+ All Categories
Home > Documents > Energy and Exergy Efficiency Comparison

Energy and Exergy Efficiency Comparison

Date post: 03-Mar-2016
Category:
Upload: efrain-valle
View: 16 times
Download: 0 times
Share this document with a friend
Description:
Energy and exergy efficiency comparison of horizontal and vertical axiswind turbines

of 12

Transcript
  • Energy and exergy efciency comparison of horizontal and vertical axis

    wind turbines

    K. Pope, I. Dincer, G.F. Naterer*

    Faculty of Engineering and Applied Science, University of Ontario Institute of Technology, 2000 Simcoe Street North, Oshawa, Ontario, Canada L1H 7K4

    a r t i c l e i n f o

    Article history:

    Received 11 February 2009

    Accepted 14 February 2010

    Available online 25 March 2010

    Keywords:

    Wind turbine

    Exergy

    Energy

    Efciency

    a b s t r a c t

    In this paper, an energy and exergy analysis is performed on four different wind power systems,

    including both horizontal and vertical axis wind turbines. Signicant variability in turbine designs and

    operating parameters are encompassed through the selection of systems. In particular, two airfoils (NACA

    63(2)-215 and FX 63-137) commonly used in horizontal axis wind turbines are compared with two

    vertical axis wind turbines (VAWTs). A Savonius design and Zephyr VAWT benet from operational

    attributes in wind conditions that are unsuitable for airfoil type designs. This paper analyzes each system

    with respect to both the rst and second laws of thermodynamics. The aerodynamic performance of each

    system is numerically analyzed by computational uid dynamics software, FLUENT. A difference in rst

    and second law efciencies of between 50 and 53% is predicted for the airfoil systems, whereas 44e55%

    differences are predicted for the VAWT systems. Key design variables are analyzed and the predicted

    results are discussed. The exergetic efciency of each wind turbine is studied for different geometries,

    design parameters and operating conditions. It is shown that the second law provides unique insight

    beyond a rst law analysis, thereby providing a useful design tool for wind power development.

    ! 2010 Elsevier Ltd. All rights reserved.

    1. Introduction

    Wind power systems have achieved signicant improvement in

    operating efciencies, making them more economically competi-

    tive with other energy generation techniques. Along with the need

    for increased sustainability in the energy sector, wind energy

    systems are increasing their market share faster than any other

    renewable energy system [1]. Horizontal axis wind turbines

    (HAWTs) have emerged as the dominant technology in modern

    wind energy technologies. In comparison to a vertical axis wind

    turbine (VAWT), a HAWT can achieve higher energy efciencies,

    thereby increasing the power production and reducing system

    expense per kW of power generated. But as the opportunity to

    expand wind capacity increases, it is important that all aspects of

    this sustainable and environmentally benign technology are fully

    developed. VAWTs have demonstrated an ability to fulll certain

    energy generation requirements that cannot be fullled by HAWTs.

    A HAWT can achieve higher efciencies, but only if the energy

    quality of the wind is high. High wind turbulence, wind uctua-

    tions, and high directional variability can cause signicant prob-

    lems for a HAWT, whereas VAWTs can operate well.

    Local or distributed power generation has attracted signicant

    attention in recent years. The resurgence of this old technology is

    partly attributed to the need for environmentally benign and

    sustainable energy systems in smaller communities. This includes

    diversifying the generation techniques and increasing the system

    efciencies, while minimizing the environmental impact. Onsite

    power generation can overcome transmission losses and land costs.

    However, densely populated locations and urban centers generally

    coincide with a low quality of wind source, including high turbu-

    lence, uctuations in intensity, and highly variable direction of the

    ow streams [2]. Variable pitch blades can improve turbine

    performance by varying the angle of attack to coincide with the

    various wind conditions, but this approach is generally not

    economically practical for small installations [3]. Fluctuating winds

    can greatly reduce a HAWT's performance as long as idling periods

    are experienced at start-up when the rotor accelerates slowly. For

    a small HAWT, a past study reported it to be 50 s at a wind speed of

    8 m/s [4]. Certain VAWT designs have the ability to operate in these

    harsh operating conditions. However, there is no clear method to

    compare these different turbine designs in various wind conditions

    that are inherent to the different operating regions.

    Typical design methodologies employ the rst law of thermo-

    dynamics for wind power system analysis and design. Empirical

    tests and experience must often be used to improve system

    performance and implementation. The process irreversibilities are

    * Corresponding author.

    E-mail addresses: [email protected] (K. Pope), ibrahim.dincer@

    uoit.ca (I. Dincer), [email protected] (G.F. Naterer).

    Contents lists available at ScienceDirect

    Renewable Energy

    journal homepage: www.elsevier .com/locate/renene

    0960-1481/$ e see front matter ! 2010 Elsevier Ltd. All rights reserved.

    doi:10.1016/j.renene.2010.02.013

    Renewable Energy 35 (2010) 2102e2113

  • not represented in the analysis [5]. A theoretical maximum ef-

    ciency can be predicted, but irreversibilities are not identied.With

    a rst law methodology, the designer includes a predetermined

    design factor to account for the irreversibilities. Past experimental

    data reported that actual ow across a wind turbine rotor is about

    20% slower than the ideal ow [6]. Predicting turbine performance

    with complicated variations in operating demands and design

    congurations reveal the deciencies with this strategy [7]. In

    contrast, the second law denes a quality of energy and quantity of

    irreversibility or loss associated with the thermodynamic process.

