You are on page 1of 30

A seminar report on

FLOW CONTROL ACTIVE AND PASSIVE

GUIDED BY Dr. J BANERJEE

SUBMITTED BY Vineet Maheshwari U07ME679 B.Tech IV Mechanical Engg.

DEPARTMENT OF MECHANICAL ENGINEERING SARDAR VALLABHBHAI NATIONAL INSTITUTE OF TECHNOLOGY ICHCHHANATH, SURAT-395 007, GUJARAT, INDIA

CERTIFICATE

This is to certify that seminar report entitled Flow Control Active and Passive submitted by Mr. Vineet Maheshwari in partial fulfilment of the requirement for the award of the degree of BACHELOR OF TECHNOLOGY IN MECHANICAL ENGINEERING of the Sardar Vallabhbhai National Institute of Technology, Surat is a record of his own work carried out under my supervision and guidance. The matter embodied in the dissertation has not been submitted elsewhere for the award of any other degree or diploma.

GUIDED BY:

Dr. J. BANERJEE Mechanical Engineering Department SVNIT, Surat -395007

TABLE OF CONTENT

1. INTRODUCTION 2. FLOW CONTROL 2.1 HISTORICAL BACKGROUND 2.2 GOALS OF FLOW CONTROL 2.3 CONTROL OF TURBULENCE 2.3.1 MECHANISM OF TURBULENCE 2.3.2 CONTROL METHODOLOGY 3. PASSIVE FLOW CONTROL 3.1 RIBLETS 3.2 VORTEX GENERATORS 4. ACTIVE FLOW CONTROL 5. MIXING ENHANCEMENT 6. CASE STUDIES 6.1 ACTIVE CONTROL OF AN INCOMPRESSIBLE AXISYMMETRIC JET USING FLAPS AND ZERO MASS-FLUX EXCITATION 6.1.1 6.1.2 6.1.3 6.1.4 EXPERIMENTAL SETUP EXPERIMENTAL PROCEDURE AND DATA ANALYSIS RESULTS AND DISCUSSION CONCLUSION

6.2 FLOW CONTROL USING LOCALIZED INDUCTION HEATING IN A VARTM PROCESS 6.2.1 6.2.2 6.2.3 6.2.4 EXPERIMENTAL SETUP EXPERIMENTAL PROCEDURE RESULTS AND DISCUSSION CONCLUSION

LIST OF FIGURES

1. ENGINEERING GOALS AND CORRESPONDING FLOW CHANGES 2. INTERRELATION BETWEEN FLOW CONTROL GOALS 3. CLASSIFICATION OF FLOW CONTROL STRATEGIES 4. RIBLET GEOMETRY 5. (A) CO-ROTATING VORTEX GENERATOR CONFIGURATION (B) COUNTER-ROTATING VORTEX GENERATOR CONFIGURATION 6. DIFFERENT VIEWS OF THE EXPERIMENTAL SETUP 7. AXIAL VELOCITY FIELD AT X/D =1.0 8. AXIAL VELOCITY FIELD, X/D = 0.5, 1.0 AND 2.0, F+ = 0.2 AND C = 0.45 9. AXIAL VELOCITY FIELD, X/D = 1.0, F+ = 0.2 AND C = 0.15, 0.45 AND 0.75 10. STREAMWISE VORTICITY, C = 0.3, F+ = 0.1, (A) = 0O, (B) = 90O, (C) = 180O AND (D) = 270O 11. EXPERIMENTAL SETUP 12. PREFORM LAYUP GEOMETRIES USED IN THE EXPERIMENTAL RUNS 13. FLOW FRONT PROGRESSION FOR THE CENTRAL LOW PERMEABILITY PATCH PREFORM LAYUP FOR (A) UNCONTROLLED (UNHEATED) CASE, AND (B) ACTIVELY CONTROLLED (HEATED) CASE PRESENTED AT TFILL/4 TIME INTERVALS. 14. FLOW FRONT GEOMETRIES FOR THE SOLID INSERT PREFORM LAYUP FOR (A) UNCONTROLLED (UNHEATED) CASE, AND (B) ACTIVELY CONTROLLED (HEATED) CASE PRESENTED AT EQUIVALENT % FILL

