You are on page 1of 17

1.

Introduction

The re-emergence of the wind as a significant source of the worlds energy must rank as one of the significant developments of the late 20th century. To understand what was happening, it is necessary to consider five main factors. First of all there was a need - An emerging awareness of the finiteness of the earths fossil fuel reserves as well as of the adverse effects of burning those fuels for energy had caused many people to look for alternatives. Second, there was the potential - Wind exists everywhere on the earth, and in some places with considerable energy density. Wind had been widely used in the past, for mechanical power as well as transportation. Third, there was the technological capacity - In particular, there had been developments in other fields, which, when applied to wind turbines, could revolutionize the way they could be used. These are the three main factors which foster the re-emergence of wind energy, but not sufficient. There needed to be two more factors, first of all a vision of a new way to use the wind, and second the political will to make it happen. At the beginning of winds re-emergence, the cost of energy from wind turbines was far higher than that from fossil fuels. With the advance in technology as a result of the continuous research over two decades makes wind energy as a mature and reliable energy technology. 1.1. History The worlds oldest windmills had a vertical axis of rotation. Braided mats were attached to the axis. The mats caused drag forces and, therefore, were taken along with the wind. In Persian windmills, an asymmetry was created by screening half the rotor with a wall. This way the drag forces could be utilised for driving the rotor(Fig.1.1a).

Fig1.1a) Persian windmill [1] b) Chinese windmill [2] In Chinese windmills - which also date back a long time -, a similar asymmetry is created by sails which rotate out of the wind on their way back, i.e. when they advance into the wind ( Fig.1.1b). These Chinese drag wheels date back to approx. 1000 AD. Similar to the Persian mills, they had a vertical axis and used braided mats as sails. However, in contrast to the Persian mills, they had the typical advantage of vertical axis windmills to utilise the wind independent of its direction. The windmill of Veranzio (Fig.1.2b) belongs to drag driven rotors with a relatively low tip speed ratio. The simplicity of this construction can be appreciated in Fig.1.2a which shows a later version of a vertical axis mill with flapping sails: The millstone is attached directly to the vertical drive shaft without redirecting the rotational movement and without an intermediate gear.

Fig.1.2 a) later version of a vertical axis mill with flapping sails [3] b) windmill of Veranzio[2] The simplicity of the vertical axis design is central to the Savonious rotor (1924,Fig.1.3a) and the Darrieus rotor (1929, Fig. 1.3b). But as late occidental versions of the vertical axis principle they utilise partially or exclusively - the lift force as their driving power.

Fig1.3 a) Savonius rotor [4] b) Darrieus rotor [5] The more recent horizontal axis windmills, such as the Dutch smock mills (Fig. 1.4), designed for a higher tip speed ratio, require a far more sophisticated construction not only for redirecting and back-gearing the rotational movement from the horizontal to the vertical axis, but also for the much more complicated bearing of the faster and heavier horizontal shaft. The oldest construction of a lift-driven horizontal axis device is the post windmill. The post mill consists of a timber support holding the vertical central post around which the boxlike buck (i.e. the mill house) turns on a pivot, Fig.1.5. Using a tail pole, the buck together with the rotor was oriented into the wind. The main shaft with the rotor is almost horizontal. The brake wheel drives via the lantern gear the vertical shaft with the millstone. Only from the 19th Century onwards, post windmills were equipped with two lantern gears for parallel milling operations of two sets of millstones. The post windmill was exclusively applied for grinding grain. First attempts made to use the wind energy to drive pumps occurred in 15 th century. The post mill had to be modified for that purpose. The driving power of the wind had to be transmitted to the pump that was

situated under the mill. The result was the wipmolen which was first used about 300 years after the post mill was first documented, and these were especially designed for drainage purposes. The revolving mill house of a wipmolen contains only the gearbox (Fig.1.6). The actual machine, e.g. a scoop wheel or Archimedean screw, is located below the pyramid-shaped support. The driving shaft is fitted through the hollow post . Later on, also grain mills were built using this principle. There is the obvious advantage of having the set of stones on the ground because no longer heavy loads, like millstones and sacks with grains and flour, had to be carried up and down in the mill house.

