscid theoretical model for flow about a rotating cylinder located near a wall consists of

Bioinspir. Biomim. 12 (2017) 016009

F Afroz et al

Figure 12. Instantaneous velocity history flat plate versus shark skin for VR=1.208 at y=2 mm and at different x-positions (a) x=45 mm, (b) x=70 mm.

the superposition of a free stream flow, doublet and vortices. Figure 5 shows how velocity ratio (VR), Reynolds number (Rex) and cylinder gap height (G/D) impact the strength of the gradient of pressure p = ¶C p . In this case, the value of coefficient C ¶x

(

)

cylinder gap to diameter ratio (G/D) was fixed at 0.75 for all the experimental conditions. The pressure calculations along the flat plate by this theoretical model were also used in previous studies (Afroz et al 2014, 2016) for predicting the zone where both laminar and turbulent boundary layer was subjected to APG. It should be mentioned that this inviscid model can be used confidently to measure the pressure gradient for viscous flows. The rotation of cylinder induces the viscous wake to move up and away from the measurement area, thereby not disturbing the measurements. The rest of the viscous flow induced by the cylinder rotation can be presented to be that of an irrotational vortex (Kundu and Cohen 1990). So the uniqueness of this viscous flow solution that results in the formation of an irrotational votex rotation allows using the inviscid model to predict the pressure gradient outside the boundary layer. It should be noted the goal of the theoretical study was to identify the measurement region of experimental study where boundary layer separation starts and the starting point of each measurement region was chosen as x=25 mm. Beyond this point, the boundary layer is subjected to an extreme APG and separation is induced on a flat surface. Figure 5(a) shows the pressure gradient plots 9

that are used for the laminar flow separation study. Figures 5(b), (c) show the pressure gradient measurements which are used for the turbulent flow separation study. 3.2. Base flow The flow was first studied to document the base flow without the presence of the cylinder for both laminar and turbulent flow. After confirming a clean laminar base flow in the measurement region, the flow was then subjected to an APG to induce a laminar separation bubble. In figure 6(a), the velocity profile for U=0.132 m s−1 at a Reynolds number, Rex » 1.3 ´ 105 is compared with the theoretical Blasius profile. The theoretical Blasius profile was calculated using Howarth (1938) solution with zero pressure gradients and it gives close agreement with the experimental velocity profile for laminar flow. The theoretical boundary layer thickness, δ, based on Blasius calculation is found to be 13.5 mm in comparison with the experimental value of 13 mm (figure 6(a)). In figure 7, boundary layer thickness (δ), displacement thickness (δ1), momentum thickness (θ) are compared for between Blasius calculation and experimental data at various x-location from leading edge within the measurement region. All the characteristic values from experimental data matches well with Blasius calculations. Also the shape factor (H) from Blasius calculation is found to be 2.59 and matches comparably well with the experimental value of 2.69.

Bioinspir. Biomim. 12 (2017) 016009

F Afroz et al

Figure 13. Instantaneous flow field for laminar flow under same amount of APG at VR=1.409 at upstream location of separation.

Likewise, a clean base flow turbulent boundary layer over the flat plate was ensured before testing under the presence of an adverse pressure gradient. According to (Prandtl 1952), the one-seventh power law is a good approximation for the velocity profile up to Rex=107. In figure 6(b), the experimental velocity profile for U=0.25 m s−1 at Rex » 2.5 ´ 105 is compared with the theoretical one-seventh power law and also one-sixth power law. As flow is tripped to turbulent and the Reynolds number is relatively low, the experimentally measured profile shows a better match with a one-sixth law.

4. Results and discussions 4.1. Laminar separation control 4.1.1. Time-averaged velocity field The time-averaged velocity contours showed that for laminar flow (Rex=1.3×105), the shark skin 10

resulted in a reduction in size of the separation bubble as compared to a smooth plate under the same amount of APG (figures 8(a), (b)). At a higher strength of APG, the separation bubble on the shark skin is again reduced in size (figures 9(a) and (b)). Figure 8 shows the LSB region bounded by mean dividing streamline passing through the separation point. Then the separation point and transition point are identified on this velocity line. From table 1 it can be seen that shark skin minimally delays the separation point, but induces earlier transition to reattach the flow thereby reducing the height and extent of the LSB. It should be noted that LSB height (h) is measured at the point of transition, T, after which height decreases due to turbulent mixing and eventual reattachment. 4.1.2. Time-averaged vorticity field Figure 10 shows the comparison of time-averaged vorticity field between the smooth plate and shark skin