    In this paper, the concept of entropy generation will be used to

    describe the magnitude of energy dissipation. Higher levels of

    entropy generation are associated with a lower level of useful

    energy. The second law requires that the amount of entropy in an

    isolated system will always increase [8]. This principle can be

    applied to a variety of engineering applications.

    Entropy-based design and exergy analysis have been shown to

    identify the maximum theoretical capability of energy system

    performance in various applications. For example, it can provide

    component-level energy management to improve diffuser perfor-

    mance [9] and reduce voltage losses within a fuel cell [10]. Exergy

    analysis has been used to diagnose inefciencies of power plants

    [11], minimize the carryover leakage irreversibilities in a power

    plant regenerative air heater [12], and many other power plant

    associated applications. These studies have shown exergy analysis

    to be very useful for improving a wide range of thermo uid

    systems. Exergy analysis also provides a design tool for increased

    accuracy and more efcient performance.

    However, there are few examples in past literature that pertain

    to wind exergy. Through an energy and exergy analysis of the

    characteristics of wind energy, it was found that differences

    between energy and exergy efciencies are approximately 20e24%

    at lowwind speeds and approximately 10e15% at high wind speeds

    [13]. Sahin et al. [14] developed a useful exergetic analysis tech-

    nique for determining the exergetic efciency of a wind turbine.

    The technique utilizes the wind chill temperature associated with

    wind velocity to predict the entropy generation of the process.

    Better turbine design and location selection can be achieved with

    the aid of such exergy analysis.

    In this paper, a comparison of second law efciencies for four

    different wind power systems will be presented. The analysis is

    intended to compare turbines that have different performance

    advantages in various operating conditions. A parametric study

    investigates the selection and associated predictions of key variables

    foreach system.Thispaperwill developa second lawanalysisofwind

    power for potentially valuable utility in the wind energy industry.

    2. Wind energy system description

    Four systems will be investigated in this paper: two airfoils

    commonly used for horizontal axis wind turbines [15,16] and two

    VAWTs (Savonius and Zephyr) operating under low quality wind

    properties (see Figs. 1 and 2). These include (i) a NACA 63(2)-215

    airfoil developed by the National Advisory Committee for Aero-

    nautics (NACA) and (ii) the Wortmann FX 63-137. The VAWTs

    include (iii) a conventional Savonius design and (iv) a more

    complex Zephyr VAWT prototype. A numerical model for each

    system is developed for the uid ow analysis. Computational Fluid

    Dynamics software, FLUENT [17], is used to predict the operation of

    each system. These numerical models will offer insight into the

    uid ow characteristics for each of the different turbines. Data

    gained from the numerical predictions will be used to examine the

    second law efciencies of wind power systems.

    2.1. System 1: NACA 63(2)-215 airfoil

    As presented in Fig. 3a, the NACA 63(2)-215 airfoil is a conven-

    tional design that creates lift with low wind speeds, making it

    suitable for use in wind turbine blades [15,16]. Throughout the

    following analysis, the approach angle (4) is estimated to be 10!. The

    incomingcomponentsof velocityareVx9.85m/s andVy1.74m/s.

    The X and Y force vectors for the lift and drag components become

    XL #sin(10!) #0.174, YL cos(10

    !) 0.985, XD cos

    (10!) 0.985, and YD sin(10!) 0.174, respectively. Table 1

    summarizes the systemassumptions used in the numerical analysis.

    The prole of the NACA 63(2)-215 airfoil is discretized with

    a structured quadrilateral cell scheme. The mesh is rened from an

    average cell edge length of 260mm at the outer region, to 14 mm at

    the airfoil surface. Rened to a total of 12,150 cells, with an average

    cell size of 0.061 m2, produces results that are independent of

    further grid renement [18]. The governing equations are the

    incompressible form of NaviereStokes equations. The standard k-3

    Nomenclature

    A Area [m2]

    B Number of blades [e]

    Cp Specic heat [kJ/kg K]

    Cpower Power coefcient [e]

    E Energy [kJ]

    ex Specic exergy [kJ/kg]_Ex Exergy rate [kW]

    I Irreversibilities [kW]

    KE Kinetic energy [kJ]

    m Mass [kg]

    P Power [kW]

    R Radius [m]

    t Time [s]

    T Temperature [!C]

    U Volume [m3]

    V Velocity [m/s]

    W Work [kJ]_W Work rate [W]_m Mass ow rate [kg/s]

    Greek

    h Energy efciency [e]

    l Tip speed ratio [e]

    r Air density [kg/m3]

    4 Approach angle [!]

    J Exergy efciency [e]

    u Humidity ratio [e]