ABSTRACT

Today there are a number of situations where it is desirable to manipulate the flow-field by means of some flow control mechanism in order to achieve an engineering goal, such as reducing the drag, enhancing the lift, enhancing the mixing of mass, momentum or energy, suppressing the flow-induced noise or a combination of two or more of these. Strategies employed to achieve these goals include: transition delay/advancement, separation prevention/provocation and turbulence suppression/enhancement. Flow control in turbulence is challenging, but of high value in various engineering areas. There are two types of flow control methods (i) energy-consuming active control methods and (ii) passive methods that do not require energy. Passive control utilises the action of macro level physical structures, like riblets and vortex generators. Active control seems to be a highly promising and versatile method. It alters the turbulence or transient flow by modifying the coherent structures. Active flow control is achieved by creating pressure waves, causing convective flow currents or by using structures incorporating electromechanical devices, like MEMS. Flow control has the biggest application in mixing enhancement. This also forms the basis of flow-induced noise reduction methodologies. The present work talks about the objectives, definition, mechanisms, types and limitations of various flow control methods. Control of turbulence is stressed upon and mechanism of turbulence is described to explain fundamentals of turbulence control strategies. Active flow control is explained in detail, owing to its greater applicability. Two case studies are put forward to further elaborate active flow control methods.

1. INTRODUCTION The ability to manipulate a flow-field actively or passively to effect a desired change is of immense technological importance, and this makes the subject being hotly pursued by scientists and engineers today. Potential benefits of realizing efficient flow-control systems are the saving of billions of dollars in annual fuel costs for land, air and sea vehicles and environmentally more competitive processes involving fluid flows. Many methods are considered to achieve transition delay, separation postponement, lift enhancement, drag reduction, turbulence augmentation and noise suppression. However there are some performance penalties, associated with each method, like cost, complexity and trade-off. 2. FLOW CONTROL It is important to observe that here the term Flow Control is in no way related to flowrate control by means of a manual or automatic valve. Rather it infers to the attempt to alter favourably the character or disposition of a flow-field. 2.1 HISTORICAL BACKGROUND German scientist Ludwig Prandtl pioneered the modern use of flow control in his paper, titled ber Flssigkeitsbewegung bei sehr kleiner Reibung (On Fluid Motion with Very Small Friction) that he presented at the Third International Congress of Mathematicians held at Heidelberg, Germany in 1904. In this paper, Prandtl introduced the boundarylayer theory, explained the mechanics of steady separation and described several experiments in which boundary layer was controlled [1]. Demonstrating active flow control, he used suction to delay boundary-layer separation from the surface of a cylinder. Vigorous research in boundary layer control was pursued in Germany just before and during the Second World War. The oil crisis of 1970s renewed the interest in novel methods of flow control to reduce friction drag even in turbulent boundary layers. In comparatively recent past, the need to reduce the emissions of greenhouse gases and to construct supermaneuverable fighter planes, faster and quieter underwater vehicles and hypersonic transport aircraft provided new challenges to researchers in the field of flow control.

2.2 GOALS OF FLOW CONTROL Flow control is employed to achieve many engineering goals. Some of the typical ones are: reducing the drag; enhancing the lift; enhancing the mixing of mass, momentum or energy; suppressing the flow-induced noise; a combination of two or more of all these. To achieve one or more of these goals, several strategies are employed. They are: transition delay/advancement; separation prevention/provocation; turbulence suppression/ enhancement [1].

Figure 1: Engineering goals and corresponding flow changes [1]

These objectives are not necessarily mutually exclusive. Let us take the example of an aircraft wing to understand this fact. If the boundary layer becomes turbulent, its resistance to separation is enhanced, and more lift could be obtained at increased incidence. On the other hand, the skin-friction drag for a laminar boundary layer can be as much as an order of magnitude less than that for a turbulent one. If transition is delayed, lower skin friction as well as lower flow-induced noise is achieved. However the laminar boundary layer can only support a very small adverse pressure gradient without separation, and subsequent loss of lift and increase in form drag occur. Once the laminar boundary layer separates, a free-shear layer forms and, for moderate Reynolds numbers, transition to turbulence takes place. Increased entrainment of high speed fluid because of the turbulent mixing may result in reattachment of the separated region and formation of a laminar separation bubble. At a higher incidence, the bubble breaks down, either

separating completely or forming a longer bubble. In either case, the form drag increases and the lift-curves slope decreases. The ultimate goal of all this is to improve the airfoils performance by increasing the lift-to-drag ratio. However induced drag is caused by lift generated at higher incidence, but form drag also increases at these angles. This discussion points to potential conflicts as one tries to achieve a particular control goal only to adversely affect another goal.