Fig.1.4 Dutch smock mills [1] 1 fantail; 2 gear wheel with brake; 3 gear for cap rotation; 4 rollers; 5 wallower; 6 main shaft; 7 sack hoist; 8 great spur wheel; 9 spindle drive; 10 millstone crane; 11 millworks with chute; 12 brake chain; 13 stone adjustment; 14 flour chute

Another mill type that was popular in southern Europe was the tower mill. The first wind mills of this kind were used for irrigation. The first documentation of these mills dates back to the

Fig.1.5 Post mill [1] 1 gear wheel with brake, 2 shaft for sack hoist, 3 hand-driven hoist, 4 rotor shaft, 5 lantern gear, 6 quant, 7 hopper, 8 millstones, 9 traverse beam, 10 brake lever, 11 brake rope, 12 hoist operating rope, 13 floor for flour, 14 saddle, 15 tailpole, 16 central post, 17 sack hoist, 18 quarter bars, 19 cross trees, 20 foundation 13th Century [8] Later versions ,had a turnable wooden cap and a four-bladed wooden rotor like the post mills. The turnable cap is the main characteristics of the Dutch smock mill which came into use in the 16th Century (Fig. 1.4).It is a further development of the tower mill as the lighter wooden construction of the octagonal tower could be easier erected on the wet Dutch marshland than the heavy stone construction of the tower mill. They were used for many purposes like drainage of the polders, to lift the water and grinding grain. A somewhat exotic development is the 17th century Paltrock mill which shows that wind energy can be utilised universally as a driving force. The whole mill (as the cap of the Dutch smock mill) rested on a live ring. This way an entire sawmill can be driven by a windwheel (Fig. 1.7). The last type of historical windmills is the American farm windmill (Western mill), which was developed in the mid-19th century. The Western mill was mainly used for providing drinking water for both people and cattle. The main characteristics of this windmill is the is the rotor rosette of a diameter between 3 and 5 m, with more than 20 metal sheet blades, situated on top of a metal lattice tower. It uses a crank shaft to drive a piston pump. The Western mill was the first windmill type with a fully automatically controlled yaw system including a storm control.

Fig1.6. Wipmelon mill [6]

Fig.1.7. Overview of historical horizontal axis windmill types 1.2. Principle behind wind energy conversion The original source of the renewable energy contained in the earths wind resource is the sun. Global winds are caused by pressure differences across the earths surface due to the uneven heating of the earth by solar radiation. The spatial variations in heat transfer to the earths atmosphere create variations in the atmospheric pressure field that cause air to move from high to low pressure. There is a pressure gradient force in the vertical direction, but this is usually cancelled by the downward gravitational force. Thus, the winds blow

predominately in the horizontal plane, responding to horizontal pressure gradients. At the same time, there are forces that strive to mix the different temperature and pressure air masses distributed across the earths surface. In addition to the pressure gradient and gravitational forces, inertia of the air, the earths rotation, and friction with the earths surface (resulting in turbulence), affect the atmospheric winds. The influence of each of these forces on atmospheric wind systems differs depending on the scale of motion considered. Wind turbine power production depends on the interaction between the rotor and the wind. The power associated with wind is its kinetic energy, and it is proportional to the air density , the cross sectional area A (perpendicular to v), and the third power of the wind velocity v. (1.1) The power of the wind is converted into mechanical power of the rotor by deceleration of the flowing air mass. On one hand, it cannot be converted completely, since this would decelerate the mass flow to zero, it would block the cross sectional (rotor) area A for the following air masses. On the other hand, if the air flows through the area without any deceleration of the wind velocity, there would be no power conversion as well. So there must be an optimum of wind energy conversion through flow deceleration between these two extremes. The maximum power is extracted by a free (i.e. unshrouded) wind turbine if the original upstream wind velocity v1 is reduced to a velocity v3 = v1/3 far downstream the rotor [8][10]. Then, the resulting velocity in the rotor plane is power extraction, the result is (1.2) with the maximum power coefficient . Even in this best case of power extraction . In that case of a theoretically maximum

without any losses, only 59 % of the wind power is extractable. 1.2.1. Drag driven rotors

The drag devices utilise the force that acts on an area a perpendicular to the wind direction, drag force. (1.3) The drag coefficient cD is the proportional constant and describes the aerodynamic quality of the body: the higher the aerodynamic quality of a body, the lower is cD and thus the corresponding drag force.

