chapter6 com

Upload: damienlerwick

Post on 02-Jun-2018

228 views

Category:

Documents


0 download

TRANSCRIPT

  • 8/10/2019 Chapter6 Com

    1/81

    CHAPTER 6 WIND LOADS C6-1

    CHAPTER 6 WIND LOADS

    Outline

    6.1 General

    6.1.1 Scope of application

    6.1.2 Estimation principle

    6.1.3 Buildings for which particular wind load or wind induced vibration is taken into account

    6.2 Horizontal Wind Loads on Structural Frames

    6.2.1 Scope of application

    6.2.2 Equation

    6.3 Roof Wind Load on Structural Frames6.3.1 Scope of application

    6.3.2 Procedure for estimating wind loads

    6.4 Wind Loads on Components/Cladding

    6.4.1 Scope of application

    6.4.2 Procedure for estimating wind loads

    A6.1 Wind Speed and Velocity Pressure

    A6.1.1 Velocity pressure

    A6.1.2 Design wind speed

    A6.1.3 Basic wind speed

    A6.1.4 Wind directionality factor

    A6.1.5 Wind speed profile factor

    A6.1.6 Turbulence intensity and turbulence scale

    A6.1.7 Return period conversion factor

    A6.2 Wind force coefficients and wind pressure coefficients

    A6.2.1 Procedure for estimating wind force coefficients

    A6.2.2 External pressure coefficients for structural frames

    A6.2.3 Internal pressure coefficients for structural frames

    A6.2.4 Wind force coefficients for design of structural frames

    A6.2.5 Peak external pressure coefficients for components/cladding

    A6.2.6 Factor for effect of fluctuating internal pressures

    A6.2.7 Peak wind force coefficients for components/cladding

    A6.3 Gust Effect Factors

    A6.3.1 Gust effect factor for along-wind loads on structural frames

    A6.3.2 Gust effect factor for roof wind loads on structural frames

    A6.4 Across-wind Vibration and Resulting Wind Load

    A6.4.1 Scope of application

  • 8/10/2019 Chapter6 Com

    2/81

    C6-2 Recommendations for Loads on Buildings

    A6.4.2 Procedure

    A6.5 Torsional Vibration and Resulting Wind Load

    A6.5.1 Scope of application

    A6.5.2 Procedure

    A6.6 Horizontal Wind Loads on Lattice Structural Frames

    A6.6.1 Scope of application

    A6.6.2 Procedure for estimating wind loads

    A6.6.3 Gust effect factor

    A6.7 Vortex Induced Vibration

    A6.7.1 Scope of application

    A6.7.2 Vortex induced vibration and resulting wind load on buildings with circular sections

    A6.7.3 Vortex induced vibration and resulting wind load on building components with circularsections

    A6.8 Combination of Wind Loads

    A6.8.1 Scope of application

    A6.8.2 Combination of horizontal wind loads for buildings not satisfying the conditions of

    Eq.(6.1)

    A6.8.3 Combination of horizontal wind loads for buildings satisfying the conditions of Eq.(6.1)

    A6.8.4 Combination of horizontal wind loads and roof wind loads

    A6.9 Mode Shape Correction Factor

    A6.9.1 Scope of application

    A6.9.2 Procedure

    A6.10 Response Acceleration

    A6.10.1 Scope of application

    A6.10.2 Maximum response acceleration in along-wind direction

    A6.10.3 Maximum response acceleration in across-wind direction

    A6.10.4 Maximum torsional response acceleration

    A6.11 Simplified Procedure

    A6.11.1 Scope of application

    A.6.11.2 Procedure

    A6.12 Effects of Neighboring Tall Buildings

    A6.13 1-Year-Recurrence Wind Speed

    Appendix 6.6 Dispersion of Wind Load

    References

  • 8/10/2019 Chapter6 Com

    3/81

    CHAPTER 6 WIND LOADS C6-3

    CHAPTER 6 WIND LOADS

    Outline

    Each wind load is determined by a probabilistic-statistical method based on the concept of

    equivalent static wind load, on the assumption that structural frames and components/cladding

    behave elastically in strong wind.

    Usually, mean wind force based on the mean wind speed and fluctuating wind force based on a

    fluctuating flow field act on a building. The effect of fluctuating wind force on a building or part

    thereof depends not only on the characteristics of fluctuating wind force but also on the size and

    vibration characteristics of the building or part thereof. These recommendations evaluate themaximum loading effect on a building due to fluctuating wind force by a probabilistic-statistical

    method, and calculate the static wind load that gives the equivalent effect. The design wind load can

    be obtained from the summation of this equivalent static wind load and the mean wind load.

    A suitable wind load calculation method corresponding to the scale, shape, and vibration

    characteristics of the design object is provided here. Wind load is classified into horizontal wind load

    for structural frames, roof wind load for structural frames and wind load for components/cladding. The

    wind load for structural frames is calculated from the product of velocity pressure, gust effect factor

    and projected area. Furthermore, a calculation method for horizontal wind load for lattice structural

    frames that stand upright from the ground is newly added. The wind load for components/cladding is

    calculated from the product of velocity pressure, peak wind force coefficient and subject area. For

    small-scale buildings, a simplified procedure can be applied.

    These recommendations introduce the wind directionality factor for calculating the design wind

    speed for each individual wind direction, thus enabling rational design considering the buildings

    orientation with respect to wind direction. Moreover, the topography factor for turbulence intensity is

    newly added to take into account the increase of fluctuating wind load due to the increase of

    fluctuating wind speed.

    Introduction of the wind directionality factor requires the combination of wind loads in along-wind,

    across-wind and torsional directions. Hence, it is decided to adopt the regulation for the combination

    of wind loads in across-wind and along-wind directions, or in torsional and along-wind directions

    explicitly. Furthermore, a prediction formula for the response acceleration of the building for

    evaluating its habitability to vibration, which is needed in performance design, and information of

    1-year-recurrence wind speed are newly added. Besides, information has been provided for the

    dispersion of wind load.

  • 8/10/2019 Chapter6 Com

    4/81

    C6-4 Recommendations for Loads on Buildings

    Notation

    Notations used in the main text of this chapter are shown here.

    Uppercase Letter

    A (m2): projected area at height Z

    RA (m2): subject area

    CA (m2): subject area of components/cladding

    0A (m2): whole plane area of one face of lattice structure

    FA (m2): projected area of one face of lattice structure

    B (m): building breadth

    1B (m): building length in span direction

    2B (m): building length in ridge direction0B , HB (m): width of lattice structure in ground and width at height H

    DB : background excitation factor for lattice structure

    1C , 2C , 3C : parameters determining topography factor gE and IE

    DC , RC , XC , YC : wind force coefficients

    'LC ,

    'TC : rms overturning moment coefficient and rms torsional moment coefficient

    eC : exposure factor, which is generally 1.0 and shall be 1.4 for open terrain with few

    obstructions (Category II). When wind speed is expected to increase due to local

    topography, this factor shall be increased accordingly.gC : overturning moment coefficient in along-wind direction

    'gC : rms overturning moment coefficient in along-wind direction

    fC : wind force coefficient. For horizontal wind loads, wind force coefficient DC defined

    in A6.2 with 9.0Z =k shall be used. For roof wind loads, wind force coefficient RC

    defined in A6.2 shall be used.

    peC : external pressure coefficient

    pe1C , pe2C : external pressure coefficients on windward wall and leeward wall

    pi

    C : internal pressure coefficient

    *piC : factor for effect of fluctuating internal pressure

    rC : wind force coefficient at resonance

    CC : peak wind force coefficient

    peC : peak external pressure coefficient

    D (m): building depth, building diameter, member diameter

    BD (m): building diameter at the base

    mD (m): building diameter at height of 3/2H

    E: wind speed profile factor

    HE : wind speed profile factor at reference height H

  • 8/10/2019 Chapter6 Com

    5/81

    CHAPTER 6 WIND LOADS C6-5

    IE : topography factor for turbulence intensity

    gE : topography factor for wind speed

    gIE : topography factor for turbulence intensity

    rE : exposure factor for flat terrain categories

    DF : along-wind force spectrum factor

    F: wind force spectrum factor

    DG : gust effect factor for along-wind load

    RG : gust effect factor for roof wind load

    H(m): reference height

    SH (m): height of topography

    TI (kgm2): generalized inertial moment of building for torsional vibration

    ZI : turbulence intensity at heightZ

    rZI : turbulence intensity at height Z on flat terrain categories

    DK : wind directionality factor

    L (m): span of roof beam

    SL (m): horizontal distance from topography top to point where height is half topography

    height

    ZL (m): turbulence scale at height Z

    M(kg): total building mass

    DM (kg): generalized mass of building for along-wind vibration

    LM (kg): generalized mass of building for across-wind vibration

    R : factor expressing correlation of wind pressure of windward side and leeward side