Figure 2: Interrelation between flow control goals [1]

2.3 CONTROL OF TURBULENCE In order to understand the flow control methodology in turbulent flow, it is first obligatory to understand the mechanism of turbulence. 2.3.1 MECHANISM OF TURBULENCE

Shear flow turbulence is neither homogeneous nor isotropic. It is dominated by a quasiperiodic sequence of large-scale structures often referred to as turbulent coherent structures. Coherent structures are not only quasi-periodic, but are different in size and shape depending on the location of these structures within the flow [2]. Furthermore, the coherent structures are born, grow and die within the boundary layer, evolving in both space and time. Although the precise dynamics involved in the turbulence activities are far from clear, it is widely assumed that the process starts with pairs of elongated, counter rotating, stream-wise vortices buried within the near-wall sub-layer region. These vortices are often referred to as hairpin vortices. These hairpin vortices exist within a strong shear layer and induce low and high speed regions between them. The low speed regions, close to the wall, termed streaks, grow downstream and develop inflectional velocity profiles. At the same time, the interface between the low and high speed fluid begins to oscillate, signalling the onset of a secondary instability. The low speed region lifts up away from the wall as the oscillation amplitude increases and the flow rapidly breaks down (bursts) into a completely chaotic motion. This process is self-generating resulting in the continuous cyclic propagation of near-wall hairpin vortices, streaks and bursts. 2.3.2 CONTROL METHODOLOGY

By comparison with laminar flow control or separation prevention, the control of turbulent flow remains a very challenging problem. Brute force suppression on turbulence via active, energy consuming control strategies is always possible, but the penalty for doing so often exceeds any potential benefits [2]. The artifice is to achieve a desired effect with minimum energy expenditure. Delaying laminar-to-turbulence transition to reasonable Reynolds numbers and preventing separation can readily be accomplished using many passive and predetermined active control strategies. Yet very few of the classical strategies are effective in controlling free-shear or wall-bounded turbulent flows. For Example, in an attempt to reduce the skin-friction drag of a body having turbulent boundary layer using global suction, the penalty associated with the control device often exceeds the saving derived from its use. Flow control is most effective when applied near the transition or separation points [3]; in other words, near the critical flow regimes, where flow instabilities magnify quickly and therefore delaying/advancing laminar-toturbulence transition and preventing/provoking separation are relatively easier tasks.

Most of the activity in turbulent flow control, both active and passive, relies on the targeted manipulation by suppression, enhancement or modification of the quasi-static coherent structures [3]. As a result of modern innovations in electronics, like microfabrication, it is possible to develop inexpensive, programmable sensor/actuator chips that have dimensions of the order of a few microns. These chips, with the help of modern computing tools, are potentially useful for constructing effective adaptive controllers. Such a system can sense and target individual coherent structures by means of various active flow control techniques. However some passive flow control methods are also effective at times.

Figure 3: Classification of Flow Control strategies

3. PASSIVE FLOW CONTROL In passive flow control, flow is affected without expending energy in the control process. Normally the passive control does not employ a feedback loop of the transition detection and manipulation. There are a number of passive control methods. Two of the most widely used ones are riblets and vortex generators. 3.1 RIBLETS These are inspired by tiny ridge-like structures on the scales of a sharks skin. They reduce skin friction drag from turbulent boundary-layer flow. On the fuselage of an aircraft, they can be seen as small longitudinal grooves, aligned in the flow direction, with typically sharp ridges separating the valleys [4]. They are known to reduce skin friction by up to 10% [3]. However geometry of riblet structure is very important to determine its drag reduction ability. Bechert and Bartenwerfer (1989) suggested guidelines for an optimal riblet geometry: (i) a sharp wedge for the rib, preferably with a radius of curvature smaller than 0.5-1% of the lateral rib spacing; (ii) a rib wedge angle as small as possible; and (iii) a valley depth of about 60% of the lateral rib spacing. The same authors state as their basic hypothesis for the drag reduction mechanism "that sharp ribs impede the instantaneous cross-flow in the viscous sub-layer which is generated by the turbulent motion. In this way the whole turbulent momentum exchange in the boundary layer is reduced, which is equivalent to a shear stress reduction".