Fig.1.8 Drag force on a plate and drag coefficients c D for some typical bodies The drag force on a rotating plate is given by (1.4) Where the actual relative velocity, w = v - u at the rotating plate is the difference between wind velocity v and

the circumferential velocity u = RM at the mean radius RM of the area A . The angular velocity due to the rotational speed n(rps) is =2n.

Fig.1.9 Simplified model of drag driven rotor The mean driving mechanical power, { ( ) } (1.5)

The expression in the braces is equal to the power coefficient CP, the aerodynamic efficiency of the rotor. It gives the portion of wind power which is converted into mechanical power. This coefficient must be lower than the theoretical maximum value cP.Betz=0.59 determined by Betz. It depends on tip speed ratio = u / v, At complete standstill (= 0) no mechanical power is extracted from the wind. Neither it is at idling with maximum rotational speed ( = 1), where the circumferential velocity is equal to the wind velocity. In between these extreme cases, the maximum power coefficient CP.Max 0.16 is reached at a tip speed ratio of about opt0.33. Merely 16% of the wind energy can be converted to mechanical energy. 1.2.2. Lift driven rotors

In airfoils the force resulting from the attacking flow has not only a drag component F D parallel to the direction of flow velocity w but also a component FL perpendicular to it, Lift force. (1.6) For small angles of attack A the lift force FL acts at approximately a quarter of the cord length c behind the leading edge.

Fig.1.10 Lift force and drag force of an airfoil and the corresponding coefficients versus angle of attack. For small angle of attack (A 100) the lift coefficient CL and the lift force (FL) is directly proportional to the angle of attack.

(1.7) There is also a drag force FD existing, but it is very small for good quality aerodynamic profiles in the range of a small angle of attack. All horizontal axis wind turbines, like the post windmill, the Dutch smock mill, are driven by the lift principle. Their power coefficients are in the range of CP.Max 0.25 and therefore significantly higher than the maximum values of the drag driven rotors. Modern horizontal axis wind turbines with good aerodynamic profiles (which show small drag coefficients) reach power coefficients up to CP.max 0.5. So they are already very close to the limit value of CP.Betz = 0.59 [9][10]. 2. Wind turbine aerodynamics. 2.1. Aerodynamic modelling Wind turbine aerodynamics concerns the modelling and prediction of aerodynamic forces, such as performance predictions of wind farms, and the design of specic parts of wind turbines, such as rotor -blade geometry. Aerodynamics is normally integrated with models for wind conditions and structural dynamics. The integrated aero elastic model for predicting performance and structural deections is a prerequisite for the design, development, and optimization of wind turbines. Aerodynamic stall (sudden loss of lift) is an important part of wind turbine operation. Wind turbines are subjected to atmospheric turbulence, wind shear, and change in wind directions in both time and in space, and effects from the wake of neighbouring wind turbines. As a consequence, the forces vary in time and space, and a stochastic description of the wind eld and a dynamical description of the blades and solid structures of the wind turbine are intrinsic parts of the aerodynamic analysis. Aerodynamic modelling is used by industry in the design of new turbines, and on state-of-the- art methods for analysing wind turbine rotors and wakes. Practical horizontal axis wind turbine designs use airfoils to transform the kinetic energy in the wind into useful energy. This section deals with the power production with the use of airfoils, to calculate an optimum blade shape for the start of a blade design and to analyse the aerodynamic performance of a rotor with a known blade shape and airfoil characteristics. A number of authors have derived methods for predicting the steady state performance of wind turbine rotors. The classical analysis of the wind turbine was originally developed by Betz and Glauert (Glauert, 1935) in the 1930s [11]. Subsequently, the theory was expanded and adapted for solution by digital computers. [12]. Momentum theory and blade element theory are developed and used to calculate the optimum blade shape for simplified, ideal operating conditions. The results illustrate the derivation of the general blade shape used in wind turbines. The combination of the two approaches, called strip theory or blad e element momentum (BEM) theory, is then used to outline a procedure for the aerodynamic design and performance analysis of a wind turbine rotor. 2.2. Momentum Theory and Blade-Element Momentum Theory 2.2.1. Basics of Momentum Theory