    DR : resonance factor for along-wind vibration

    LR : resonance factor for across-wind vibration

    TR : resonance factor for torsional vibration

    ReR : resonance factor for roof beam

    DS : size effect factor

    0U (m/s): basic wind speed

    1U (m/s): 1-year-recurrence 10-minute mean wind speed at 10m above ground over flat and

    open terrain

    1HU (m/s): 1-year-recurrence wind speed

    500U (m/s): 500-year-recurrence 10-minute mean wind speed at 10m above ground over

    flat and open terrain

    HU (m/s): design wind speed

    *LcrU ,

    *TcrU : non-dimensional critical wind speed for aeroelastic instability in across-wind

    and torsional directions

    rU (m/s): resonance wind speed*TU : non-dimensional wind speed for calculating torsional wind load

  • 8/10/2019 Chapter6 Com

    6/81

    C6-6 Recommendations for Loads on Buildings

    *rU : non-dimensional resonance wind speed

    DW (N): along-wind load at height Z

    LW (N): across-wind load at height Z

    TW (Nm): torsional wind load at height Z

    LCW (N): across-wind combination load

    RW (N): roof wind load

    SCW (N): wind load on components/cladding obtained by simplified method

    SfW (N): wind load on structural frames

    rW (N): wind load at height Z

    SX (m): distance from leading edge of topography to construction site

    Z(m): height above ground

    bZ , GZ (m): parameters determining exposure factor

    Lowercase Letter

    Dmaxa , Lmaxa (m/s2), Tmaxa (rad/s

    2): maximum response acceleration in along-wind,

    across-wind and torsional directions at top of building

    b (m): projected width of member

    f (m): rise

    1f (Hz): The smaller of Lf and Tf

    Df , Lf , Tf (Hz): natural frequency for first mode in along-wind, across-wind and torsional

    directions

    Rf (Hz): natural frequency for first mode of roof beam

    aDg , aLg , aTg : peak factors for response accelerations in along-wind, across-wind and

    torsional directions

    Dg , Lg , Tg : peak factors for wind loads in along-wind, across-wind and torsional

    directions

    h (m): eaves height

    1k : factor for aspect ratio

    2k : factor for surface roughness

    3k : factor for end effects

    4k : factor for three demensionality

    Ck : area reduction factor

    rWk : return period conversion factor

    Zk : factor for vertical profile for wind pressure coefficients or wind force coefficients

    l (m): smaller value of H4 and B , minimum value of H4 , 1B and 2B , member

    length

    a1l (m): smaller value of H and 1B

    a2l (m): smaller value of H and 2B

  • 8/10/2019 Chapter6 Com

    7/81

    CHAPTER 6 WIND LOADS C6-7

    Hq (N/m2): velocity pressure at reference height H

    Zq (N/m2): velocity pressure at height Z

    r(year): design return period

    Rer : coefficient of variation for generalized external pressure

    x (m) : distance from end of component

    Greek Alphabet

    : exponent of power law for wind speed profile

    : exponent of power law for vibration mode

    : load combination factor

    , L , T : mass damping parameter for vortex induced vibration, across-wind vibration

    and torsional vibrationD , L , T : mode correction factor for vortex induced vibration, across-wind vibration and

    torsional vibration

    D , L , T : critical damping ratio for first translational and torsional modes

    R : critical damping ratio for first mode of roof beam

    : solidity

    : mode correction factor of general wind force

    U : 0500 /UU

    : first mode shape in each direction

    D (Hz): level crossing factor

    (): roof angle, angle of attack to member

    S (): inclination of topography

    (kg/m3): air density

    S (kg/m3): building density which is )/( BmDHDM

    LT : correlation coefficient between across-wind vibration and torsional vibration

    6.1 General

    6.1.1 Scope of application

    (1) Target strong wind

    Most wind damage to buildings occurs during strong winds. The wind loads specified here are

    applied to the design of buildings to prevent failure due to strong wind. The strong winds that occur in

    this country are mainly those that accompany a tropical or extratropical cyclone, and down-bursts or

    tornados. The former are large-scale phenomena that are spread over about 1000km in a horizontal

    plane, and their nature is comparatively well known. Down-bursts are gusts due to descending air

    flows caused by severe rainfall in developed cumulonimbus. Since the scale of these phenomena are

    very small, few are picked up by the meteorological observation network. It is known that tornados are

  • 8/10/2019 Chapter6 Com

    8/81

    C6-8 Recommendations for Loads on Buildings

    small-scale phenomena several hundred meters wide at most having a rotational wind with a rapid

    atmospheric pressure descent. The characteristics of the strong wind and pressure fluctuation caused

    by tornados are not known. The number of occurrences of down-bursts and tornados is relatively large,

    but their probability of attacking a particular site is very small compared with that of the tropical or

    extratropical cyclones. However, the winds caused by down-bursts and tornados are very strong, so

    they often fatally damage buildings. These recommendations focus on strong winds caused by tropical

    or extratropical cyclones. However, the minimum wind speed takes into account the influence of

    tornadoes and down-bursts.

    (2) Wind loads on structural frames and wind loads on components/cladding

    The wind loads provided in these recommendations is composed of those for structural frames and

    those for components/cladding. The former are for the design of structural frames such as columns and

    beams. The latter are for the design of finishings and bedding members of components/cladding andtheir joints. Wind loads on structural frames and on components/cladding are different, because there

    are large differences in their sizes, dynamic characteristics and dominant phenomena and behaviors.

    Wind loads on structural frames are calculated on the basis of the elastic response of the whole

    building against fluctuating wind forces. Wind loads on components/cladding are calculated on the

    basis of fluctuating wind forces acting on a small part.

    Wind resistant design for components/cladding has been inadequate until now. They play an

    important role in protecting the interior space from destruction by strong wind. Therefore, wind

    resistant design for components/cladding should be just as careful as that for structural frames.

    6.1.2 Estimation principle

    (1) Classification of wind load

    A mean wind force acts on a building. This mean wind force is derived from the mean wind speed

    and the fluctuating wind force produced by the fluctuating flow field. The effect of the fluctuating

    wind force on the building or part thereof depends not only on the characteristics of the fluctuating

    wind force but also on the size and vibration characteristics of the building or part thereof. Therefore,

    in order to estimate the design wind load, it is necessary to evaluate the characteristics of fluctuating

    wind forces and the dynamic characteristics of the building.

    The following factors are generally considered in determining the fluctuating wind force.

    1) wind turbulence (temporal and spatial fluctuation of wind)

    2) vortex generation in wake of building

    3) interaction between building vibration and surrounding air flow

  • 8/10/2019 Chapter6 Com

    9/81

    CHAPTER 6 WIND LOADS C6-9

    Figure 6.1.1 Fluctuating wind forces based on wind turbulence and vortex generation in wake of

    building

    Fluctuating wind pressures act on building surfaces due to the above factors. Fluctuating wind

    pressures change temporally, and their dynamic characteristics are not uniform at all positions on the

    building surface. Therefore, it is better to evaluate wind load on structural frames based on overall

    building behavior and that on components/cladding based on the behavior of individual building parts.

    For most buildings, the effect of fluctuating wind force generated by wind turbulence is predominant.

    In this case, horizontal wind load on structural frames in the along-wind direction is important.

    However, for relatively flexible buildings with a large aspect ratio, horizontal wind loads on structural

    frames in the across-wind and torsional directions should not be ignored. For roof loads, the

    fluctuating wind force caused by separation flow from the leading edge of the roof often predominates.

    Therefore, wind load on structural frames is divided into two parts: horizontal wind load on structural

    frames and roof wind load on structural frames.

    Figure 6.1.2 Classification of wind loads

    (2) Combination of wind loads

    Wind pressure distributions on the surface of a building with a rectangular section are asymmetric

    even when wind blows normal to the building surface. Therefore, wind forces in the across-wind and

    torsional directions are not zero when the wind force in the along-wind direction is a maximum.

    wind load onstructural frames

    wind load on

    components/cladding

    horizontal wind load

    roof wind load

    along-wind load

    across-wind load

    torsional wind loadwind load

    simplifiedprocedure

    small-scale building

    wind load onstructural frames

    wind load oncomponents/cladding

    vibration direction vorticeswind turbulence

    a) fluctuating wind force caused by

    wind turbulence

    vibration direction

    b) fluctuating wind force caused by

    vortex generation in wake of building

  • 8/10/2019 Chapter6 Com

    10/81

    C6-10 Recommendations for Loads on Buildings

    Combination of wind loads in the along-wind, across-wind and torsional directions have not been

    taken into consideration positively so far, because the design wind speed has been used without

    considering the effect of wind direction. However, with the introduction of wind directionality,

    combination of wind loads in the along-wind, across-wind and torsional directions has become

    necessary. Hence, it has been decided to adopt explicitly a regulation for combination of wind loads in

    along-wind, across-wind and torsional directions.