Figure 4: Riblet Geometry

3.2 VORTEX GENERATORS The control of the flow around an airfoil at high angle of attack is also of strong interest. Such a flow is encountered in several flight phases (take-off, landing and manoeuvre) and can lead to partial or large separation, with the well known consequences (vibration, drag increase and loss of control). To delay and even suppress this separation, it is necessary to bring momentum into the boundary layer (BL) in order to allow it to sustain the strong adverse pressure gradient (APG) [3]. The idea is thus to enhance the exchange of momentum between the boundary layer and the outer flow with the help of optimal coherent structures, triggered by an optimal actuator. Induced vortices are useful in creating such an effect. To create these vortices, vortex generators are employed. These

generators are thin plates of triangular or trapezoidal shape, placed normal to the surface and at a lateral angle to the flow. Vortex generators may be arranged in either co-rotating or counter-rotating configuration, as per the requirement. In their downstream region, they produce arrays of vortices that may be co-rotating or in pairs of counter-rotating vortices. These induced vortices transport low momentum fluid upward (away from the wall) and high momentum fluid downward.

Figure 5: (a) Co-rotating vortex generator configuration, (b) Counter-rotating vortex generator configuration.

4. ACTIVE FLOW CONTROL Active control seems to be the most promising and versatile flow control method. Some amount of energy is expended in active control, but when used with proper optimisation, it can far surpass the benefits of passive control. It must always be kept in mind that the penalty in terms of cost does not exceed the benefit obtained with active control. Active control of a turbulent flow is possible by periodic excitation at one or more frequencies with positive (amplitude) growth rate. The control behaviour is determined and characterized by the frequency, the amplitude, the phase position and the waveform, which controls the content of higher harmonics [1]. The physical mechanism of excitation is in the modification of the coherent structures. Thus excitation at 'low' frequencies (near the local characteristic passage frequency of the coherent structures) acts directly on their strengthening and stabilization, causing local increase and spread in the region of their formation and decrease in the stabilization range. Conversely, 'high' frequency forcing acts on the coherent structures of higher frequency and small dimension, i.e. those near the trailing edge (in any case upstream of the region to be influenced). There it causes either premature transition or at least fast growth with subsequent decay and only asymptotic recovery. High frequency excitation is often applied where suppression of turbulence activity is asked for.

Technical principle of producing and introducing periodic excitation is based on the modification of the concentrated vorticity sheet near the trailing edge (or nozzle), where the receptivity of the flow is at its maximum. This is possibly done by one of the many active control mechanisms. A few of them are: Pressure waves by (i) acoustic means, e.g. a loudspeaker, which may be driven externally by a frequency generator, or by the flow itself, providing an external feedback loop or (ii) by a resonance mechanism within or outside of the flow regime (e.g. the test section) [5]. The important parameter here is the fluctuating pressure gradient at the trailing edge. Shedding/modulating of vorticity (mechanically activated ribbon or flap), vibration (or rotation) of the trailing edge (nozzle) [5]. Suction and injection of fluid at critical points. Heating fluid to result in desired convective currents. Use of Micro Electro Mechanical Systems (MEMS) for sensing turbulent or laminar behaviour of flow and responding accordingly with actuator signalcontrolled motion of specialized structures, with feedback mechanism.

While talking about the mechanism of active flow control, the following points must be kept in mind [5]: The efficiency of excitation is locally determined and it is limited by the saturation behaviour of the structures to be influenced, depending essentially on the amplitude and the mode of excitation frequency. The location of maximum excitation effect depends on the frequency (usually at larger x for lower frequencies), being a function of the characteristic frequency of the neutral flow. The generally limited local extent of the controlled region may be extended by excitation of more than one frequency. The effect depends then on the individual amplitudes of the two frequencies and on their phase relation. Exciting a flow at certain (instability) modes or their combinations may produce special effects.