The basic tool for understanding wind turbine aerodynamics is the momentum theory in which the ow is assumed to be in viscid, incompressible, and axisymmetric. The momentum theory basically consists of control volume integrals for conservation of mass, axial and angular momentum balances , and energy conservation: (Conservation of mass) (2.1)

, (Conservation of axial momentum) (2.2) (Conservation of angular momentum) (Energy equation) (2.3) (2.4)

Where, V = ( ux, ur, u ) is the velocity vector in the axial, radial, and azimuthal direction = the density of air A = Outward-pointing area vector of the control volume p = Pressure T = the axial force (thrust) acting on the rotor Q = Torque P =Power extracted from the rotor. Dimensionless parameters to characterize the aerodynamic operation of a wind turbines are, Tip- speed ratio: Thrust coefficient:
CT

=
=

(2.5) (2.6) (2.7)

Power coefficient: C P Where,

=Angular velocity of rotor A=Rotor area R=Radius of the rotor Uo =Wind speed Based on the simple one-dimensional (1D) momentum theory developed by Rankine (1865), W. Froude (1878) , and R.E. Froude (1889), Betz (1920) showed that the power that can be extracted from a wind turbine is given by Cp = 4a (1 a) 2 (2.8)

Where a = 1 u/Uo is referred to as the axial interference factor (fractional decrease in wind velocity between the free stream and the rotor plane), and u denotes the axial velocity in the rotor plane. Differentiating the power coefcient with respect to the axi al interference factor, the maximum obtainable power is given as CpMax=16/ 27 ~0. 593 for a = 1/ 3. This result is usually referred to as the Betz limit or the Lanchester-Betz-Joukowsky limit and states the upper maximum for power extraction, which is that no more than 59.3% of the kinetic energy contained in a stream tube having the same cross section as the disc area can be converted to useful work by the disc. However, it does not include the losses due to rotation of the wake, and therefore it represents a conservative upper maximum [14].

Figure2.1: Operating parameters for a Betz turbine, U velocity of undisturbed air; U 4 air velocity behind the rotor; C P power coefficient; C T thrust coefficient

The thrust coefficient for an ideal wind turbine is equal to 4a (1 a).CT has a maximum of 1.0 when a=0.5 and the downstream velocity is zero. At maximum power output (a=1/3), C T has a value of 8/9. A graph of the power and thrust coefficients for an ideal Betz turbine and the non-dimensionalised downstream wind speed are illustrated in Figure2.1. This idealized model is not valid for axial induction factors greater than 0.5. In practice, as the axial induction factor approaches and exceeds 0.5, complicated flow patterns that are not represented in this simple model result in thrust coefficients that can go as high as 2.0 The Betz limit , CpMax=16/27, is the maximum theoretically possible rotor power coefficient[15]. In practice, three effects lead to a decrease in the maximum achievable power coefficient. 2.3. Wind Rotation of the wake behind the rotor; Finite number of blades and associated tip losses; Non-zero aerodynamic drag. turbine wake rotation

In the case of a rotating wind turbine rotor, the flow behind the rotor rotates in the opposite direction to the rotor, in reaction to the torque exerted by the flow on the rotor. The generation of rotational kinetic energy in the wake results in less energy extraction by the rotor than would be expected without wake rotation. The extra kinetic energy in the wind turbine wake will be higher if the generated torque is higher. Slow-running wind turbines (with a low rotational speed and a high torque) experience more wake rotation losses than high-speed wind machines with low torque. Power in this case is a function of the axial and angular induction factors (a and a) and the tip speed ratio. The axial and angular induction factors determine the magnitude and direction of the air flow at the rotor plane. The local speed ratio ( r) is a function of the tip speed ratio ( ) and radius. Tip speed ratio () = Local speed ratio ( r) = Angular inductance factor a given by [16], a = * + (2.10) = (2.9)

Where a is axial induction factor, for maximum power production, the value of r given by r =

(2.11)

This equation defines the axial induction factor for maximum power as a function of the local tip speed ratio. From this a = Expression for power coefficient, * + (2.12)

The lower limit of integration, a1, corresponds to the axial induction factor for r =0 and the upper limit, a2, corresponds to the axial induction factor at r =

2 =

(2.13)

From equation 2.11 a1=0.25 give r =0, it is clear from equation 13 that a2 =1/3is the upper limit of the axial induction factor a, giving an infinitely large tip speed ratio.