    (3) Wind directionality factor

    Occurrence and intensity of wind speed at a construction site vary for each wind direction with

    geographic location and large-scale topographic effects. Furthermore, the characteristics of wind

    forces acting on a building vary for each wind direction. Therefore, rational wind resistant design can

    be applied by investigating the characteristics of wind speed at a construction site and wind forces

    acting on the building for each wind direction. These recommendations introduce the winddirectionality factor in calculating the design wind speed for each wind direction individually. In

    evaluating the wind directionality factor, the influence of typhoons, which is the main factor of strong

    winds in Japan, should be taken into account. However, it was difficult to quantify the probability

    distribution of wind speed due to a typhoon from meteorological observation records over only about

    70 years, because the occurrence of typhoons hitting a particular point is not necessarily high. In these

    recommendations, the wind directionality factor was determined by conducting Monte Carlo

    simulation of typhoons, and analysis of observation data provided by the Metrological Agency.

    (4) Reference height and velocity pressure

    The reference height is generally the mean roof height of the building, as shown in Fig.6.1.3. The

    wind loads are calculated from the velocity pressure at this reference height. The vertical distribution

    of wind load is reflected in the wind force coefficients and wind pressure coefficients. However, the

    wind load for a lattice type structure shall be calculated from the velocity pressure at each height, as

    shown in Fig.6.1.3.

    Figure 6.1.3 Definition of reference height and velocity pressure

    (5) Wind load on structural frames

    The maximum loading effect on each part of the building can be estimated by the dynamic response

    analysis considering the characteristics of temporal and spatial fluctuating wind pressure and the

    lattice type structurehigh-rise buildingdomehouse

    qZ

    qH

    qHqH

    Z

    H

    HH

  • 8/10/2019 Chapter6 Com

    11/81

    CHAPTER 6 WIND LOADS C6-11

    dynamic characteristics of the building. The equivalent static wind load producing the maximum

    loading effect is given as the design wind load. For the response of the building against strong wind,

    the first mode is predominant and higher frequency modes are not predominant for most buildings.

    The horizontal wind load (along-wind load) distribution for structural frames is assumed to be equal to

    the mean wind load distribution, because the first mode shape resembles the mean building

    displacement. Specifically, the equivalent wind load is obtained by multiplying the gust effect factor,

    which is defined as the ratio of the instantaneous value to the mean value of the building response, to

    the mean wind load. The characteristics of the wind force acting on the roof are influenced by the

    features of the fluctuating wind force caused by separation flow from the leading edge of the roof and

    the inner pressure, which depends on the degree of sealing of the building. Therefore, the

    characteristics of roof wind load on structural frames are different from those of the along-wind load

    on structural frames. Thus, the roof wind load on structural frames cannot be evaluated by the sameprocedure as for the along-wind load on structural frames. Here, the gust effect factor is given when

    the first mode is predominant and assuming elastic dynamic behavior of the roof beam under wind

    load.

    (6) Wind load on components/cladding

    In the calculation of wind load on components/cladding, the peak exterior wind pressure coefficient

    and the coefficient of inner wind pressure variation effect are prescribed, and the peak wind force

    coefficient is calculated as their difference. Only the size effect is considered. The resonance effect is

    ignored, because the natural frequency of components/cladding is generally high. The wind load on

    components/cladding is prescribed as the maximum of positive pressure and negative pressure for

    each part of the components/cladding for wind from every direction, while the wind load on structural

    frames is prescribed for the wind direction normal to the building face. Therefore, for the wind load on

    components/cladding, the peak wind force coefficient or the peak exterior wind pressure coefficient

    must be obtained from wind tunnel tests or another verification method.

    (7) Wind loads in across-wind and torsional directions

    It is difficult to predict responses in the across-wind and torsional directions theoretically like

    along-wind responses. However, a prediction formula is given in these recommendations based on the

    fluctuating overturning moment in the across-wind direction and the fluctuating torsional moment for

    the first vibration mode in each direction.

    (8) Vortex induced vibration and aeroelastic instability

    Vortex-induced vibration and aeroelastic instability can occur with flexible buildings or structural

    members with very large aspect ratios. Criteria for across-wind and torsional vibrations are provided

    for buildings with rectangular sections. Criteria for vortex-induced vibrations are provided for

    buildings and structural members with circular sections. If these criteria indicate that vortex-induced

    vibration or aeroelastic instability will occur, structural safety should be confirmed by wind tunnel

    tests and so on. A formula for wind load caused by vortex-induced vibrations is also provided for

    buildings or structural members with circular sections.

  • 8/10/2019 Chapter6 Com

    12/81

    C6-12 Recommendations for Loads on Buildings

    (9) Small-scale buildings

    For small buildings with large stiffness, the size effect is small and the dynamic effect can be

    neglected. Thus, a simplified procedure is employed.

    (10) Effect of neighboring buildings

    When groups of two or more tall buildings are constructed in proximity to each other, the wind flow

    through the group may be significantly deformed and cause a much more complex effect than is

    usually acknowledged, resulting in higher dynamic pressures and motions, especially on neighboring

    downstream buildings.

    (11) Assessment of building habitability

    Building habitability against wind-induced vibration is usually evaluated on the basis of the

    maximum response acceleration for 1-year-recurrence wind speed. Hence, these recommendations

    show a map of 1-year-recurrence wind speed based on the daily maximum wind speed observed atmeteorological stations and a calculation method for response acceleration.

    (12) Shielding effect by surrounding topography or buildings

    When there are topographical features and buildings around the construction site, wind loads or

    wind-induced vibrations are sometimes decreased by their shielding effect. Rational wind resistant

    design that considers this shielding effect can be performed. However, changes to these features during

    the buildings service life need to be confirmed. Furthermore, the shielding effect should be

    investigated by careful wind tunnel study or other suitable verification methods, because it is generally

    complicate and cannot be easily analyzed.

  • 8/10/2019 Chapter6 Com

    13/81

    CHAPTER 6 WIND LOADS C6-13

    Figure 6.1.4 Flow chart for estimation of wind load

    Start

    Outline of building

    6.1.3 Buildings to be designed for particular

    wind load or wind induced vibration

    (1) across-wind, torsional wind loads

    (2) vortex induced vibration,

    aeroelastic instability

    A6.12 Effects of neighboring tall buildings

    A6.1 Wind speed and velocity pressure

    A6.11 Simplified procedure

    A6.1.1 Velocity pressure

    A6.1.2 Design wind speed

    A6.1.3 Basic wind speed

    A6.1.4 Wind directionality factor

    A6.1.5 Wind speed profile factor

    A6.1.6 Turbulence intensity and turbulence

    scale

    A6.1.7 Return period conversion factor

    6.2 Horizontal wind load

    End

    A6.8 Combination of wind loads

    Wind tunnel experiment

    6.3 Roof wind load 6.4 Wind load on

    components/cladding

    Wind load on components/claddingWind load on structural frames

    A6.2.2 External wind pressure coefficient

    A6.2.3 Internal pressure coefficients

    A6.2.4 Wind force coefficients

    A6.2.1 Procedure for estimating wind force coefficients

    A6.2.5 Peak external pressure coefficients

    A6.2.6 Factor for effect of fluctuating internal

    pressuresA6.2.7 Peak wind force coefficient

    A6.4 Across-wind load

    A6.5 Torsional wind load

    A6.3.2 Gust effect factor

    for roof wind loads

    A6.3.1 Gust effect factor

    for along-wind loads

    A6.6 Horizontal wind loads on lattice

    structural frames

    A6.7 Vortex induced vibration

    A6.10 Response acceleration

    A6.13 1-year-recurrence wind speed

  • 8/10/2019 Chapter6 Com

    14/81

    C6-14 Recommendations for Loads on Buildings

    6.1.3 Buildings for which particular wind load or wind induced vibration need to be taken into

    account

    (1) Buildings for which horizontal wind loads on structural frames in across-wind and torsional

    directions need to be taken into account

    Horizontal wind loads on structural frames imply along-wind load, across-wind load and torsional

    wind load. Both across-wind load and torsional wind load must be estimated for wind-sensitive

    buildings that satisfy Eq.(6.1). Figure 6.1.5 shows the definition of wind direction, 3 component wind

    loads and building shape.

    along-wind

    across-windwind

    torsion

    Figure 6.1.5 Definition of load and wind direction

    Both across-wind vibration and torsional vibration are caused mainly by vortices generated in the

    buildings wake. These vibrations are not so great for low-rise buildings. However, with an increase in

    the aspect ratio caused by the presence of high-rise buildings, a vortex with a strong period uniformly

    generated in the vertical direction, and across-wind and torsional wind forces increase. However, with

    increase in building height, the natural frequency decreases and approaches the vortex shedding

    frequency. As a result, resonance components increase and building responses become large. In

    general, responses to across-wind vibration and torsional vibration depending on wind speed increase

    more rapidly than responses to along-wind vibration. Under normal conditions, along-wind responses

    to low wind speed are larger than across-wind responses. However, across-wind responses to high

    wind speed are larger than along-wind responses. The wind speed at which the degrees of along-wind

    response and across-wind response change places with each other differs depending on the height,

    shape and vibration characteristics of the building. The condition with regard to the aspect ratio of

    Eq.(6.1) has been established through investigation of the relationship between the magnitude of

    along-wind loads and across-wind loads for flat terrain subcategory II and a basic wind speed of 40m/s

    assuming 180kg/m3 building density, )024.0/(11 Hf = (Hz) natural frequency of the primary mode

    and 1% damping ratio for an ordinary building. Therefore, it is desirable to estimate across-wind and

    torsional wind loads even for buildings of light weight and small damping to which Eq.(6.1) is not

  • 8/10/2019 Chapter6 Com

    15/81

    CHAPTER 6 WIND LOADS C6-15

    applicable.