    Subscripts

    0 Ambient

    2 Denition 2 (exergy)

    B Benz limit

    D Drag

    dest Destruction

    eff Effective

    KE Kinetic energy

    L Lift

    ph Physical

    x Horizontal vector

    y Vertical vector

    K. Pope et al. / Renewable Energy 35 (2010) 2102e2113 2103

  • model is used to simulate turbulence in the ow eld. This is

    a widely used model that provides reasonable accuracy and

    a robust ability to represent a wide range of ow regimes [17]. The

    features of the airfoil solver are summarized below:

    $ Finite volume method with a segregated solver;

    $ Standard ke3 turbulence model;

    $ Standard wall functions for near-wall treatment;

    $ Air density is constant, 1.225 kg/m3;

    $ Air viscosity is constant, 1.7894 % 10#5 kg/m s;

    $ Backow turbulence intensity, 12%;

    $ Velocity-inlet e velocity specication method;

    $ Inlet x-velocity, 9.85 m/s;

    $ Inlet y-velocity, 1.74 m/s;

    $ Airfoil roughness height, 500 mm;

    $ Pressureevelocity coupling e SIMPLE;

    $ Turbulence kinetic energy e second order upwind; and

    $ Turbulence dissipation rate e second order upwind.

    2.2. System 2: FX 63-137 airfoil

    This airfoil has been examined widely in past literature. Its

    characteristics will be used to investigate the aerodynamic behav-

    iour of wind turbines [15,16]. In comparison to other commonwind

    turbine airfoil proles, this design provides a relatively complex

    conguration. The pronounced tail curvature provides a substantial

    contrast to System 1 that will be evaluated (see Fig. 3b).

    The same solver features, including the governing and turbu-

    lence equations, as with System 1 e NACA 63(2)-215 ewill be used

    for this system. Variations in airfoil surface roughness are pre-

    sented in Fig. 4. Convergence in both lift and drag coefcients are

    Fig. 1. Zephyr vertical axis wind turbine (a) illustration and (b) geometrical variables.

    Fig. 2. Sample mesh discretization of VAWT e a) rotor, b) stator and surrounding sub-

    domain. Fig. 3. Velocity contours (m/s) for (a) NACA 63(2)-215 and (b) FX 63-137 airfoils.

    K. Pope et al. / Renewable Energy 35 (2010) 2102e21132104

  • evident after 250 mm. For both airfoils, a CL/CD ratio will be used for

    the power coefcient predictions from a roughness height of

    500 mm. The mesh is rened from an average of 289 mm at the

    outer region, to 0.014 mm at the airfoil surface. It is rened to

    12,195 cells, with an average cell size of 0.061 m2.

    2.3. System 3: savonius VAWT

    The Savonius VAWT is a common design that is capable of

    reaching substantial efciencies. There is signicant past literature

    on this design. For example, extensive experiments were per-

    formed by Saha [19], whereby 16 Savonius models with identical

    aspect ratios were compared in wind tunnel tests. The study

    included an investigation of the optimum number of rotor blades

    (two or three), number of stages (one, two, or three), and the best

    blade shape (twisted or semicircular). The author found that a two-

    stage design, with two twisted rotor blades, preformed the best,

    achieving a maximum power coefcient of 0.32. Biswas et al. [20]

    investigated experimentally the overlap effect of rotors on a Savo-

    nius wind turbine. The authors achieved a power coefcient of 0.37.

    A hybrid Savonius e Darrieus turbine achieved a maximum Cpowerof 0.51, in a comparative study described by Gupta et al. [21]. In the

    current paper, a basic semicircular design will be used, with a 10%

    overlap (see Fig. 5a).

    The domain is discretized into 13,693 cells with an average cell

    size of 0.019 m2. The mesh is rened from 500mm at the boundary,

    to 50 mm at the rotor surface. Again, the same governing and

    turbulence equations, as for previous cases, will be used for this

    system. However, considerable differences in some solver features

    are required. A transient mesh formulation is used to simulate the

    rotor motion. This allows the rotor to rotate 50 time steps per

    revolution, with a time step size of 0.0209 s. The numerically pre-

    dicted average moment induced on the rotor blades will be used to

    determine the power coefcient. The wall roughness height is

    assumed negligible, and the incoming velocity is simulated to be

    Vx 10 m/s and Vy 0.

    2.4. System 4: Zephyr VAWT

    As illustrated in Figs. 1, 2 and 5b, the Zephyr turbine represents

    a more complex VAWT. The features of this turbine allow it to

    perform in both low wind and high turbulence conditions. Thus,

    several turbines can operate in close proximity of each other in

    urban areas. The maximum power coefcient of this turbine is

    relatively low, in comparison to other systems. However, it is an

    ideal candidate to operate in low quality winds and offers a good

    contrast to the Savonius VAWT and airfoil HAWT systems. The

    stator sub-domain is discretized with an edge length of 191 mm at

    the exterior, which is gradually rened to an edge length of

    6.35 mm at the rotor. This yields approximately 117,000 cells, with

    an average area of 217 mm2. At this resolution, the results become

    independent of grid spacing [22]. The analysis of this system

    employs the same solver features, including the governing and

    turbulence equations, as the Savonius VAWT.