Figure2.2 shows the Theoretical maximum power coefficient as a function of tip speed ratio for an ideal Horizontal axis wind turbine, with and without wake rotation. The results show that, the higher the tip speed ratio, the closer the CP can approach the theoretical maximum

Figure2.2: Theoretical maximum power coefficient as a function of tip speed ratio for an ideal Horizontal axis wind turbine, with and without wake rotation. The axial and angular induction factors for a turbine with a tip speed ratio of 7.5 are shown in Figure2.3. It can be seen that the axial induction factors are close to the ideal value of 1/3 till near hub. Angular induction factors are close to zero in the outer parts of the rotor, but increase significantly near the hub.

Figure2.3: Induction factors for an ideal wind turbine with wake rotation 2.4. Wind turbine blade aerodynamics Wind turbine blades are shaped to generate the maximum power from the wind at the minimum cost. Design is driven by the aerodynamic requirements, but economics mean that the blade shape is a compromise to keep the cost of construction within a reasonable limit. Wind varies every second due to turbulence developed due to environmental changes. It also blows more strongly higher above the ground than closer to it, due to surface friction. There is a variation of loads on the blades of a wind turbine as they rotate; hence there is a need for the aerodynamic and structural design conditions to merge with that are rarely optimal. But there is an optimum amount of power to be extracted from a given disc diameter, a theoretical maximum of 59% of the winds power to be captured (this is called Betzs limit). Here we are introducing how airfoils can be used to approach this theoretically achievable power extraction. 2.4.1. Airfoil Theory Airfoils are structures with specific geometric shapes that are used to generate mechanical forces due to the relative motion of the airfoil and a surrounding fluid .Wind turbine blades use airfoils to develop mechanical power. Airfoil nomenclature is shown in Figure2.4. The mean camber line is the locus of points halfway between the upper and lower surfaces of the airfoil. The most forward and rearward points of the mean camber line are on the leading and trailing edges,

respectively. The straight line connecting the leading and trailing edges is the chord line of the airfoil, and the distance from the leading to the trailing edge measured along the chord line is designated the chord ,c, of the airfoil. The camber is the distance between the mean camber line and the chord line, measured perpendicular to the chord line. The thickness is the distance between the upper and lower surfaces, also measured perpendicular to the chord line. The angle of attack, , is defined as the angle between the relative wind (Urel) and the chord line. The geometric parameters that have an effect on the aerodynamic performance of an airfoil include: the leading edge radius, mean camber line, maximum thickness and thickness distribution of the profile, and the trailing edge angle.

Figure2.4: Airfoil nomenclature The mean camber line is the locus of points halfway between the upper and lower surfaces of the airfoil. The most forward and rearward points of the mean camber line are on the leading and trailing edges, respectively. The straight line connecting the leading and trailing edges is the chord line of the airfoil, and the distance from the leading to the trailing edge measured along the chord line is designated the chord ,c, of the airfoil. The camber is the distance between the mean camber line and the chord line, measured perpendicular to the chord line. The thickness is the distance between the upper and lower surfaces, also measured perpendicular to the chord line. The angle of attack, , is defined as the angle between the relative wind (Urel) and the chord line. The geometric parameters that have an effect on the aerodynamic performance of an airfoil include: the leading edge radius, mean camber line, maximum thickness and thickness distribution of the profile, and the trailing edge angle. Wind turbine blades work by generating lift due to their shape. The more curved side generates low air pressures while high pressure air pushes on the other side of the aerofoil. The net result is a lift force perpendicular to the direction of flow of the air. The lift force increases as the blade is turned to present itself at a greater angle to the wind. This is called the angle of attack. At very large angles of attack the blade stalls and the lift decreases again. So there is an optimum angle of attack to generate the maximum lift.