    Furthermore, for flat-plane buildings with small torsional stiffness or buildings with large

    eccentricity whose translational natural frequency and torsional natural frequency approximate each

    other, it is also desirable to estimate the torsional wind loads even where Eq.(6.1) is not applicable to

    those buildings.

    The discriminating conditional formula shown in this chapter was derived for a building with a

    rectangular plane. It is possible to apply Eq.(6.1) to a building with a plane that is slightly different

    from rectangular by regarding B and D roughly as projected breadth and a depth. For values of B

    and D changed in the vertical direction, the wind force acting on the upper part has a major effect on

    the response. Therefore, a representative value for the upper part should used for the computation.

    Under normal conditions, a value in the vicinity of 2/3 of the building height is chosen in most cases.

    The computation of Eq.(6.1) using a smaller value for the upper part yields a conservative value.(2) Vortex resonance and aeroelastic instability

    It is feared that aeroelastic instabilities such as vortex-induced vibration, galloping and flutter occur

    in buildings with low natural frequency and are high in comparison with their breadth and depth, as

    well as in slender members. The conditions for estimation of aeroelastic instability in both across-wind

    vibration and torsional vibration for building with rectangular planes as well as the conditions for

    estimation of vortex-induced vibrations for a building with a circular plane are given based on wind

    tunnel test results and the field measurement results1)-6)

    . The method for estimating the wind load for a

    building with a circular plan when vortex-induced vibration occurs is shown in A6.7. It may well be

    that vortex-induced vibration and aeroelastic instability will occur in a slender building with a

    triangular or an elliptical plan. Therefore, attention must be paid to this.

    The first condition required for estimating aeroelastic instability and vortex-induced vibration is the

    aspect ratio ( BDH/ or m/DH ). Aeroelastic instability as well as vortex-induced vibration does

    not occur easily in buildings with a small aspect ratio. Under this recommendation, the aspect ratio for

    estimating both aeroelastic instability and vortex-induced vibration was set to 4 or more and 7 or more,

    respectively. The second condition for estimating non-dimensional wind speed is ( BDfU/ or

    m/ fDU ). The occurrence of aeroelastic instability and vortex-induced vibration is dominated by the

    non-dimensional wind speed, which is determined by the representative breadth of the building, its

    natural frequency and wind speed. The non-dimensional critical wind speed for aeroelastic instability

    depends upon the mass damping parameter, which is determined by the side ratio, the turbulence

    characteristics of an approaching flow and the mass and damping ratio of a building. Thus, the

    non-dimensional critical wind speed with regard to the estimation of aeroelastic instability of a

    building with a rectangular plane was provided as the function for those parameters. The

    non-dimensional wind speed for vortex-induced vibration of a building with a circular plan is almost

    independent of this parameter. Therefore, the value for non-dimensional critical wind speed is fixed.

    The non-dimensional wind speed for estimating aeroelastic instability and vortex-induced vibration is

    set at 0.83(=1/1.2) times the non-dimensional critical wind speed. This is because it is known that

  • 8/10/2019 Chapter6 Com

    16/81

    C6-16 Recommendations for Loads on Buildings

    aeroelastic instability or vortex-induced vibration occurs within a period shorter than 10 min, which is

    the evaluation time for wind speed prescribed in this recommendation, and that the uncertainty of the

    non-dimensional wind speed including errors in experimental values is taken into account.

    Furthermore, the damping ratio of a building is required for the computation of the buildings mass

    damping parameter. It is thus recommended that the damping ratio of a building be estimated through

    reference to Damping in Buildings7)

    .

    6.2 Horizontal Wind Loads on Structural Frames

    6.2.1 Scope of application

    This section describes horizontal wind loads on structural frames in the along-wind direction. The

    along-wind load is generally composed of a mean component caused by the mean wind speed, aquasi-static component caused by relatively low frequency fluctuation and a resonant component

    caused by fluctuation in the vicinity of the natural frequency. For many buildings, only the first mode

    is taken into account as the resonant component. The procedure described in this section can estimate

    the equivalent static wind load producing the maximum structural responses (load effects of stress and

    displacement) using the gust effect factor. The equivalent static wind load is also divided into the mean

    component, quasi-static component and resonant component. Although the vertical profiles for these

    components are different from each other, it is assumed that all profiles similar to that of the mean

    component are provided.

    6.2.2 Estimation method

    Equation (6.4) for horizontal wind loads is derived from a gust effect factor method, which includes

    the effect of along-wind dynamic response due to atmospheric turbulence of approaching wind. The

    gust effect factor is a magnifying rate of the maximum instantaneous value to the mean building

    responses. Davenport, who first proposed the gust effect factor, calculated this factor based on the

    displacement at the highest position of a building8)

    . However, in these recommendations the gust effect

    factor based on the overturning moment of a base9), which can rationally estimate the design wind load

    for a building, was employed. Projected area A is the area projected from the wind direction for the

    portion concerned, as shown in Fig.6.2.1, and for wind load at a unit height being taken into account,

    projected area A becomes projected breadth B .

  • 8/10/2019 Chapter6 Com

    17/81

    CHAPTER 6 WIND LOADS C6-17

    D

    wind

    Figure 6.2.1 Projected area

    6.3 Roof Wind Load on Structural Frames

    6.3.1 Scope of application

    Roof wind loads on structural frames should be estimated from load effects of wind forces that act

    on roof frames. The properties of wind forces acting on roofs are influenced by the external pressures,

    which are affected by the behavior of the separated shear layers from leading edges, and the internal

    pressures, which are affected by the buildings permeability. This section describes equations to be

    applied to roof frames of buildings with rectangular plan without dominant openings, where the

    correlation between fluctuating external pressures and fluctuating internal pressures can be ignored.

    A light roof like a hanging roof might generate aerodynamically unstable oscillations. These

    oscillations may be generated in roof frames that satisfy the conditions of 3/ LfU

    and 15.0H

  • 8/10/2019 Chapter6 Com

    18/81

    C6-18 Recommendations for Loads on Buildings

    6.4 Wind Loads for Components/Cladding

    6.4.1 Scope of application

    Wind loads on components/cladding need to be designed for parts of buildings; finishings of roofs

    and external walls; bed members such as purlins, furring strips and studs; roof braces; and tie beams

    subject to strong effects of intensive wind pressure. These wind loads are also applied to the design of

    eaves and canopies.

    6.4.2 Procedure for estimating wind loads

    Wind loads on components/cladding are derived from the difference between the wind pressures

    acting on the external and internal faces of a building, and are calculated from Eq.(6.6). Peak wind

    force coefficients C

    C corresponding to the peak values of fluctuating net pressures, defined by thedifference between external and internal pressures, are given by Eq.(A6.15) for convenience. For

    buildings such as free-standing canopy roofs, where the top and bottom surfaces are both exposed to

    wind, the peak wind force coefficients CC are derived directly from the actual peak values of

    pressure differences, as shown in section A6.2.7.

    External pressure coefficients provided in the Recommendations correspond to the most critical

    positive and negative peak pressures on each part of a building irrespective of wind direction.

    Therefore, when the wind loads are calculated by considering the directionality of wind speeds, the

    peak pressure or force coefficients for each wind direction are needed, which should be determined

    from appropriate wind tunnel experiments or some other method12)

    .

    The subject areaACdepends on the item to be designed. When designing the finishing of roofs and

    external walls, the supported area of the finishing is used, and when designing the supports of the

    finishing, the tributary area of the supports is used.

    A6.1 Wind speed and velocity pressure

    A6.1.1 Velocity pressure

    The velocity pressure, which represents the kinetic energy per unit volume of the air flow, is the

    basic variable determining the wind loading on a building.. It corresponds to the rise in pressure from

    the free stream (atmospheric ambient static pressure) to the stagnation point on the windward face of

    the building, and is defined as ( ) 221 U , where U is the wind speed.It is only necessary to consider the velocity pressure as the basic variable of wind loading when

    static effects of the wind are examined. However, it is more appropriate to adopt wind speed as the

    basic variable when dynamic wind effects are under discussion. Thus, wind speed is adopted in the

    recommendations as the basic variable for calculating wind loading. The design velocity pressure, Hq ,

    which is based on the design wind speed HU at the reference height H , is defined in Eq.(A6.1).