Bioinspir. Biomim. 12 (2017) 016009

F Afroz et al

Figure 14. Instantaneous flow field for laminar flow under same amount of APG at VR=1.409 inside the separation region.

under same amount of APG. The separated laminar boundary layer forms a free shear layer outside the U=0 velocity line and contains the clockwise, negative vorticity from the original boundary layer flow. At this higher strength of APG, the shark skin lowers the extent to which the vorticity layer is pulled away from the wall as compared to that of smooth plate. 4.1.3. Reynolds stress One method to locate the turbulent transition point for a separation bubble is to measure the normalized

( ) where the starting point of

Reynolds stress –

u¢v¢ U2

turbulent transition, T in separated shear layers, is defined as the point where the normalized Reynolds stress

(– ) is more than 0.001 (Volino and u¢v¢ U2

11

Hultgren 2000, Ol et al 2005, Burgmann et al 2006, Hu and Yang 2008). As such the contour levels of Reynolds stress above 0.001 are only shown in figures 11(a) and (b). As seen in figure 11, early transition takes place and higher Reynolds stress occurs for the shark skin cases. For a strong APG (figure 11(b)), shark skin results in comparatively higher Reynolds stress. As also seen in figures 8 and 9, shark skin controls the separation better for the stronger APG case. We thus hypothesize that there might be a threshold value of reverse flow for which shark skin bristling is actuated to better control flow separation; this threshold would be achieved more regularly for the higher APG case. 4.1.4. Instantaneous velocity field The instantaneous velocity history gives more insight as to the laminar flow separation control of shark skin.

Bioinspir. Biomim. 12 (2017) 016009

F Afroz et al

Figure 15. Instantaneous velocity history flat plate Vs shark skin for VR=1.409 at y=2 mm and at different x-positions (a) x=45 mm, (b) x=70 mm.

Figure 16. Contour of backflow coefficient of laminar flow under at Rex » 1.3 ´ 105 and at different amount of APG, (a), VR=1.209, (b) VR=1.409.

The u-velocity history (for t=9.6 s) very near the wall (y=2 mm) at different x positions for two different magnitudes of APG are presented in figures 12 and 15. Here two x-locations are chosen: one just after the beginning of separation (x=45 mm) and the other near the transition zone of separation (x=70 mm). 12

Figure 12(a) shows that the peak magnitudes of reverse flow velocity vary within −0.5 to −1.2 cm s−1 for the flat plate case, whereas for the shark skin cases, most of the peak reverse flow values are less than −0. 5 cm s−1. In figure 12(b), the peak of reverse flow velocity varies within −0.5 to −1.5 cm s−1 for smooth plate cases.

Bioinspir. Biomim. 12 (2017) 016009

F Afroz et al

Figure 17. Contour of backflow coefficient of turbulent flow under at different amount of APG, Rex » 2.5 ´ 105, VR=2.13, (b) Rex » 2.5 ´ 105, VR=2.66, and (c) Rex » 3.0 ´ 105, VR=3.10.

Here also the peak reverse flow value for shark skin surface does not exceed −0.5 cm s−1. For the smooth plate case, the instantaneous flow velocity shows that velocity is negative for most of the time. However, the shark skin causes a higher frequency of flow direction change and a lower peak value of backflow. Figures 13 and 14 present the instantaneous flow field for an interval of 0.2 s over a smooth and shark skin surface at the two strengths of APG. For the smooth plate, the instantaneous shape and size of the bubble does not change much with time. But for the shark skin surface, the instantaneous shape of the bubble is obviously changing with time. It starts with a large scale vortex structure that sheds off with a decrease in size of the separated region, and then the bubble starts growing in size. Instantaneous flow velocity of shark skin under that higher strength of APG is also observed in figures 15(a), (b). At x=45 mm, the peak reverse flow goes up to −3 cm s−1 for the smooth case. The amount of positive velocity increases for shark skin surface compare to the cases in figure 12. But at x=70 mm, the reverse velocity peak goes as high as −5 cm s−1 and it shows more control of separation. At this point, the velocity remains positive for most of the time and the positive velocity peaks are higher in magnitude. 13