    3. Formulation of rst and second law efciencies

    This section will compare the rst law efciency (h) with the

    second law efciency (J). The energy efciency is dened as the

    ratio of useful work to the difference in kinetic energy,

    Table 1

    Signicant system assumptions.

    V1 [m/s] Radius [m] Area [m2] Wout Wout,2

    NACA 63(2)-215 airfoil 10 1 p % 12 f(CL, CD) DKE

    FX 63-137 airfoil 10 1 p % 12 f(CL, CD) DKE

    Savonius VAWT 10 1 2 f(T, l) DKE

    Zephyr VAWT 10 0.762 0.7622 f(T, l) DKE

    Fig. 4. Airfoil performance with variable surface roughness.

    Fig. 5. Velocity contours (m/s) for (a) Savonius and (b) Zephyr VAWTs.

    K. Pope et al. / Renewable Energy 35 (2010) 2102e2113 2105

  • h _WoutP

    (1)

    The exergy efciency refers to the ratio of useful work to the

    exergy of the wind,

    J _Wout

    _Exflow(2)

    In general, the energy balance equation for a wind turbine can

    be represented by

    KE1 Wout KE2 (3)

    where KE is the kinetic energy of the ow stream and Wout is the

    useful work produced by the turbine. The equation describes the

    energy content of air as it passes through a two-dimensional plane.

    It is based on the kinetic energy of wind, which is

    KE 1

    2mV2 (4)

    where KE is energy, m is mass, and V is the wind velocity.

    The mass of wind is difcult to measure, so a more convenient

    variable is volume (U), as related to mass bym r U, where r is the

    density of air. The volume is expressed by the product of the cross-

    sectional area perpendicular to the wind (A) and the horizontal

    length of incoming wind (L). The horizontal length of incoming

    wind is then expressed as LU t. This re-arrangement results in the

    more convenient following expression,

    KE 1

    2rAtV3 (5)

    Since power is related to energy by P E/t, the above expression is

    more commonly used in the following form,

    P 1

    2rAV3 (6)

    where P is the magnitude of power, r is the density of air, A is the

    cross-sectional area perpendicular to wind, and V is the wind

    velocity. This expression describes total kinetic power of a wind

    stream. It is used for the production of wind power maps that are

    used for turbine placement and resource estimation. Although KE1can be readily determined from velocity measurements or predic-

    tions, many problems are associatedwith determining KE2. The exit

    velocity from a wind turbine is extremely difcult to measure, as it

    is highly variable and it quickly dissipates in all directions. As

    a result, empirical work outputs are generally required to deter-

    mine the wind turbine efciency (h).

    Methods for describing the independent variables in the energy

    analysis are relatively straightforward. The density of air is most

    affected by temperature and relative humidity of the air. Temper-

    ature can be readily measured and computed. As a result, density is

    normally selected solely on this measure (i.e., r (T), where T air

    temperature). The air moisture content (u) also affects wind

    density. A more accurate measure of density is r (T, u). However,

    this is not commonly modelled in practice.

    A second law analysis includes the ow irreversibilities associ-

    ated with the system. The exergy balance equation can be

    expressed as

    _Ex1 _Wout _Ex2

    _Exdest (7)

    where _Exdest represents the exergy destruction associated with the

    process. It is a representative measure of the irreversibilities

    involved with the process. This methodology offers a useful alter-

    native measure of turbine efciency that includes the

    irreversibilities, which were not included in the rst law analysis.

    The exergy of ow, _Exflow, can be dened as the maximum attain-

    able work acquired as the air ows through the turbine [7]. The

    relevant terms include physical _Exph and kinetic exergy _ExKE,

    _Exflow _Exph

    _ExKE (8)

    Physical exergy includes the enthalpy and entropy changes

    associated with the turbine operation, expressed as [14]

    _Exph _m

    !CPT2 # T1 T0

    "CPln

    "T2T1

    ## Rln

    "P2P1

    #

    #CP$T0 # Taverage

    %T0

    #&(9)

    where Pi P0 ) r=2V2 and T1 and T2 are determined through the

    wind chill temperature, based on a model developed by Zecher

    [23]. The rst term in the square brackets represents the change in

    ow enthalpy, while the second term characterizes the ow irre-

    versibilities of the system. The irreversibilities associated with the

    system can be determined by

    I #T0DS (10)

    or

    I_ T0

    "CPln

    "T2T1

    ## Rln

    "P2P1

    ##

    _mCP$T0 # Taverage

    %T0

    #(11)

    The kinetic component of the ow exergy is equivalent to the

    difference of kinetic energy through the turbine (i.e., DKE. From

    Eq. (3), the change in kinetic energy can also be expressed by the

    work output of the turbine. This methodology offers a signicant

    opportunity to improve the wind turbine design and enhance the

    site selection by supplementing the information provided by the

    Table 2

    Signicant numerical assumptions and predictions.

    l P0 [kPa] T0 [C] CL CD CM Cpower

    NACA 63(2)-215 airfoil 4 101.3 25 1.15 0.0385 e 0.44

    FX 63-137 airfoil 4 101.3 25 1.87 0.0357 e 0.47

    Savonius VAWT 0.5 101.3 25 e e 0.0294 0.18

    Zephyr VAWT 0.5 101.3 25 e e 0.0147 0.11

    0

    2

    4

    6

    8

    10

    1 2 3 4

    V2[m/s]

    Definition of V2

    NACA 63(2)-215 Airfoil FX 63-137 Airfoil Savonius VAWTZephyr VAWT

    Fig. 6. Denitions of V2 for a variety of wind power systems.