  • 8/10/2019 Chapter6 Com

    19/81

    CHAPTER 6 WIND LOADS C6-19

    Air density varies with temperature, ambient pressure and humidity. However, the influence of

    humidity is usually neglected. In these recommendations, the air density is taken as 22.1= (kg/m3),

    which corresponds to a temperature of 15C and an ambient pressure of 1013 hPa.

    A6.1.2 Design wind speed

    The wind speed at a construction site is a function of its geographical location, orography or

    large-scale topographic features (e.g. mountain ranges and peninsulas) as well as the ground surface

    conditions (e.g. size and density of obstructions such as buildings and trees), and small-scale

    topographic features (e.g. escarpments and hills). The height above ground level is also a factor. Of

    these factors, the geographical location and large-scale topographical features are reflected in the

    values of basic wind speed 0U and wind directionality factor DK . The influences of surface

    roughness, small-scale topographical features and elevation are reflected in the wind speed profilefactor HE .

    Designers are required to decide the wind load level by considering the buildings social importance,

    occupancy, economic situation and so on. This is represented by the return period conversion factor

    rWk . The basic wind load defined in 2.2 is that corresponding to the 100-year-recurrence wind speed,

    which is calculated from Eq.(A6.2) by substituting 1rW =k . The wind directionality factor DK , a

    newly introduced parameter in this version, makes the design more rational by considering the

    dependencies of the wind speed, the frequency of occurrence of extreme wind and the aerodynamic

    property on wind direction. The wind directionality factor DK is affected by the frequency of

    occurrence and the routes of typhoons, climatological factors, large-scale topographic effects and so

    on.

    If the design ignores wind directionality effects, the design wind speed HU can be calculated by

    substituting 1D =K in Eq.(A6.2).

    A6.1.3 Basic wind speed

    The basic wind speed 0U is the major variable in Eq.(A6.2) for calculating the design wind speed.

    The wind speed at a construction site is influenced by the occurrence of typhoon and monsoon, the

    longitude and latitude of the location and large-scale topographical effects. The basic wind speed

    reflects the effects of these factors. The value of the basic wind speed corresponds to the

    100-year-recurrence 10-minute-mean wind speed over a flat and open terrain (category II) at an

    elevation of 10m. Figure A6.1.1 shows the procedure for making the basic wind speed map. As the

    first step of the procedure, data from different metrological stations were adjusted or corrected to

    reduce them to a common base considering the directional terrain roughness. Then extreme value

    analyses were conducted for mixed wind climates of typhoon winds and non-typhoon winds. For

    typhoon winds, a Monte-Carlo simulation based on a typhoon model was also conducted for each

    meteorological station in Japan. Although the analysis was conducted with consideration of wind

    directionality effect, the basic wind speed was considered as a non-directional value. Instead, the wind

  • 8/10/2019 Chapter6 Com

    20/81

    C6-20 Recommendations for Loads on Buildings

    directionality effect was reflected by introducing the wind directionality factor, which is defined as the

    wind speed ratio for a certain wind direction to the basic wind speed, as defined in A6.1.4.

    Records of wind speed and direction

    (for all meteorological stations from

    1961 to 2000)

    Modeling of typhoon pressure fields

    (based on data from 1951 to 1999)

    Monte Carlo simulation of typhoon

    winds (for 5000 years)

    Extreme wind probability distribution

    due to typhoons

    Extraction of independent storm

    (including the 2nd higher and less)

    Extreme wind probability distribution

    due to non-typhoon winds

    Synthesis of extreme value

    distributions

    Evaluation of terrain category

    (considering historical variation)

    Basic wind speed map

    Reduction to the common base

    Extreme value probability

    analysis for mixed wind climates

    Figure A6.1.1 Procedure for making basic wind speed map

    1) Data for analysis

    Data of wind speed, wind direction and anemometer height from the Japan Meteorological Business

    Support Center (Daily observation climate data from 1961-2000, Observation history at metrological

    stations) were used for analysis. The daily observation climate data from 1961-1990 and the

    Geophysical Review of 1951-1999 by the Japan Meteorological Agency were referred for modeling

    the pressure fields and tracks of typhoons, respectively. For homogenization of the wind speed records,

    data measured by different types of anemometers were corrected to those of propeller type

    anemometers13).

    2) Evaluation of directional terrain roughness and homogenization of wind speed

    The wind speed records at the meteorological stations were homogenized, that is to say, converted

  • 8/10/2019 Chapter6 Com

    21/81

    CHAPTER 6 WIND LOADS C6-21

    into data at a height of 10m over terrain category II by utilizing a method for evaluating the terrain

    roughness from the pseudo-gust factor (ratio of daily maximum instantaneous wind speed divided by

    daily maximum wind speed) and elevation of the measurement point14)

    . The details of the method are

    as follows. The pseudo-gust factors were first averaged according to the year and wind direction. Then,

    referring to the averaged pseudo-gust factors, a terrain roughness category was identified in which the

    same gust-factor was given using the profiles of mean wind speeds (defined in A6.1.5) and turbulence

    intensity (defined in A6.1.6). For this calculation, the terrain roughness category was treated as a

    continuous variable.

    Figure A6.1.2 shows examples of the annual variance of terrain roughness for four dominant wind

    directions measured at Fukuoka Meteorological Station, in which the symbols are for the calculated

    values and the lines are the results of linear approximation. The value of roughness category was

    assumed to be between I and V. This shows that the roughness category changes due to urbanizationand the roughness category varies with wind direction.

    Historical changes of the directional terrain roughness were utilized for homogenization of wind

    speed records at meteorological stations and calibration of wind speeds near the ground surface in the

    extreme value analysis and the typhoon model.

    Figure A6.1.2 Examples of evaluation for terrain roughness

    3) Extreme value analysis in mixed wind climates

    The extreme value analysis in mixed wind climates15)

    was applied to extreme wind data generated

    by different wind climates, for instance, typhoons and monsoons. In this method, the extreme wind

    records were divided into groups and independently fitted by extreme value distributions, and the

    combined distribution was obtained assuming the independency of each extreme distribution.

    Based on typhoon track data, the measuring records were divided into typhoon and non-typhoon

    winds, that is, if it was within 500 km of the typhoon center, the wind climate was considered as

    Flatterraincategories V

    IV

    III

    II

    I Flatterraincategories V

    IV

    III

    II

    I

    Flatterraincategories V

    IV

    III

    II

    I Flatterraincategories V

    IV

    III

    II

    I

  • 8/10/2019 Chapter6 Com

    22/81

    C6-22 Recommendations for Loads on Buildings

    typhoon, and otherwise as non-typhoon. The wind speed data measured in a typhoon area were

    analyzed by Monte-Carlo simulation based on a typhoon model to obtain the extreme value

    distribution, while those measured in a non-typhoon area were analyzed by the modified Jensen &

    Franck method16)

    in which wind speed data smaller than the highest value were also included as

    independent storms for analysis.

    4) Typhoon simulation technique

    In Japan, typhoons are the dominant wind climates generating strong winds that need to be taken

    into account in wind resistant design, due to their high wind speeds and large influence areas. An

    average of 28 typhoons occur annually, of which roughly 10% land. Typhoons sometimes do not pass

    near metrological stations, so severe wind damage may occur without large wind speeds being

    observed. In order to improve the instability of the statistical data (sampling error), a typhoon

    simulation method was adopted for evaluating the strong wind caused by typhoons.Figure A6.1.3 shows a general procedure of this typhoon simulation method. The pressure fields of

    typhoons are modeled by several parameters, i.e. central pressure depth, radius to maximum winds,

    moving speed, etc. The non-exceedance probability of strong wind in the target area is evaluated by

    generating virtual typhoons according to the results of statistical analysis of pressure field parameters.

    This Monte-Carlo simulation method is considered in recommendations of other countries. For

    example, in the ASCE17)standard, simulation is required as a principle for evaluation of the design

    wind speed in hurricane-prone regions. In this standard, the simulation results were adopted as the

    value of basic wind speed. In order to improve the accuracy of typhoon simulation18), correlations

    between gradient winds and near-ground winds and correlations among parameters of typhoon

    pressure fields in each area are considered.

  • 8/10/2019 Chapter6 Com

    23/81

    CHAPTER 6 WIND LOADS C6-23

    moving speed and

    direction

    radius of maximum wind

    wind speed fieldgradient wind

    surface wind

    central pressure

    depth

    rate of occurrenceinitial position

    Probability distributions

    statistics of

    historical typhoons

    return period

    pressure field

    rate of occurrence

    initial position

    moving velocity

    central pressure depth

    radius of maximum wind

    correlation of wind speed

    and direction based onobserved records

    Figure A6.1.3 General procedure for typhoon simulation

    The non-exceedance probability of the annual maximum wind speed caused by a typhoon was

    obtained from the typhoon simulation. For strong wind not caused by a typhoon, extreme value

    analysis was conducted on data observed from 1961-2000. The results obtained from typhoon and

    non-typhoon conditions were combined to evaluate the return period of annual maximum wind speed.