4.2. Turbulent separation control 4.2.1. Contour of backflow coefficient The backflow coefficient (χ) is a measurement of the percentage of time the flow is reversed. Figure 16 shows a reduction in the size of the backflow coefficient region on shark skin as compared to the smooth plate under the same amount of APG. Contours of backflow coefficient show that in the ydirection upwards from the wall, the amount of backflow increases to its maximum value and then start decreasing again. Also in the x-direction after a certain distance, the amount of backflow reaches a maximum value and then starts decreasing. Backflow contours also agree with the fact that the smooth wall case is steady laminar flow separation reaching 100% backflow in the front of the separation bubble, while the shark skin case is unsteady. For the cases of shark skin, the maximum amount of backflow is lower than the smooth plate under the same APG. Also in the xdirection, the amount of backflow reaches to a maximum value earlier for shark skin cases and after that starts decreasing. Turbulent boundary layer separation is unsteady in nature and does not have a single stationary separation point like the smooth wall laminar separation bubble. The unsteady nature of separated flow

Bioinspir. Biomim. 12 (2017) 016009

F Afroz et al

Figure 18. Contour of backflow coefficient of turbulent flow under at Rex » 3.0 ´ 105 and at VR=2.65, (a) smooth plate, (b) shark skin.

Table 2. Comparison of turbulent separation over flat plate and shark skin under different amount of APG.

Rex

VR

Smooth plate XS

Shark skin XS

2.5 ´ 105

2.13 2.66

35 30

40 36

3 .0 ´ 105

2.65 3.1

62.5 31

51

forming in a turbulent boundary layer is evident by the contours of back-flow coefficient (χ) (figures 17 and 18). Quantitative definitions as to the detachment state near the wall can be defined based on this backflow (Simpson et al 1981a, 1981b, Simpson 1996). The separation points defined are: incipient detachment which occurs where an instantaneous backflow 1% of the time (χ=0.01) is measured; intermittent transitory detachment occurs for an instantaneous backflow 20% of the time (χ=0.20); finally detachment (D) occurs for an instantaneous backflow 50% of the time (χ=0.50). The backflow coefficient contour shows that shark skin is capable of controlling turbulent separation. Shark skin reduces the size of the backflow coefficient region and results in delayed separation as compared to a smooth plate under the same amount of APG. Detachment points (XS) for smooth plate and shark skins under different amounts of APG are listed in table 2. 14

For the lower strength of APG (figure 17(a)), the amount of backflow reaches as high as 70% on the smooth plate, whereas this value reduces to 50% for shark skin. At higher APG values (figures 17(b), (c) and 16(b)) the shark skin gave better control of separation as the height of the backflow contour and the maximum percentage of backflow reduces significantly. Perhaps at a higher strength of APG, the boundary layer is induced to pull further away from the wall, but the higher magnitude of flow reversal may cause greater bristling of the shark scales thereby inducing greater separation control. 4.2.2. Time-averaged vorticity field The time-averaged vorticity field shows that shark skin gives better control under higher APG as shown in figure 19(b). For lower APG, there is less difference between the pulling of the boundary layer from shark skin as compared to the smooth plate. But at higher APG, the shark skin definitely controls the pulling away of the boundary layer from the wall. 4.2.3. Time-averaged velocity field Turbulent separation is very unsteady in nature and a time-averaged mean flow analysis will not reveal the unsteady, instantaneous flow behavior of the separation bubble (Na and Moin 1998). But if the velocity fields can be accurately averaged then it can give a steady separation bubble. From TR-DPIV analysis the time-averaged velocity are shown in figure 20. The

Bioinspir. Biomim. 12 (2017) 016009

F Afroz et al

Figure 19. Time-averaged vorticity field under at different amount of APG, (a) Rex » 2.5 ´ 105, VR=2.13, (b) Rex » 2.5 ´ 105, VR=2.66.