Figure2.5 Lift & drag vectors [17] Drag force is parallel to the wind flow, and also increases with angle of attack. If the aerofoil shape is good, the lift force is much bigger than the drag, but at very high angles of attack, especially when the blade

stalls, the drag increases dramatically. So at an angle slightly less than the maximum lift angle, the blade reaches its maximum lift/drag ratio. The best operating point will be between these two angles.

Figure2.6 Blade at low, medium & high angles of attack[17] The wind blowing from a different angle called apparent wind. The apparent wind is stronger than the true wind but its angle is less favourable: it rotates the angles of the lift and drag to reduce the effect of lift force pulling the blade round and increase the effect of drag slowing it down. It also means that the lift force contributes to the thrust on the rotor. The result of this is that, to maintain a good angle of attack, the blade must be turned further from the true wind angle.

Figure2.7 Apparent wind angles[7] The speed at which the turbine rotates is a fundamental choice in the design, and is defined in terms of the speed of the blade tips relative to the free wind speed (i.e. before the wind is slowed down by the turbine). This is called the tip speed ratio. High tip speed ratio means the aerodynamic force on the blades (due to lift and drag) is almost parallel to the rotor axis, so relies on a good lift/drag ratio. The lift/drag ratio can be affected severely by dirt or roughness on the blades. Low tip speed ratio would seem like a better choice but i t results in lower aerodynamic efficiency, due to two effects. Because the lift force on the blades generates torque, it has an equal but opposite effect on the wind, tending to push it around tangentially in the other direction. The result is that the air downwind of the turbine has swirl, i.e. it spins in the opposite direction to the blades. That swirl represents lost power so reduces the available power that can be extracted from the wind. Lower rotational speed requires higher torque for the same power output, so lower tip speed results in higher wake swirl losses. The other reduction in efficiency at low tip speed ratio comes from tip losses, where high-pressure air from the upwind side of the blade escapes around the blade tip to the low-pressure side, thereby wasting energy. The higher lift force on a wider blade also translates to higher loads on the other components such as the hub and bearings, so low tip speed ratio will increase the cost of these items. All this means that turbine designers typically compromise on tip speed ratios in the region of 7-10, so at design wind speed (usually 2- 5 metres

per second) the blade tip can be moving at around 20 m/s (approximately 20 miles per hour). There are practical limits on the absolute tip speed too: at these speeds, bird impacts and rain erosion start to become a problem for the longevity of the blades and noise increases dramatically with tip speed [17].

Figure2.10 Swirl in the wake [17]. 2.4.2. Non dimensional parameters Air flow over an airfoil produces a distribution of forces over the airfoil surface. The flow velocity over airfoils increases over the convex surface resulting in lower average pressure on the suction side of the airfoil compared with the concave or pressure side of the airfoil. Viscous friction between the air and the airfoil surface slows the air flow to some extent next to the surface. The resultant of all of these pressure and friction forces is usually resolved into two forces and a moment that act along the chord at a distance of c/4 from the Leading edge. Figure2.5 shows forces and moments on an airfoil section. The represents angle of attack ;c is the chord. The direction of positive forces and moments is indicated by the direction of the arrow.

Figure2.5 Forces and moments on an airfoil section Lift forcedefined to be perpendicular to direction of the oncoming air flow. The lift force is a consequence of the unequal pressure on the upper and lower airfoil surfaces. Drag forcedefined to be parallel to the direction of the oncoming air flow. The drag force is due both to viscous friction forces at the surface of the airfoil and to unequal pressure on the airfoil surfaces facing toward and away from the oncoming flow. Pitching moment defined to be about an axis perpendicular to the airfoil cross-section The most important non-dimensional parameter for defining the characteristics of fluid flow conditions is the Reynolds number. The Reynolds number, Re, is defined by: (2.14) Where is the fluid density is fluid viscosity, that characterize the scale of the flow. the kinematic viscosity, and U and L are velocity and length