    Figure A6.1.4 shows an example of the maximum wind speed evaluated at K city.

    r

  • 8/10/2019 Chapter6 Com

    24/81

    C6-24 Recommendations for Loads on Buildings

    Figure A6.1.4 Example of maximum wind speed evaluated at K city

    5) Map of basic wind speedThe contour line of 100-year-recurrence wind speed was somewhat complicated even though the

    data obtained in 4) had been homogenized according to surface roughness, wind direction, etc. This

    was assumed to be due to the influences of local topography and structures surrounding the

    metrological station and the applicability of the homogenization models. To remove such local effects,

    spatial smoothing was conducted.

    In addition, the lower limit of wind speed was set to 30m/s. It is difficult to include the effects of

    tornado and downburst in the analysis.

    6) 100-year-recurrence wind speed in winter

    100-year-recurrence wind speed in winter is necessary for combination of wind loads and snow

    loads. As for the basic wind speed, 100-year-recurrence wind speed in winter reflects only the effects

    of large-scale topography. Figure A6.1.5 is a spatially smoothed wind speed map made for the

    100-year-recurrence wind speed at metrological stations during the snow season (from December to

    March). The procedure for making this map is the same as that for Fig.A.6.1.1, except that the typhoon

    simulation method is not used. Thus, the wind directionality factor should not be used ( 1D =K ) here.

    For return period factor rWk mentioned in A6.1.7, there are small differences in U among wind

    speeds in winter for different meteorological stations. An average value of 1.1U = can be applied

    for calculating rWk in Eq.(A6.12).

  • 8/10/2019 Chapter6 Com

    25/81

    CHAPTER 6 WIND LOADS C6-25

    Figure A6.1.5 100-year-recurrence 10-minutes mean wind speed at 10m above ground over a flat

    and open terrain in winter (m/s)

  • 8/10/2019 Chapter6 Com

    26/81

    C6-26 Recommendations for Loads on Buildings

    A6.1.4 Wind directionality factor

    Meteorological stations in Japan have approximately 70 years of records at most. However, the

    annual average number of typhoon landfalls in Japan is only three, so the number of typhoons

    included in the records of a particular site is very limited. When the records are divided into 8 sectors

    of azimuth, each sector have very few typhoon data, so sampling error is very large. Thus, typhoon

    effect should be considered when wind directionality factor is determined. In these recommendations,

    Monte-Carlo simulation for typhoon winds and statistical analysis on the non-typhoon observation

    data had been conducted to obtain the wind directionality factor.

    There are two types of wind directionality factors. One defines a wind directionality factor that

    changes with direction, as shown in BS6399.219) and AS/NZS 1170.220),21), except for the

    cyclone-prone regions. The other defines a constant reduction coefficient regardless of wind direction,

    as in the ASCE

    17)

    standard. For the latter, it is hard to reflect directional design wind speeds in designpractice. In these recommendations, wind directionality factor was defined for each direction as for the

    former type, so as to achieve reasonable wind resistant design.

    Wind directionality factor was provided on the assumption that the wind load is calculated

    according to the following procedure.

    (1) Where the aerodynamic shape factors for each wind direction are known from appropriate wind

    tunnel experiments, the wind directionality factor DK , which is used to evaluate wind loads on

    structural frames and components/cladding for a particular wind direction, shall take the same value as

    that for the cardinal direction whose 45 degree sector includes the wind direction. In this case, the

    wind tunnel experiments should be conducted for detailed change of directional characteristics for the

    aerodynamic shape factors of the structure.

    (2) Where the aerodynamic shape factors in A6.2 are used

    1) When assessing the wind loads on structural frames, two conditions are considered: whether or not

    the aerodynamic shape factors depend on wind direction.

    a) Where the aerodynamic shape factors are dependent on wind directions, four wind directions

    should be considered that coincide with the principal coordinate axis of the structure. If the wind

    direction is within a 22.5 degree sector centered at one of the 8 cardinal directions, the value of the

    wind directionality factor DK for this direction should be adopted (Fig.A6.1.6(a)). If the wind

    direction is outside the 22.5 degree sector, the larger of the 2 nearest cardinal directions should be

    adopted (Fig.A6.1.6(b)). For lattice structures, the effect of inclined wind on the wind force

    coefficient can be considered directly, so the same measures as for above rectangular cylinders are

    adopted for the 4-leg square plane (8 directions) and 3-leg triangular plane (6 directions).

    b) Where the aerodynamic shape factors are independent of wind directions, e.g. a structure that

    has a circular sectional plan, the wind directionality factor DK shall take the same value as for the

    cardinal direction whose 45 degree sector includes the wind direction.

    2) When assessing wind loads on cladding according to the peak wind pressure coefficient in A6.2,those obtained under the condition of 1D =K should be used for design because the maximum peak

  • 8/10/2019 Chapter6 Com

    27/81

    CHAPTER 6 WIND LOADS C6-27

    pressure coefficient of all directions is shown in these recommendations.

    The wind directionality factors for the 8 cardinal directions shown in Table A6.1.1 were originally

    obtained at 16 directions. When the 16 directional values are converted into 8 cardinal directional ones,

    the values are determined to be the maximum of those for the relevant direction and its two

    neighboring directions. Therefore, the value for a given direction represents the influence of a 67.5

    degree sector centered on that direction. For a building with rectangular horizontal section, the wind

    force coefficients for the wind directions normal to the building faces are given by these

    recommendations. When the wind direction considered is at an intermediate position between two

    cardinal directions shown in the table, the greater value of the two neighboring directions is adopted.

    This means that the value considers the influence from a 112.5 degree sector. In addition, considering

    the effects of tornado and downburst, which are difficult to take into account, the lower limit of wind

    directionality factor is given as 0.85.

    KD=0.9

    NW

    0.85

    W

    1.0

    SW

    0.95

    NE

    0.95

    E

    0.85

    SE

    0.85S

    0.9

    wind directionN

    0.9

    (a) Where the wind direction falls in a 22.5 degree sector as shown in Table A6.1.1

    N

    0.9larger value of 0.9 and 0.95

    KD= 0.95

    NW

    0.85

    W

    1.0

    SW

    0.95

    NE

    0.95

    E

    0.85

    SE

    0.85S

    0.9

    wind direction

    (b) Where the wind direction does not fall in a 22.5 degree sector as shown in Table A6.1.1

    Figure A6.1.6 Selection of the wind directionality factor (when using the wind force coefficient of

    buildings with rectangular horizontal sections defined in these recommendations)

  • 8/10/2019 Chapter6 Com

    28/81

    C6-28 Recommendations for Loads on Buildings

    Where wind directionality effects are not considered, this corresponds to the condition where the

    wind directionality factors equal unity for all directions. This leads a conservative design compared to

    the condition when the wind directionality effects are considered.

    Whether or not wind directionality effects are considered corresponds to whether or not wind

    directionality factors are adopted. As shown in Table A6.1.1, the wind directionality factors are less

    than unity, and are defined as values for evaluating 100-year-recurrence wind loads. It is possible to

    achieve a more rational design by considering the orientation of the building plan from the viewpoint

    of wind directionality factor. In other words, the wind loads are conservative if wind directionality

    factor is not considered. However, the amount of this overestimation depends on the orientation of the

    building, and not constant for all buildings. When wind directionality effects are considered, because

    the wind directionality factor is less than unity, the wind loads will be smaller than those predicted by

    conventional method, in which wind directionality is not taken into account. Designers should beconscious of the fact that safety level decreases when wind directionality factor is utilized.

    The wind directionality factors defined in these recommendations are valid only for locations near

    major metrological stations. The wind directionality factor defined in Table A6.1.1 can be applied to

    construction sites near metrological stations, but they cannot be applied to construction sites far from

    metrological stations and influenced by large-scale topography. For these situations, special

    consideration should be given, for instance, by not using the wind directionality factors i.e. by setting

    1D =K .

    A6.1.5 Wind speed profile factor

    (1) Effects of terrain roughness and topography on wind speed profile

    Wind speed near the ground varies with terrain roughness, i.e. buildings, trees, etc., and topography.

    The friction force from terrain roughness and the concentration or blockage effects from topography

    influence the atmospheric boundary layer from the ground to the gradient height. In the

    recommendations, the influence of surface roughness on the wind speed profile over flat terrain is

    expressed by rE , while the influence of small-scale topographical features is represented by gE .