height of the negative stream wise velocity zone of turbulent separation over the shark skin is smaller than that of the smooth plate for different amounts of APG. For higher strength of APG (figures 20(b) and (c)), a significant reduction in the size of the negative velocity region was observed. 4.2.4. Instantaneous velocity field Figure 21 shows the instantaneous velocity history of a point very close to the wall under different strengths of APG. As the APG increases, the peak of negative flow velocity increase from −0.1 m s−1 to −0.35 m s−1 for the smooth case. And as the magnitude of APG increases, shark skin results in better control of turbulent boundary layer separation. It can be also seen for the shark skin surface under various amount of APG in figure 21 that the periods of backflow are much shorter in duration with lower peak magnitudes, resulting in a significantly smaller total time of reversal. As the shark skin is subjected to an increased strength of APG, the reversed flow region is further reduced from reaching a higher magnitude thereby controlling flow separation.

5. Summary The maximum speed limit of our water tunnel facility is much lower than the actual speed of a real shark. However both in the experiments and real swimming conditions, the scale bristling takes place in bottom 15

5% of boundary layer thickness. Let us consider a real shark swimming at 10 m s−1 and the Reynolds number at the flank region (x=1 m from the nose) would be approximately 1 ´ 107 . The equation of flat plate 0.16x turbulent boundary layer thickness is dturb = 1 and Rex 7

in the real swimming case would be approximately 16 mm. On the flank region, the flexible scales of crown length of 0.22 mm can bristle up to 48°. This gives a scale protrusion height upon actuation of 0.16 mm, which is within the lower 1% of turbulent boundary layer thickness. The tripped turbulent boundary layer thickness in our experiment is close to 31 mm and the protruding scale would reach within 1% of the boundary layer to interact with the buffer layer. An estimate of the laminar boundary layer thickness is 5x given by d lam = Re , and our experimental laminar x

boundary layer thickness is measured approximately as 13 mm. Therefore, the protruding scale would be within the bottom 2% of the boundary layer thickness. Thus even though both flows are not similar in Re, the flow control actuation in both cases reaches into the bottom 1%–5% of boundary layer thickness. The fundamental turbulent production happens in the viscous sublayer and buffer zone of the boundary layer (nominally within 1% of boundary layer) and thus the scales within this lower 1% of the boundary layer are capable of interacting with the reversed flow, hypothesized to occur in the low speed streaks. As the denticles bristle to form cavities between individual scales, the presence of cavities is known to induce strong mixing

Bioinspir. Biomim. 12 (2017) 016009

F Afroz et al

Figure 20. Time-averaged velocity (u/U) contour for turbulent flow separation at different amount of APG, (a) Rex » 2.5 ´ 105, VR=2.13, (b) Rex » 2.5 ´ 105, VR=2.66, and (c) Rex » 3.0 ´ 105, VR=3.10.

events which is highly dependent on the cavity Reynolds number (Rec, calculated based on scale height and freestream velocity). If a shark swims at a speed of 10 m s−1 and the crown length of a scale is 0.22 mm, the Rec would be 2200. But for the present experimental study, Rec for laminar flow is 26.4 and approximately 60 for the turbulent case. As the cavity Reynolds number is very low in the present experimental studies, turbulent mixing in and out of the cavities formed between the bristled scales may not be strongly induced. Thus the primary mechanism of flow control for the experiments is hypothesized to be the scale bristling inhibiting flow reversal. It should be noted also that the shark skin also caused earlier transition to turbulence in the laminar separation bubble. In this case, the instantaneous flow velocity measurements near the wall indicated an unsteady laminar point of flow separation. This unsteady behavior is the result of the shark skin inhibiting the flow reversal near the wall in an unsteady manner as the scales are bristled up during flow reversal events. Also, the time history of the velocity inside the laminar separation region shows a reduction in the peak 16

magnitude of reversing flow for the shark skin compared to the smooth plate under the same APG. For higher APG, the shark skin gives better control of laminar separation and the overall forward velocity is higher in magnitude. In the cases of the turbulent boundary layer separation, the reverse shear is hypothesized to erect the scales and the bristled scales obstruct the reversing flow reducing the overall backflow coefficient. Thus even for turbulent flow, turbulent mixing induced by the presence of the scales may not be overly present in this experiment as a key mechanism to control the flow separation. Rather the reversed flow bristles the shark scales, inhibits backflow and thus prevents the reverse flow from reaching higher magnitudes. The fact that the shark skin controlled the flow separation more effectively for higher APG cases indicates potentially more bristling leading to more flow control as the backflow contour height and the maximum percentage of backflow reduced significantly at larger APG. Overall, our hypothesis that shortfin mako shark skin can control the flow separation has been proven, and a logical hypothesis based on scale bristling has