    K. Pope et al. / Renewable Energy 35 (2010) 2102e21132106

  • rst law analysis. The specic exergy destruction can be dened

    as

    exdest T0DS

    rAV(12)

    For the airfoils, _Wout will be determined by the model of Wilson

    et al. [24]. This is an empirical technique that correlates the turbine

    power coefcient with the lift (CL), drag (CD), number of blades (B),

    and tip speed ratio (l).

    Cpower

    "16

    27

    #l

    24l 1:32

    l#820

    )2B

    23

    35#1

    #0:57l

    2

    CLCD

    $l 12B

    % (13)

    The lift and drag coefcients are estimated from numerical

    predications obtained at the specied tip speed ratio and blade

    number. For the VAWTs, _Wout is determined from the product of

    a numerically predicted torque and the simulated rotational

    velocity.

    Fig. 7. Energy and exergy efciencies based on (a) kinetic energy, (b) ow exergies, (c) V2 maintained constant, (d) V2/V1 maintained constant and (e) Benz efciencies.

    K. Pope et al. / Renewable Energy 35 (2010) 2102e2113 2107

  • 4. Results and discussion

    In this section, the results of the rst and second law analyses

    will be presented. The numerical simulations are used to obtain

    performance predictions of the various wind systems and provide

    useful information about the uid ow behaviour. For the airfoils,

    the numerically predicted coefcients of lift and drag are used to

    predict the maximum power coefcient through Eq. (12). A

    numerically predicted moment coefcient is used to determine the

    power coefcient of the VAWTs. The results of the numerical study

    are summarized in Table 2.

    Using the second law analyses for wind turbines requires V2. A

    standard metric for nding V2 is crucial for maintaining consistent

    exergetic efciency predictions for different wind turbine designs

    and operating conditions. In this paper, four different denitions for

    determining V2 are presented. Each denition is investigated with

    respect to its reference variables and operating conditions. The

    second law predictions achieved from each denition are the

    compared with the rst law predictions and discussed with regards

    to accuracy. The ability to represent the performance of different

    turbine designs and operating conditions with a standard metric is

    a signicant contribution of the second law. Each denition of V2and its ability to uphold these criteria will be identied. Four

    denitions below of V2 will be used in the analysis (see Fig. 6),

    $ Denition 1: Point specic, low wind velocity (measured one

    chord behind the turbine);

    $ Denition 2: Specic effective velocity (V2 V2,eff % Aeff/A);

    $ Denition 3: Average velocity of useful area (V2 V2,useful,ave);

    $ Denition 4: Average of the low velocity stream.

    A parametric study of the reference conditions and operating

    wind conditions will be presented for each of the four wind energy

    systems. The results will be compared for each denition of V2. This

    includes (i) point specic low velocity, (ii) specic effective velocity,

    (iii) average velocity of useful area, and the (iv) average of the low

    velocity stream.

    4.1. Denition 1: point specic, low wind velocity (measured one

    chord behind the turbine)

    Denition 1 is advantageous because it can be determined in

    a consistent manner for different designs. Also, this denition lends

    itself well to physical measurements. Virtually identical values of V2are predicted for the VAWTs, with similar values also suggested for

    the airfoils. The resultant airfoil second law efciencies are 15% and

    17% for the NACA 63(2)-215 and FX 63-137, respectively. A 51e50%

    distinction between the rst and second law efciencies is

    predicted for the airfoil designs. A different trend is forecasted for

    the VAWTs. The second law efciency for the Savonius VAWT is

    predicted to be 17%, a 6% difference from the rst law predictions.

    Comparably, a 10% exergy efciency, which is 9% different than the

    rst law predictions, is forecasted for the Zephyr VAWT.

    Fig. 8. Energy and exergy efciencies with varying pressure for (a) point specic low velocity, (b) specic effective velocity, (c) effective velocity and (d) average low velocity.

    K. Pope et al. / Renewable Energy 35 (2010) 2102e21132108

  • Results of the energetic and exergetic analysis using Denition 1

    for determining V2 are presented in Figs. 7e12a. Variations to the

    reference conditions P0 and T0 are presented in Figs. 8a and 9a,

    respectively. Theminor effects of altering P0 are included in the rst

    law efciencies for the VAWTs, but they do not affect the airfoils.

    The discrepancy is caused by the method of determining _Wout, or

    more specically, the power coefcient. The analysis of the airfoils

    uses an empirical correlation, which is independent of the refer-

    ence pressure. In contrast, for the energy analysis, the VAWT

    predictions include the effects of local pressure on the local wind

    density, when predicting the power coefcient. The plot of varying

    reference temperature includes the majority of the standard

    operating conditions, with an equal distribution from 25 !C, stan-

    dard reference temperature. Similar to the variable P0, it can be

    observed that although the second law efciencies depend on the

    choice of reference conditions, it allows different wind power

    systems to be compared with one descriptive parameter.