Three-dimensional airfoils have a finite span and force and moment coefficients are affected by the flow around the end of the airfoil. Two-dimensional airfoil data, on the other hand, are assumed to have an infinite span (no end effects). Two-dimensional data are measured in such a way that there is indeed no air flow around the end of the airfoil in the test section. Force and moment coefficients for flow around two-dimensional objects are usually designated with a lower case subscript, as in Cd for the two-dimensional drag coefficient. In that case, the forces measured are forces per unit span. Lift and drag coefficients that are measured for flow around three- dimensional objects is usually designated with an upper case subscript, as in CD. Rotor design usually uses two-dimensional coefficients, determined for a range of angles of attack and Reynolds numbers, in wind tunnel tests. The two-dimensional lift coefficient is defined as: (2.15) The two-dimensional drag coefficient is defined as: (2.16)

And the pitching moment coefficient is: (2.17)

Where is the density of air, U is the velocity of undisturbed air flow, A is the projected airfoil area (Chord x span), c is the airfoil chord length, and l is the airfoil span. Other dimensionless coefficients that are important for the analysis and design of wind turbines include the power and thrust coefficients and the tip speed ratio, The pressure coefficient: used to analyse airfoil flow, (2.18) The surface roughness ratio: (2.19) 2.4.3. Flow over an airfoil The lift, drag, and pitching moment coefficients of an airfoil are generated by the pressure variation over the airfoil surface and the friction between the air and the airfoil. As the air flow accelerates around the rounded leading edge, the pressure drops, resulting in a negative pressure gradient. As the air flow approaches the trailing edge, it decelerates and the surface pressure increases, resulting in a positive pressure gradient. If, given the airfoil design and the angleof attack, the air speeds up moreover the upper surface than over the lower surface of the airfoil, then there is a net lift force. Similarly, the pitching moment is a function of the integral of the moments of the pressure forces about the quarter chord over the surface of the airfoil. Drag forces are a result of both the pressure distribution over the airfoil and the friction between the air flow and the airfoil. The component of the net pressure distribution in the direction of the air flow results in the drag due to the pressure. Drag due to friction is a function of the viscosity of the fluid and dissipates energy into the flow field. Drag also causes the development of two different regions of flow: one farther from the airfoil surface, where frictional effects are negligible and the boundary layer, immediately next to the airfoil surface, where frictional effects dominate. In the boundary layer, the velocity increases from zero at the airfoil surface to that of the friction-free flow outside of the boundary layer. The boundary layer on a wind turbine blade may vary in thickness from a millimetre to tens of centimetres.

The pressure gradient of the flow has a significant effect on the boundary layer that pressure gradient may be a favourable pressure gradient (positive in the direction of the flow) or an adverse pressure gradient (against the flow). Flow in the boundary layer is accelerated or decelerated by the pressure gradient. In the boundary layer, the flow is also slowed by surface friction. Thus, in an adverse pressure gradient and with the help of surface friction, the flow in the boundary layer may be stopped or it may reverse direction. These results in the flow separating from the airfoil, causing a condition called stall. Boundary layers that have already transitioned to turbulent flow are less sensitive to an adverse pressure gradient than are laminar boundary layers, but once the laminar or turbulent boundary layer has separated from the airfoil, the lift drops. An airfoil can only efficiently produce lift as long as the surface pressure distributions can be supported by the boundary layer.[8]

Figure 2.6 Effects of favourable (decreasing) and adverse (increasing) pressure gradients on the boundary layer[17] Effects of favourable (decreasing) and adverse (increasing) pressure gradients on the boundary layer Is shown in Figure2.6 .Wind turbine airfoils operate in the turbulent planetary boundary layer but the scale of the turbulent fluctuations in the atmosphere is much larger than the scale of the turbulence in the boundary layer of a wind turbine airfoil. The flow in the boundary layer is only sensitive to fluctuations on the order of the size of the boundary layer itself. Thus, the atmospheric turbulence does not affect the airfoil boundary layer directly. It may affect it indirectly through changing angles of attack, which will change the flow patterns and pressure distributions over the blade surface.

You might also like