Bioinspir. Biomim. 12 (2017) 016009

F Afroz et al

Figure 21. Instantaneous velocity history inside the turbulent separation region flat plate versus shark skin for y=2 mm and x=45 mm, (a) Rex » 2.5 ´ 105, VR=2.13(b) Rex » 2.5 ´ 105, VR=2.66 (c) Rex » 3.0 ´ 105, VR=3.10.

been presented. The large difference of Rec suggests that mixing induced by the scales would be much higher for real swimming flow conditions. However, the passive flow separation control mechanism dependent on inhibiting flow reversal is a mechanism somewhat independent of the Re as long as the reversing flow is strong enough toinitiate scale bristling. Therefore, the primary separation control mechanism for lower Rec in our experiments may not be because of turbulent mixing. However, for real shark swimming speeds, or at higher Re, the degree of flow control could be substantially increased as turbulent mixing induced by the cavities could provide an additional mechanism to bring higher momentum flow closer to the surface inside the turbulent boundary layer.

6. Conclusions Shortfin Mako shark skin from the flank region with highly flexible scales was tested under various 17

amounts of APG for its separation control capability. The time-averaged DPIV results showed that shark skin is capable of controlling both laminar and turbulent separation on flat, non-moving surface. This indicates the presence of a passive mechanism for separation control and, although scale bristling could not be observed during experiments, the inhibiting effect on the near-wall flow reversal by the presence of the scales was measured. The presence of the shark skin on the flat plate also results in a smaller separation region and a delay in the separation point under different magnitudes of APG. Instantaneous velocity history also reveals that for a higher strength of APG, shark skin results in better separation control in both laminar and turbulent flow cases. Flow separation degrades the stability and performance of vehicles in a wide range of engineering applications. This newly discovered flow separation control technique of flexible shark scales has the potential to inspire similar man-made microscale surfaces for

Bioinspir. Biomim. 12 (2017) 016009

F Afroz et al

future separation control applications in both air and water.

Acknowledgments Support under NSF grant 0932352 is gratefully acknowledged. First author Farhana Afroz was also supported by a scholarship through the Alabama EPSCoR Graduate Research Scholars Program.

References Afroz F 2014 Laminar and turbulent boundary layer separation control of Mako shark skin Doctoral Dissertation The University of Alabama TUSCALOOSA Afroz F, Sharif M A and Lang A 2016 Numerical study of adverse pressure gradient generation over a flat plate using a rotating cylinder J. Fluids Struct. 62 187–208 Bechert D, Bruse M and Hage W 2000a Experiments with threedimensional riblets as an idealized model of shark skin Exp. Fluids 28 403–12 Bechert D, Bruse M, Hage W and Meyer R 2000b Fluid mechanics of biological surfaces and their technological application Naturwissenschaften 87 157–71 Bechert D, Bruse M, Hage W, van der Hoeven J T and Hoppe G 1997 Experiments on drag-reducing surfaces and their optimization with an adjustable geometry J. Fluid Mech. 338 59–87 Bechert D, Hoppe G and Reif W-E 1985 On the drag reduction of the shark skin 1985 AIAA Shear Flow Control Conf. Bixler G D and Bhushan B 2013 Shark skin inspired low-drag microstructured surfaces in closed channel flow J. Colloid Interface Sci. 393 384–96 Burgmann S, Brücker C and Schröder W 2006 Scanning PIV measurements of a laminar separation bubble Exp. Fluids 41 319–26 Bushnell D B 1985 Turbulent drag reduction for external flows Aircraft Drag Prediction and Reduction AGARD-R-723 Paper No.5 Gad-el-Hak M 1989 Flow control Appl. Mech. Rev. 42 261–93 Gaster M 1969 The Structure and Behaviour of Laminar Separation Bubbles (HM Stationery Office) 1–32 Hart D P 2000 PIV error correction Exp. Fluids 29 13–22 Howard F G and Goodman W L 1985 Axisymmetric bluff-body drag reduction through geometrical modification J. Aircr. 22 516–22 Howarth L 1938 On the solution of the laminar boundary layer equations Proc. R. Soc. A 164 547–79 Hu H and Yang Z 2008 An experimental study of the laminar flow separation on a low-Reynolds-number airfoil J. Fluids Eng. 130 051101 Huang H, Dabiri D and Gharib M 1997 On errors of digital particle image velocimetry Meas. Sci. Technol. 8 1427 Kundu P K and Cohen I M 1990 Fluid Mechanics (San Diego, CA: Academic) 638 Lang A, Motta P, Habegger M L and Hueter R 2012 Shark skin boundary layer control Natural Locomotion in Fluids and on Surfaces (Berlin: Springer)