    Variations to the input velocity, V1, are presented in Figs. 10a and

    11a. These gures illustrate the effects of variations of inlet velocity,

    while comparing the assumptions that (a) V2 is constant, or (b) V2/

    V1 is maintained constant. The rst law analysis fails to predict

    changes to the system efciency when this crucial operating

    condition is altered. In comparison, despite the assumption of V2,

    the second law predicts an increased efciency with higher values

    of V1. The linear trend displayed by V2 has a higher variability,

    suggesting that this denition of V2 increases reliability with the

    assumption of V1/V2 C. More importantly, the second law

    efciency reveals variability in the efciency of each system for

    different wind conditions. This is valuable information when

    designing a turbine for a wide variety of wind conditions, or

    selecting a turbine for a specic site with one of the many different

    possible operating requirements.

    Sahin et al. [14] proposed that the shaft work is used to estimate

    the kinetic component of ow exergy. The same value is used to

    represent the work output, thereby assuming that the turbine is

    100% energy efcient. This paper offers two alternative methods,

    using one of the proposed denitions for obtaining V2 to represent

    the change in kinetic energy. This includes the effects of DKE on (a)

    its inclusion in _Exflow as_ExKE, and (b) assuming DKE Wout. To

    differentiate, the second law efciencies in (a) are denoted as J2.

    Fig. 7a illustrates the variations in the incoming velocity, where hKEand JKE represent the rst and second law efciencies, with

    DKE Wout. A high level of variability is expected with the de-

    nition of V2, thereby providing signicant information about the

    effects of the denition. Furthermore, these plots present the point

    specic change in kinetic energy, as stated by the rst law. Fig. 7c

    and d present the methods of predicted ow exergy with

    a varying inlet velocity. A comparison with Figs. 10a and 11a

    predicts a 50e53% difference in the rst and second law efciencies

    for the airfoil systems, and 44e55% for the VAWTs, at reference

    conditions.

    Fig. 7e illustrates the results of a second law analysis of the Benz

    limit. The theoretical maximum energy efciency is obtained with

    the Benz limit. With the rst law, the Benz limit is a constant value,

    Fig. 9. Energy and exergy efciencies with varying temperature for (a) point specic low velocity, (b) specic effective velocity, (c) effective velocity and (d) average low velocity.

    K. Pope et al. / Renewable Energy 35 (2010) 2102e2113 2109

  • independent of operating conditions. However, combining the Benz

    limit theory with second law analysis provides a theoretical

    maximum efciency that includes the effect of irreversibilities,

    resulting in a dependence on design and operating conditions.

    Dening it here as 0:59 _Woutt= _Exflow, the second law Benz limit

    with both methods of obtaining _Exflow (i.e., JB, and J2,B) is pre-

    sented. Signicant variability between the VAWTs and the airfoils is

    predicted by JB from 29% to 59%, while the range of J2,B is only

    28e32%. Table 3 summarizes the rst and second law efciencies,

    predicted for the reference operating conditions.

    4.2. Denition 2: specic effective velocity (V2 V2,eff % Aeff/A)

    This denition suggests a high level of accuracy with the second

    law, as it attempts to specify the acting ow stream on the turbine.

    This denition can provide a high level of comparability between

    different turbine designs and congurations. However, the precision

    of analysis could be a problem with this denition, as it requires

    a level of intuition from the analyst. Figs. 8e12bpresent the results of

    varied reference conditionsandoperatingconditions forDenition2.

    A noticeable reduction in variability is experienced between the

    airfoils from Denition 2. Also, the exergetic variability between the

    airfoils and VAWTs is reduced. Difculties in dening the effective

    area for an airfoil could be a contributing factor to this result.

    In this paper, the effective area was assumed as the high speed

    ow streamdirected above the airfoil, with the total area taken to be

    the chord length. This denition does not fully represent the

    differences in geometry between the airfoils, as the effects of tail

    curvature are not fully represented. Dening the effective area for

    the VAWT is comparatively straightforward. A low velocity ow

    stream is evident in the locations where a signicant kinetic force is

    applied. An effective area is assumed as the cross-sectional area of

    this ow stream, with a total area assumed to be the turbine diam-

    eter. Anotable result fromDenition2 is illustrated in Fig.11b,where

    the airfoil second law efciencies exhibit a non-linear reduction in

    efciency, falling rapidly after the referencewind velocity of 10m/s.

    The VAWT second law efciencies display a slight linear increase.

    ThehighvaluesofV2 suggested fromDenition2produce the lowest

    predictions ofJ throughout the study.

    4.3. Denition 3: average velocity of useful area (V2 V2,useful,ave)

    Denition 3 predicts the largest range in V2, with 9.5 m/s for the

    NACA 63(2)-215 airfoil, compared to 3.1 m/s for the Zephyr VAWT.