    (2) Wind speed profile over flat terrain

    Terrain roughness causes a gradual decrease in wind speed toward the ground. The domain than is

    influenced by terrain roughness is called the boundary layer, where the wind speed profile changes

    with terrain roughness category. The boundary layer depth increases with fetch length, which means

    that the wind speed profile extends to a higher elevation downstream. In addition, the boundary layer

    tends to develop faster when the terrain is rougher.

    For a fully developed boundary layer, the velocity profile can be represented by a power law or a

    logarithmic law. The following power law is adopted in the recommendations:

    )(

    0

    0ZZ

    Z

    ZUU = (A6.1.1)

    where ZU (m/s) is the mean wind speed at height Z(m), 0ZU (m/s) is the mean wind speed at height

  • 8/10/2019 Chapter6 Com

    29/81

    CHAPTER 6 WIND LOADS C6-29

    0Z , and is the power law exponent.

    It has been realized from many observation data that the power law exponent becomes greater as the

    terrain becomes rougher.

    However, it is rare for the terrain roughness to be uniform over a long fetch. Roughness conditions

    usually vary. When the terrain roughness changes suddenly, a new boundary layer develops according

    to the new terrain roughness which gradually propagates with elevation and fetch, such that wind

    speeds above this new boundary layer remain unchanged after the roughness change. Thus, the wind

    speed profile corresponding to the new roughness condition can not be applied to the high elevation.

    This tendency is particularly obvious when the wind flows from the sea to city center, where the

    roughness changes suddenly from smooth to rough. After a fetch of approximately 3km (or 40H ) the

    new boundary layer is considered fully developed. Hence, in the recommendations, the roughness

    condition in the region of the smaller of 40H and 3km upstream from the construction site isconsidered when the roughness category, shown in Table A6.2 is to be determined.

    The influence of terrain roughness becomes smaller at higher elevations. In the recommendations, it

    is assumed that the design wind speed at GZ is not influenced by terrain roughness, and is

    considered constant for convenience. However, it does not mean that wind speeds at elevations greater

    than GZ are really constant. Since the boundary layer depth becomes greater when the terrain

    roughness increases, GZ is assumed to increase with terrain category, as shown in Table A6.3.

    However, GZ is defined just for the utilization of the power law for different terrain categories,

    because the velocity profile is actually unknown in detail at higher elevations. It is different from the

    boundary layer depth.

    CFD studies on the wind speed profile in urban area show that the wind speed below a certain

    height bZ does not follow the power law when the ratio of building plan area to regional area is over

    a few percent, as shown in Fig.A6.1.7. The wind speed profile here is complex due to nearby buildings.

    For heights below bZ , the wind speed at bZ is usually the maximum, so the wind speeds in this

    region are assumed to equal to that at bZ , which is defined in Table A6.3, in order to arrive at a safer

    design. For heights above bZ , the power law can approximate the mean wind speed profile.

    Zb

    height

    Figure A6.1.7 Mean wind speed profile in urban area

  • 8/10/2019 Chapter6 Com

    30/81

    C6-30 Recommendations for Loads on Buildings

    Figure 6.1.8 shows an example of mean wind speed profiles measured in natural wind22)

    , in which

    the wind speed profiles measured simultaneously at coastal and inland locations are compared. As

    mentioned before, the wind speed near the ground decelerates due to the inland terrain roughness. As a

    result, there is great difference between the wind speed profiles in the two locations.

    The exposure factor rE of the flat terrain, shown in A6.1.5(2) 2), is defined with the above

    considerations included. Figure A6.1.9 shows rE for each terrain category. The exposure factor is the

    ratio of wind speed at a given height Z for each terrain category to the wind speed at 10m over

    terrain roughness category II.

    Mean wind speed (m/s)

    Figure A6.1.8 Example of mean wind speed profiles measured simultaneously at the coast of Tokyo

    bay and a suburban residential area 12km away22)

    Exposure factor Er

    terrain category

    Figure A6.1.9 Exposure factor rE

  • 8/10/2019 Chapter6 Com

    31/81

    CHAPTER 6 WIND LOADS C6-31

    Figure A6.1.10 shows an example of terrain roughness categories.

    Terrain category I represents open sea or lake, or unobstructed coastal areas on land.

    Terrain category II is defined as terrain with scattered obstructions up to 10m high. Rural areas are

    representative. Low rise building areas also belongs to this category, if the building area ratio (total

    building plan area divided by regional area) is less than 10.0%.

    Terrain category III is characterized be closely spaced obstructions up to 10m high, or by sparsely

    spaced medium-rise buildings of 4-9 stories. Suburban residential areas, manufacturing districts, and

    wooded fields are typical of this category. The area where the building area ratio is between 10% and

    20%, or the building area ratio is larger than 10% while the high-rise building ratio (plan area of

    buildings higher than 4 stories divided by total area of buildings) is less than 30% belongs to this

    category. The example in Fig.A6.1.10(c) is an area with a building area ratio of 30% and a high-rise

    building ratio of 5-20%.

    (a) Terrain category I (b) Terrain category II

    (c) Terrain category III (d) Terrain category IV

    (e) Terrain category V

    Figure A6.1.10 Example of surface roughness (Photos provided by Kindai Aero Inc.)

  • 8/10/2019 Chapter6 Com

    32/81

    C6-32 Recommendations for Loads on Buildings

    Terrain category IV is mainly where many 4-9 story buildings stand. Local central cities are typical

    of this category. Areas with a building area ratio larger than 20%, and a high-rise building ratio larger

    than 30% belong to this category.

    In terrain category V, tall buildings of 10 or more stories are close together at a high density. Central

    regions of large cities such as Tokyo and Osaka belong to this category.

    In an area where the building purpose, floor area ratio and building coverage ratio are the same, the

    terrain can usually be considered uniform. Typically, in the wide area around the construction site,

    the terrain roughness is not usually identical. It is common for several terrain categories to co-exist.

    When the terrain roughness changes downstream, a new boundary layer gradually develops, and the

    developing process depends on whether the change is from smooth to rough or rough to smooth.

    Figure A6.1.11 illustrates approximately the development of a new boundary layer with a terrain

    roughness change from smooth to rough. When the terrain roughness changes from smooth to rough,the new boundary layer develops slowly, so the fully developed boundary layer over the new

    roughness can not be anticipated if the fetch downstream is not long enough. As a result, a wind speed

    profile corresponding to the new roughness category can not be adopted. Thus, if there is a terrain

    roughness change from smooth to rough within a distance of the smaller of 40H and 3km upstream

    of the construction site, the terrain category at the upstream region before the roughness change will

    be adopted as the terrain category for the construction site.

    Figure A6.1.11 Developing process of new boundary layer when terrain roughness changes from

    smooth to rough

    In determining the terrain category for a given wind direction, the upwind area inside a 45 degree

    sector within a distance of the smaller of 40H and 3km of the construction site will be counted.

    When there is a terrain roughness change upwind of the construction site, a weighting average of

    the wind speed profile on roughness and the fetch distance is conducted in AS/NZS 1170.2 20) to

    determine the exposure factor.

    However, in the recommendations, the overall terrain roughness in the upwind sector is adopted as

    the terrain category in this direction if there is no sudden roughness change. Generally, the wind load

    will be overestimated when a smoother surface roughness category is utilized.

    3 ~ 5kmSmooth Rough

    developing internalboundary layer

  • 8/10/2019 Chapter6 Com

    33/81

    CHAPTER 6 WIND LOADS C6-33

    For an urban area centered on a railway station, larger buildings are closely spaced near the station.

    Figure A6.1.12 shows an example of how to determine the terrain category if a construction site is

    near a railway station, in which the roughness changes from smooth to rough downstream. In this case,

    where there is a sudden roughness change within a distance of the smaller of 40H and 3km upwind

    of the construction site, the smoother terrain category upwind before the terrain roughness change will

    be selected.

    Wind

    Category I

    Category III

    smaller of

    40H and 3kmThe terrain roughness

    in this wind direction

    should be recognized as

    category I.

    Figure A6.1.12 Selection of terrain category (with terrain roughness change from smooth to rough)

    If the terrain roughness changes from rough to smooth, the terrain category after the terrain

    roughness change is selected. However, if there is a smooth area in a rough area, e.g. a park in a

    downtown area, it is sometimes necessary to consider the acceleration of wind speed near the ground

    downstream.

    Generally, careful consideration should be given in the determination of terrain category, because of

    the arbitrariness.

    (3) Topography factor

    When air flow passes escarpments or ridge-shaped topography as shown in Fig.A6.1.13, the flow is

    blocked on the front of the escarpment and the mean wind speed decreases. Then the flow starts to

    accelerate uphill, resulting in a mean wind speed larger than that of the flat terrain from the middle of

    the upwind slope to the top of the topographic feature. If the upwind slope is not large enough, the

    mean wind speed is larger than that over the flat terrain over a long region downstream of the hill top.

    However, if the upwind slope is sufficiently steep to establish separation downstream of the hill top,

    the wind speed downstream of the hill top near the ground is smaller than that of the flat terrain.