5. MIXING ENHANCEMENT One of the most important engineering goals of flow control is mixing-enhancement. In nature, the efficient turbulence mixing is responsible for the nearly uniform distribution of oxygen, carbon dioxide and internal energy in the earths atmosphere and oceans. Without that more-or-less uniform distribution, seasonal and latitudinal temperature changes would be even more extreme. Oxygen would be more concentrated on the equatorial plane, and carbon dioxide would be more concentrated in industrial and urban centres. The subject of mixing enhancement is also very important to entire industries. Flow control can be used to augment mixing in both free-shear and wall-bounded flows by increasing the effective area through which transport takes place, by setting off resonant flow instabilities, by advancing laminar-to-turbulence transition, and by enhancing the turbulence once the shear flow is already turbulent [1]. The effective generation of secondary flows, recirculation zones, or flow unsteadiness is an additional tool to enhance mixing in both laminar and turbulent flows. Successful examples of this strategy include the use of vortex generators on airplane wings and coiled tubes in heat exchangers. There is a drag penalty associated with increasing the surface area as well as with all other methods used to enhance mixing, and attempts must be made to minimise this penalty. Many mixing enhancement strategies are used today. Some of them are: Increase in the surface area of the solid using external and internal fins. Set off resonant flow instabilities. Advance lamina to turbulent transition Enhance turbulence Reynolds stresses Induce appropriate secondary flows, recirculation zones, or flow unsteadiness in both laminar and turbulent regimes. Effect chaotic advection in laminar flows, as early transition is not possible in this case.

6. CASE STUDIES In order to understand the mechanism of active flow control, it is important to study a few cases where it has been demonstrated to deliver desirable effects. Here I present some research work that makes active control mechanism easier to understand. 6.1 ACTIVE CONTROL OF AN INCOMPRESSIBLE AXISYMMETRIC JET USING FLAPS AND ZERO MASS-FLUX EXCITATION (SINGH ET AL, AIAA PAPER 2010-4417 [6]) I contributed, as a summer intern, in these investigations at Institut fr Strmungsmechanik und Technische Akustik (ISTA), Technical University Berlin, Germany. An active flow control method of an axis-symmetric jet was investigated which, when activated, generated stream-wise vortices and thus enhanced mixing of the jet flow with the ambient. The perimeter of the jet was equipped with six small flaps deflected away from the stream. Zero mass-flux perturbations were being used to excite the flow. These excitations were introduced in the flow through slots at the base of the flaps. Each of the flaps could be excited independently. In these investigations, the effect of an array of six individually controllable flaps on the global jet behaviour was addressed. Each of the flaps could be excited in phase or with pre-fixed phase shift. Effects of frequency and amplitude on the flow momentum, stream-wise vorticity, circulation and turbulence for a fixed flap deflection angle were part of the investigation. A stereo-PIV setup was used to acquire complete flow field information. The emphasis was placed on mapping the development of the trailing vortices in order to quantify the mixing achieved. 6.1.1 EXPERIMENTAL SETUP

The wind tunnel used for the experiments was a low speed, circular cross-section, open circuit tunnel with an open-air jet. The exit jet diameter was 90mm and it gave a maximum Reynolds number of 90,000 based on jet diameter. The lip of the axisymmetric jet was equipped with six small flaps deflected away from the stream at an angle of 30. The chord length of the flaps was 15mm. The flaps incorporate a flow control slot (15x1.5mm) and each slot was connected to a speaker via a flexible tube. A sine-wave was supplied to the speakers to produce the desired frequency and amplitude through which zero mass flux excitation was introduced to the flow in the axial direction. The control slot had been calibrated to get the desired amplitude of the excitations. The velocity measurements were carried out using stereo-PIV in planes perpendicular to the axis of the jet at axial locations of x/D = 0.25, 0.5, 1.0, 2.0.

(a)

(b)

(c)
Figure 6: (a), (b) and (c) show different views of the experimental setup.

6.1.2

EXPERIMENTAL PROCEDURE AND DATA ANALYSIS

Complete flow field measurements were carried out using a stereo-PIV setup, at a Reynolds number of 31000. The amplitude of the excitation was quantified by the nondimensional parameter C, the momentum coefficient. It is defined as the ratio of the momentum added by the control slot to the momentum of the main jet. The calibration of the excitation amplitude was carried out using a hotwire, which was positioned directly in front of the slot, oriented parallel to it. For each excitation frequency, the peak velocity of the jet was determined as a function of the AC voltage of the excitation signal supplied to the speakers. The amplitude of the excitation was varied in a range previously determined by earlier works. Within this range, a frequency scan was carried out with reduced frequency F+ (dimensionless excitation frequency). The stereo-PIV acquisition was phase-locked with the actuator signal and data at 16 different phases were acquired. The software used for data acquisition and processing was VidPiv, developed by Intelligent

Laser Application, GmbH. Further processing and analysis of results were done with previously developed scripts in MATLAB. 6.1.3 RESULTS AND DISCUSSION

Figure 7 shows the velocity field at x/D =1.0 in the absence of any control. The axial velocity profile resembles an unaltered jet, showing that there is no effect of the presence of slots on the jet.