    A high dependence on the streamline conguration of the turbine

    is obtained by this denition. A high level of precision is attainable,

    compared with Denition 2, since the value is independent of size

    for the effective area. The analysis of Denition 2 reveals an output

    for J that is evenly distributed at the reference conditions and

    throughout most of the varied operating conditions. Similar to

    Denition 2, the relatively high values of V2 for the airfoil translate

    into low second law efciencies. However, the VAWTs are less

    affected. From Fig. 11c, the prole of an airfoil can signicantly

    affect the second law efciency. The basic prole of the NACA 63(2)-

    215 exhibits only a slight reduction in second law efciency,

    compared to the FX 63-137. A decreasing trend, which increases its

    Fig. 10. Energy and exergy efciencies with varying V1, constant V2 for (a) point specic low velocity, (b) specic effective velocity, (c) effective velocity and (d) average low velocity.

    K. Pope et al. / Renewable Energy 35 (2010) 2102e21132110

  • rate of reduction throughout the plot, is predicted for the more

    complex FX 63-137. At the upper boundary of V1, the second law

    efciency of the FX 63-137 airfoil falls below the value for the

    Savonius VAWT.

    4.4. Denition 4: average of the low velocity stream

    This denition suggests a relatively even distribution of V2amongst the various systems. This translates into evenly distrib-

    uted second law efciencies, at the given reference operating

    conditions illustrated in Figs. 8e12d. Similar to the other deni-

    tions, little effect is exhibited from altering the reference pressure.

    This denition suggests a high variability between the airfoil

    second law efciencies. Similar to Denition 3, the second law

    efciency for the NACA 63(2)-215 airfoil deteriorates rapidly

    throughout the range of V1, dropping below the Savonius second

    law efciency within the operating velocity. The declining NACA 63

    (2)-215 airfoil second law efciency suggests that this denition

    can give insight into the interdependence of the geometric prole,

    operating conditions, and turbine performance.

    Fig. 12 compares the various denitions of V2 in terms of exdest,

    assuming V2/V1 is maintained constant. Many of the previous

    results can be understood through the exergy destruction. The rst

    denition of V2 predicts similar exergy destruction for the airfoils

    and VAWTs, respectively. It appears that this denition fails to fully

    represent the differences in ow irreversibilities between all

    systems. Better results would likely occur from Denitions 3 and 4,

    whereby a greater range of variability is predicted between the

    various systems.

    These case studies have compared the rst and second law

    efciencies of four wind energy systems. Difculties exhibited

    with applying the exergetic analysis have been identied and

    preliminary solutions were obtained. The denition of V2 has been

    identied as a key component for implementing the second law in

    regular wind power analysis, optimization and design. The second

    law provides a valuable design tool that can help improve the

    efciency and economic cost of wind power. Improvements to

    system output, development and installation can contribute to

    wind power systems alleviating the substantial demand from

    traditional non-renewable energy sources. There is an urgent

    need to alter these current consumption and production patterns.

    Combustion of vast quantities of fossil fuels for power production

    is responsible for numerous environmental problems. The emis-

    sions from hydrocarbon combustion contain pollution agents

    including NOx, SOx, CO and CO2. These chemicals are connected to

    a variety of environmental degradation problems, including acid

    rain, smog, and greenhouse gases that contribute to climate

    change.

    To ensure that wind energy capacity is fully utilized, the

    turbine design must be optimized to operate in various wind

    conditions. A second law analysis can contribute to improving the

    wind turbine design, system efciency and power output. Signif-

    icant reductions in the environmental impact of energy generation

    methods can be achieved through efciency improvements via the

    second law. Wind power can provide a sustainable contribution to

    Fig. 11. Energy and exergy efciencies with varying V1, constant V1/V2: (a) point specic low velocity, (b) specic effective velocity, (c) effective velocity, (d) average low velocity.

    K. Pope et al. / Renewable Energy 35 (2010) 2102e2113 2111

  • society's energy needs. The minimal impact caused by wind

    turbine manufacturing, installation, maintenance, and operation

    can be further reduced through efciency improvements and

    enhanced design methodologies that use the second law of

    thermodynamics.

    5. Conclusions

    In this paper, the rst and second lawswere used to compare the

    performance of a variety of wind power systems. Exergy analysis

    was shown to allow a diverse range of geometric and operating

    designs to be compared with a commonmetric. The results indicate

    a 50e53% difference in rst and second law efciencies for the

    airfoil systems, and 44e55% for the VAWTs. Exergy is a useful

    parameter in wind power engineering, as it can represent a wide

    variety of turbine operating conditions, with a single uniedmetric.

    Through exergy methods, better site selection and turbine design

    can improve system efciency, decrease economic cost, and

    increase capacity of wind energy systems.

    Acknowledgements

    The authors gratefully acknowledge the nancial support of this

    research provided by Zephyr Alternative Power Inc. and the Natural

    Sciences and Engineering Council of Canada.

    References

    [1] WWEA. Wind turbines generate more than 1% of the global electricity.Charles-de-Gaulle-Str, 5, 53113 Bonn, Germany: World Wind Energy Associ-ation; 2008.