18

Lang A, Motta P, Habegger M L, Hueter R and Afroz F 2011 Shark skin separation control mechanisms Mar. Technol. Soc. J. 45 208–15 Lang A, Motta P, Hidalgo P and Westcott M 2008 Bristled shark skin: a microgeometry for boundary layer control? Bioinsp. Biomim. 3 046005 Lang A W, Bradshaw M T, Smith J A, Wheelus J N, Motta P J, Habegger M L and Hueter R E 2014 Movable shark scales act as a passive dynamic micro-roughness to control flow separation Bioinsp. Biomim. 9 036017 Lin J C 2002 Review of research on low-profile vortex generators to control boundary-layer separation Prog. Aerosp. Sci. 38 389–420 Motta P, Habegger M L, Lang A, Hueter R and Davis J 2012 Scale morphology and flexibility in the shortfin mako Isurus oxyrinchus and the blacktip shark Carcharhinus limbatus J. Morphology 273 1096–110 Na Y and Moin P 1998 Direct numerical simulation of a separated turbulent boundary layer J. Fluid Mech. 374 379–405 Oeffner J and Lauder G V 2012 The hydrodynamic function of shark skin and two biomimetic applications Journal Exp. Biol. 215 785–95 Ol M V, Mcauliffe B R, Hanff E S, Scholz U and Kähler C 2005 Comparison of laminar separation bubble measurements on a low Reynolds number airfoil in three facilities AIAA Paper 5149 2005 Prandtl L 1952 Essentials of Fluid Dynamics: with Applications to Hydraulics Aeronautics, Meteorology, and other Subjects (Hafner) Raschi W and Tabit C 1992 Functional aspects of placoid scales: a review and update Mar. Freshwater Res. 43 123–47 Simpson R L 1996 Aspects of turbulent boundary-layer separation Prog. Aerosp. Sci. 32 457–521 Simpson R L, Chew Y-T and Shivaprasad B 1981a The structure of a separating turbulent boundary layer: I. Mean flow and Reynolds stresses J. Fluid Mech. 113 23–51 Simpson R L, Chew Y-T and Shivaprasad B 1981b The structure of a separating turbulent boundary layer: II. Higher-order turbulence results J. Fluid Mech. 113 53–73 Sohn K H, Shyne R J and Dewitt K J 1998 Experimental investigation of boundary layer behavior in a simulated low pressure turbine ASME 1998 Int. Gas Turbine and Aeroengine Congress and Exhibition American Society of Mechanical Engineers, V001T01A010-V001T01A010 Stevens J D 2008 The biology and ecology of the shortfin mako shark, Isurus oxyrinchus Sharks Open Ocean: Biology, Fisheries and Conservation ed M D Camhi, E K Pikitch and E A Babcock (Oxford: Blackwell) 87–94 Volino R J and Hultgren L S 2000 Measurements in separated and transitional boundary layers under low-pressure turbine airfoil conditions ASME Turbo Expo 2000: Power for Land, Sea, and Air American Society of Mechanical Engineers, V003T01A065-V003T01A065 Willert C E and Gharib M 1991 Digital particle image velocimetry Exp. Fluids 10 181–93 Zhao D-Y, Huang Z-P, Wang M-J, Wang T and Jin Y 2012 Vacuum casting replication of micro-riblets on shark skin for drag-reducing applications J. Mater. Process. Technol. 212 198–202

Smile Life

When life gives you a hundred reasons to cry, show life that you have a thousand reasons to smile

Get in touch

© Copyright 2015 - 2024 PDFFOX.COM - All rights reserved.