Figure 7: Axial velocity field at x/D =1.0

Figures 8- (a), (b) and (c) show development of flow over three axial locations x/D = 0.5, 1.0 and 2.0 respectively at F+ = 0.2 and C = 0.45. It can be seen that the introduction of excitation causes the flow near the flaps to get distorted and move towards the flaps at six flap locations, one of them marked as P in the first figure. As we move downstream, the distortions begin to appear in regions between the flaps as well. This indicates the presence of a symmetrical system of stream-wise vortices. On going away further, these regions of distortion begin to merge rapidly.

Figure 8: Axial velocity field, x/D = 0.5, 1.0 and 2.0, F+ = 0.2 and C = 0.45

Figure 9: Axial velocity field, x/D = 1.0, F+ = 0.2 and C = 0.15, 0.45 and 0.75

Figure 9 shows the effect of change in amplitude of excitation. Cases (a), (b) and (c) represent C = 0.15, 0.45 and 0.75 respectively at F+ = 0.2 and x/D = 1.0. It can be easily seen that increase in amplitude of excitation increases the deflection of flow towards the flaps but this effect saturates at still higher amplitudes.

Figure 10: Streamwise vorticity, C = 0.3, F+ = 0.1, (a) = 0o, (b) = 90o, (c) = 180o and (d) = 270o

Figure 10 shows strength of stream-wise vortices at C = 0.3, F+ = 0.1 at four phase angles. At 0o, two pairs of counter-rotating vortices can be observed near each flap. At 90o, a circular system of vortices is formed in the shear layer, with the pairs close to the flaps travelling into the ambient air. At 180, the vortices that were previously observed have moved out further and have already dissipated to a large extent. At 270o, strong pairs could be seen near the flaps. 6.1.4 CONCLUSION

Active flow control using zero mass-flux excitation is applied to study mixing characteristics in a circular jet equipped with six finite span flaps along its perimeter. The above discussions result in following main conclusions: Zero mass-flux excitation parallel to the flow is effective in attaching the flow to the flaps and also in generating stream-wise vortices. The location as well as the strength of these vortices strongly depends on the excitation frequency. The effect of increasing the excitation amplitude saturates at higher amplitudes.

6.2 FLOW CONTROL USING LOCALIZED INDUCTION HEATING IN A VARTM PROCESS (R. J. JOHNSON ET AL; ELSEVIER, 2007 [7]) Voids formed during the mould filling stage of the vacuum assisted resin transfer moulding (VARTM) process become defects in the fabricated parts. Active flow control is one way to eliminate these defects by guiding the flow along a desired path during the mould filling stage of the process. A flow front following control strategy was implemented in a lab-scale experimental setup and tested on several pre-form layups exhibiting spatial permeability variation, as well as in the case of pre-forms with mould inserts. Results of these studies demonstrated that active flow control is capable of reducing the fill time, improving the flow front uniformity throughout the duration of the mould fills, and eliminating dry spot formation.