    [2] Rohatgi J, Barbezier G. Wind turbulence and atmospheric stability e theireffect on wind turbine output. Renewable Energy 1999;16(1):908e11.

    [3] Ameku K, Nagai BM, Roy JN. Design of a 3 kW wind turbine generator withthin airfoil blades. Experimental Thermal and Fluid Science 2008;32:1723e30.

    [4] Mayer C. The starting behaviour of a small horizontal-axis wind turbine.Renewable Energy 2001;22:411e7.

    [5] Wang F, Bai L, Fletcher J, Whiteford J, Cullen D. Development of small domesticwind turbine with scoop and prediction of its annual power output. Renew-able Energy 2008;33:1637e51.

    [6] Hirahara H, Hossain MZ, Kawahashi M, Nonomura Y. Testing basic perfor-mance of a very small wind turbine designed for multi-purposes. RenewableEnergy 2005;30:1279e97.

    [7] Dincer I, Rosen M. Exergy: energy, environment & sustainable development.Burlington, Massachusetts: Elsvier Ltd.; 2007.

    [8] Cengel YA, Boles MA. Thermodynamics: an engineering approach. 5th ed.Nevada, Reno: McGraw Hill; 2006.

    [9] Adeyinka O, Naterer GF. Entropy-based metric for component-level energymanagement: application to diffuser performance. International Journal ofEnergy Research 2005;29(11):1007e24.

    [10] Naterer GF. Entropy based design of fuel cells. Journal of Fuel Cell Science andTechnology 2006;3(2):165e74.

    [11] Uche J. Exergy costs and inefciency diagnosis of a dual-purpose power anddesalination plant. Journal of Energy Resources Technology 2006;128(3):186e92.

    Fig. 12. Specic exergy destruction with varying V1, constant V2/V1 for (a) point specic low velocity, (b) specic effective velocity, (c) effective velocity and (d) average low velocity.

    Table 3

    Various predicted rst and second law efciencies.

    h [%] J [%] J2 [%] JB [%] J2,B [%] hKE [%] JKE [%] JKE,2 [%]

    NACA 63(2)-215

    airfoil

    44 21 22 29 30 39 19 20

    FX 63-137 airfoil 47 23 22 30 28 59 29 27

    Savonius VAWT 18 17 10 55 32 94 87 51

    Zephyr VAWT 11 10 6 59 32 93 93 59

    K. Pope et al. / Renewable Energy 35 (2010) 2102e21132112

  • [12] Jassim R. Exergy analysis of carryover leakage irreversibilities of a power plantregenerative air heater. Proceedings of the Institution of Mechanical Engi-neers, Part A: Journal of Power and Energy 2004;218(1):23e32.

    [13] Sahin, AD, Dincer, I, and Rosen, MA. Development of new spatio-temporalwind exergy maps, Proceedings of ASME2006 Mechanical EngineeringCongress and Exposition, November 5e10, Chicago, Illinois, USA; 2006

    [14] Sahin AD, Dincer I, Rosen MA. Thermodynamic analysis of wind energy.International Journal of Energy Research 2006;30:553e66.

    [15] Selig MS. Wind tunnel aerodynamic tests of six airfoils for use on small windturbines. Journal of Solar Energy Engineering 2004;126(4):986e1001.

    [16] Martinez GG, Sorensen JN, Shen WZ. 3D boundary layer study on a rotatingwind turbine blade. Journal of Physics - Conference Series 2007;75(1):012032e9.

    [17] ANSYS Inc. Fluent 6.3 users guide. Canonsburg, Pennsylvania 15317, USA:ANSYS Inc, Southpointe, 275 Technology Drive; 2006.

    [18] ANSYS Inc. Gambit 2.3.16 users guide. Canonsburg, Pennsylvania 15317, USA:ANSYS Inc, Southpointe, 275 Technology Drive; 2005.

    [19] Saha U. Optimum design conguration of Savonius rotor through wind tunnelexperiments. Journal of Wind Engineering and Industrial Aerodynamics2008;96(8):1359e75.

    [20] Biswas A, Gupta R, Sharma KK. Experimental investigation of overlap andblockage effects on three-bucket Savonius rotors. Wind Engineering 2007;31(5):313e68.

    [21] Gupta R, Biswas A, Sharma KK. Comparative study of a three-bucket Savoniusrotor with a combined three-bucket Savoniusethree-bladed Darrieus rotor.Renewable Energy 2008;33:1974e81.

    [22] Pope K, Naterer GF, Tsang E. Effects of rotor-stator geometry on vertical axiswind turbine performance. Ottawa, Canada.: Canadian Society for MechanicalEngineers 2008 forum, June 5-8; 2008.

    [23] Zecher JB. A new approach to an accurate wind chill factor. Bulletin of theAmerican Meteorology Society 1999;80(9):1893e9.

    [24] Wilson RE, Lissaman PB, Walker SN. Aerodynamic performance of windturbines. Energy research and development administration, ERDA/NSF/04014e76/1; 1976

    K. Pope et al. / Renewable Energy 35 (2010) 2102e2113 2113


Recommended