6.2.1

EXPERIMENTAL SETUP

The VARTM setup used in this study consisted of many integrated components as shown by the schematic in Fig. 11(a). The mould was a 12 in. square with a line inlet on one end and a line vent on the other. The mould was made of Plexiglas, with the line inlet connected by tubing to a resin container, and the line vent connected to a resin trap. Vacuum was pulled on the resin trap by a ColeParmer venturi pump supplied with pressurized shop air through a filter and regulator. This arrangement was found to be capable of drawing a maximum sustained vacuum of _77 kPa. The experiments described below were conducted using a glycerine and water mixture as the working fluid in lieu of an actual resin-catalyst system. The viscosity of the mixture measured using a Brookfield rotational viscometer was 0.7 Pa s at 25oC. This viscosity was chosen to match the viscosity of Shell Epon resin 828 with Shell Epicure 3274 catalyst and Shell HELOXY modifier 505. The viscosity of glycerine, however, was found to be less temperature sensitive than that of the resin mixture. This substitution of working fluids therefore gives conservative estimates of the effects of local heating during the VARTM process. The experiments were conducted using two types of pre-form: Owens-Corning M8610 continuous strand mat and a 7 oz/yd2, 13X13 tows per in. woven fibreglass mat. The woven mat had a permeability that was significantly lower than that of the random mat; the permeability differential was used to create a relative flow lag in the test cases reported in this work. Fig.11(b) shows an expanded view of the mould cross-section: the bottom layer is the Plexiglas mould mentioned above, the upper layer is the vacuum bag used to create the second mould side, in between these layers, the fibrous pre-form materials sandwich a woven carbon fibre layer that forms the susceptor for induction heating. Flow front monitoring during mould filling was accomplished through the use of a Panasonic gray scale CCD camera, mounted under the transparent Plexiglas mould and connected to a personal computer for data acquisition and control in LabVIEW, via a National Instruments IMAQ frame grabber. Gray scale threshold levels were used to process the captured images and to determine flow front positions. Flow front visualization was assisted by the use of a coloured die in the resin to enhance the contrast

between the filled and unfilled regions of the mould. An Ameritherm NOVA 1.0 induction heater was used to locally heat the resin flow. The heating station and induction coil setup were suspended above the vacuum bag side of the VARTM setup on a Micos xy positioning stage assembly. The positioning stages were powered by stepper motors connected to a National Instruments stepper motor driver and National Instruments motion control computer card, and were controlled using the LabVIEW software. This arrangement allowed for precise positioning of the induction coil to any point over the mould allowing for localized heating of the flowing resin. A range of scanning speeds were achieved using the positioning stage and motion control, the fastest of which being much faster than the mould filling rates which allowed for active control of the mould filling process.

Figure 11: Experimental setup

6.2.2

EXPERIMENTAL PROCEDURE

The active control of flow using localized induction heating in the VARTM process has two main requirements: power control and motion control. Induction coil power must be varied in real time to provide significant viscosity reduction for flow permeation in low permeability areas while ensuring that material temperatures are limited to within a prescribed value so as to minimize the effects of the cure reaction if an actual resin system were used. The induction coil must also be moved to insure that heating is applied to appropriate regions of the moulded part. In these studies, only x and y (in plane) motions were considered, while the induction coil remained fixed in the z-direction so as to maintain the desired constant coil susceptor gap during moulding of the flat part. A requirement of the heating control method considered here was that the material temperatures must be limited to prevent premature gelation of the resin during the filling stage of the process. To this end, an upper bound for material temperature was specified as 100oC in this study, with a view to limiting the degree of cross-linking. Control of the motion of the coil in the y-direction (along the overall flow direction) was such that the induction coil always followed the flow front, lagging just enough to not heat any unfilled areas. Heating ahead of a low permeability area could cause the flow to follow the path of least resistance and go around the low permeability area entirely, thereby potentially leading to a dry spot; heating just behind the flow front could reduce the possibility of encountering this problem. The desired motion of the coil in the x-direction (width-wise direction) was to supply heating to areas lagging behind the mean flow front. The exact positioning of these areas was determined by the preform layup, but in order to insure that the control could handle real-time variability in perform permeability, which are not predictable, this coil position needed to be determined from the flow front locations sensed in the mould. The flow front information captured by the CCD camera was used to calculate the percent fill in the y-direction at several discrete locations across the width (x-direction) of the mould. The location of the minimum calculated value, or equivalently the maximum lag, was taken as the desired x location of the coil for that control interval.

6.2.3

RESULTS AND DISCUSSION

Experiments were performed on nine different perform layups presented in Figs. 12(a) (i). All nine layup configurations have overall dimensions of 30.48cmX30.48cm. Experiments were performed on the preform layups for both controlled (heated) and uncontrolled (unheated) cases. Flow front information was collected in each experiment at 5 s intervals in the form of binary image captures.

Figure 12: Preform layup geometries used in the experimental runs

An example of the collected data is presented in its raw format in Fig. 13 for the preform layup presented in Fig. 12(e). The Figure shows the preform layup in the first frame and the progress of the mold filling in the remaining five frames at even intervals of time. The top portion represents the unheated experiment, while the lower portion represents the corresponding actively controlled experiment with localized induction heating on demand. Note that the fill time, tfill, is different for the two experiments.

Figure 13: Flow front progression for the central low permeability patch preform layup for (a) uncontrolled (unheated) case, and (b) actively controlled (heated) case presented at tfill/4 time intervals.

In both experiments the flow starts from the line inlet and progresses towards the central low permeability (woven mat) patch as indicated by the black areas. A lag develops as the flow enters the low permeability region; this lag continues to grow as the flow passes through the woven mat area. In the unheated case, the flow lag is sufficiently large such that the flow pinches off a dry spot at the exit side of the low permeability area. In the heated case, this dry spot is eliminated via active control whereby the resin saturated preform is appropriately heated to reduce the local resin viscosity sufficiently to decrease the flow lag. This comparison shows the distinct advantage of an active flow control scheme in preventing the formation of dry spots, which are defects in the fabricated composite parts.

Figure 14: Flow front geometries for the solid insert preform layup for (a) uncontrolled (unheated) case, and (b) actively controlled (heated) case presented at equivalent % fill.

Fig. 14 shows different flow shapes at the ending time and how they correspond to % fills and non-dimensional error values. The heated and unheated fills of a preform layup with a sold insert are depicted at fill percentages of 70, 80, 90 and at the end of fill. Here, the solid insert in a continuous strand mat layup is designated by the hatched area. As the flow, indicated in black, progresses around the solid insert, it takes on a pronounced V-shaped lag that remains until the ending fill in the unheated case; however, in the heated case this V-shape lag is quickly reduced and eliminated by the ending fill. Note that the error at the end of fill in the unheated case is considerable and in the heated case is almost zero. Also the ending percent fill of the unheated case is close to 90, whereas in the heated case it is close to 100%.

6.2.4

CONCLUSION

Results show that improved flow patterns can be achieved in heterogeneous and complex geometry layups through the use of active control. The use of the active control was further demonstrated to avoid entrapment of dry spots within the preform as well as improve the flow shape as the flow reaches the end of the mould where further voids are commonly formed. In addition to improving flow uniformity, the control was shown to significantly reduce fill times in heterogeneous preform layup cases, compared to the fill times in the absence of any control.

REFERENCES

[1]

Flow control: passive, active and reactive flow management, Mohamed Gad-elHak, Cambridge University Press, 2000. Mohamed Gad-el-Hak, The Taming of the Shrew: Why is it so difficult to Control Turbulence?, Active flow control: papers contributed to the conference Active Flow Control 2006, Berlin, Germany, September 27 to 29, 2006. Clyde Warsop, Current status and prospects for turbulent flow control, Aerodynamic drag reduction technologies: proceedings of the CEAS/DragNet European Reduction Conference, 19-21 June 2000, Potsdam, Germany. P. R. Viswanath, Aircraft viscous drag reduction using riblets, Progress in Aerospace Sciences, Volume- 38 (2002), pp. 571-600. H.E. Fiedler, H. H. Fernholz, On management and control of turbulent shear flows, Progress in Aerospace Science, Volume- 27 (1990), pp. 305-387. Singh, Y., Mueller-Vahl, H., Greenblatt, D., Nayeri, C.N., Paschereit, C.O., Active Control of an Incompressible Axisymmetric Jet using Flaps and Zero Mass-flux Excitation, AIAA Paper 2010-4417, 2010. R. J. Johnson, R. Pitchumani, Flow control using localized induction heating in a VARTM process, Composites Science and Technology (Elsevier), Volume- 67 (2007), pp. 669-684.

[2]

[3]

[4]

[5]

[6]

[7]

ACKNOWLEDGEMENT

I feel it as a great privilege in expressing my deepest and most sincere gratitude to my supervisor, Dr. J. Banerjee for his valuable suggestions and guidance during the seminar work period, without which this work would not have been accomplished. I would also like to thank all the professors and other non-teaching staffs for their kind help in carrying out this work. Last but not the least; I would like to thank the world-wide researchers working in the field of Flow Control who have done pioneering work in the field on which the seminar work is based. I am honoured to be provided with this excellent opportunity. My experience while the seminar work was amazing. It is one of those which I will certainly never forget. It was a great opportunity to research on a topic in which I had interest through academic and other readings but had never got a chance to do. The seminar work has no doubt helped me explore in greater depths the field of Flow Control. It has further strengthened my bondage to the field.

You might also like