  • 8/10/2019 Chapter6 Com

    34/81

    C6-34 Recommendations for Loads on Buildings

    Figure A.6.1.13 Change of mean wind speed over an escarpment (thin solid line and thick solid line

    are for the mean wind speed over flat terrain and escarpments respectively)

    Equation A6.5 for the topography factor is based on the results of wind tunnel experiments of

    two-dimensional escarpments and ridge-shaped topography with different slopes23), 24), 25)

    . The

    experiments were carried out with an approach flow corresponding to terrain category II. The models

    corresponded to escarpments and ridge-shaped topography with heights between several tens of meters

    to 100m with smooth surfaces. The ratio of the mean wind speed over the escarpments to the

    counterpart over flat terrain was obtained from the experiments. The height Z in Eq.(A6.5) is the

    height from the local ground surface over the topographic feature. The slope angle is defined with the

    aid of the horizontal distance from the top of the topographic feature to the point where the height is

    half the topography height.

    Although, the wind speed decreases upwind of the escarpment and in the separation region

    downstream of steep topography, the topography factor in these regions is defined as 1 in the

    recommendations, as shown in Figs.A6.1.14 and A6.1.15, because only acceleration of wind speed is

    considered24)

    .

    Figure A6.1.14 Wind speed-up ratio over a two-dimensional escarpment with an inclination angle of

    60 degrees. The symbols are for the experimental results, and the solid lines are for

    Eq.(A6.5)

  • 8/10/2019 Chapter6 Com

    35/81

    CHAPTER 6 WIND LOADS C6-35

    Figure A6.1.15 Wind speed-up ratio over a two-dimensional ridge-shaped topography with

    inclination angle of 30 degrees. The symbols are for the experimental results, and

    the solid lines are for Eq.(A6.5)

    Tables A6.4 and A6.5 show the values of the parameters in Eq.(A6.5) for the escarpment and

    ridge-shaped topography determined from experiment. For a particular location and a particular slope

    angle, not shown in these tables, the topography factor can be obtained by linear interpolation. The

    following is an example of the procedure for calculating the topography factor of a 50-degree

    escarpment, at a location with a distance ss 6.1 HX = downstream of the top of the escarpment at a

    height s5.1 HZ= .

    Calculate the topography factor 1gE and 2gE at 1/ ss =HX and 2 for the inclination

    angle of 45 degrees from Eq.(A6.5), and then calculate the topography factor 12gE at

    6.1/ ss =HX by linear interpolation according to the following equation:

    2g1g12g 6.04.0 EEE +=

    Calculate the topography factor 34gE for the inclination angle of 60 degrees in the same

    way as for the inclination angle of 45 degrees.

    Conduct linear interpolation for topography factors 12gE and 34gE , with respect to the

    inclination angle to achieve the topography factor at an inclination angle of 50 degrees

    and 6.1/ ss =HX from the following equation.

    34g12gg3

    1

    3

    2EEE +=

    If the inclination angle is less than 7.5 degrees, the topography effect can be neglected.

    The topography factor calculated from Eq.(A6.5) is shown in Figs.A6.1.14 and A6.1.15 by a solid

    line. It agrees well with the experimental data at all sections with speedup..

    Equation (A6.5) is for the condition in which the air flow passes at right angles to the

    two-dimensional escarpments and ridge-shaped topography. However, strict two-dimensional hills do

    not exist, and flow does not always pass escarpments and ridge-shaped topography at right angles.

    However, even in these conditions, Eq.(A6.5) can be applied if the terrain extends a distance of several

  • 8/10/2019 Chapter6 Com

    36/81

    C6-36 Recommendations for Loads on Buildings

    times the height of the topographic feature in the traverse direction. In addition, as has been shown in

    experimental and CFD studies, the speed-up ratio of two-dimensional topography is greater than that

    of three-dimensional topography, and so application of Eq.(A6.5) to three-dimensional topography is

    conservative26)

    .

    Complex terrain may increase the wind speed in valleys, which is not considered in this equation. In

    such cases, it is recommended to investigate the topography factor by wind tunnel or CFD studies

    when the construction site is very complex.

    Figure A6.1.16 Interpolation procedure for calculating topography factor with inclination angle of

    50 degrees and 6.1/ ss =HX

    A6.1.6 Turbulence intensity and turbulence scale

    Natural wind speed fluctuates with time. The wind speed )(tU at a point, shown in Fig.A6.1.17,

    can be separated into a mean wind speed component U and a fluctuating wind speed component

    )(tu in the longitudinal direction as well as )(tv and )(tw in the cross wind directions. Usually, the

    longitudinal fluctuating wind speed component )(tu is important for design of buildings, so only the

    characteristics of )(tu are defined in the recommendations. For long-span structures such as bridges

    and for tall slender buildings, the vertical and lateral fluctuating wind-speed components )(tw and)(tv are also sometimes important.

  • 8/10/2019 Chapter6 Com

    37/81

    CHAPTER 6 WIND LOADS C6-37

    Figure A6.1.17 Mean wind and component of turbulence

    (1) Turbulence intensity

    1) On flat terrain

    Wind speed fluctuation can be expressed quantitatively by a statistical approach. Turbulenceintensity I indicates the turbulence level and it is defined in the following equation as the ratio of

    standard deviation of the fluctuating component u to the mean wind speed U.

    UI u

    = (A6.1.2)

    Turbulence is generated by the friction on the ground and drag on surface obstacles, and is

    influenced by the terrain roughness just as is the mean wind speed profile. Figure A6.1.18 shows the

    turbulence intensity observed in the natural wind and the recommended values calculated from

    Eq.(A6.8).

    eq.(A6.8) eq.(A6.8) eq.(A6.8) eq.(A6.8) eq.(A6.8)

    Turbulence intensityIrZ Turbulence intensityIrZ Turbulence intensityIrZ Turbulence intensityIrZ Turbulence intensityIrZ

    Figure A6.1.18 Observed turbulence intensity27)

    and recommended value

    The turbulence intensity ZI at height Z above the ground, is defined in Eq.(A6.7), in which the

    turbulence intensity rZI on flat terrain expressed in Eq.(A6.8), and the topography factor gIE , shown

    in Tables A6.6 and A6.7, is considered separately.

    2) Topography factor for turbulence intensity

    Not only the mean wind speed, but also the wind speed fluctuation is influenced by topography.

    Terrain category I Terrain category II Terrain category III Terrain category IV Terrain category V

  • 8/10/2019 Chapter6 Com

    38/81

    C6-38 Recommendations for Loads on Buildings

    Especially in the separation region, there is an obvious increase in the standard variation of the wind

    speed fluctuating component )(tu (fluctuating wind speed hereafter) compared to that on flat terrain,

    as Figs.A.6.1.19 and A6.1.20 show. Mean and fluctuating wind speed variation are closely related..

    The location of the maximum fluctuating wind speed generally corresponds to the location where the

    vertical gradient of mean wind speed is maximum. The region where the fluctuating wind speed is

    greater than the flat terrain counterpart is generally inside the separation region when the mean wind

    speed is smaller than that on flat terrain.

    Figure A6.1.19 Topography factor for fluctuating wind speed on an escarpment with inclination

    angle of 60 degrees. The symbols are for the experimental results, and the thick

    solid lines are for Eq.(A6.10).

    Figure A6.1.20 Topography factor for fluctuating wind speed on ridge-shaped topography with

    inclination angle of 30 degrees. The symbols are for the experimental results, and

    the thick solid lines are for Eq.(A6.10).

    In the recommendations, the topography factor for turbulence intensity is defined as the ratio of the

    topography factor for fluctuating wind speed to the topography factor for mean wind speed.

    Topography factor for fluctuating wind speed is defined in Eq.(A6.10), in which the values of the

    parameters besides 1C , 2C and 3C are identical to those in Eq.(A6.5) for the topography factor for

    mean wind speed. Equation (A6.10) is based on the results of wind tunnel experiments on escarpments

    and ridge-shaped topography, as for Eq.(A6.5). The experiments were carried out with an approach

    flow corresponding to terrain category II. The models corresponded to escarpments and ridge-shaped

  • 8/10/2019 Chapter6 Com

    39/81

    CHAPTER 6 WIND LOADS C6-39

    topography with a height of about 50m23), 24), 25)

    . Topography factors of mean wind speed and

    fluctuating speed are defined to be greater than 1 without considering the decrease in mean wind speed

    and fluctuating wind speed due to topography effects24)

    . However, when the topography factor for

    fluctuating wind speed is smaller than that for mean speed, the topography factor for turbulence

    intensity will be smaller than 1.

    Fluctuating wind speed near the ground becomes greater on the leeward slope of escarpments or

    ridge-shaped topography. In these regions the mean wind speed is smaller, which results in the

    maximum instantaneous wind speed being smaller than that for flat terrain in this area, as shown in

    Fig.A6.1.21. Because the decrease in mean wind speed is not considered in A6.1.5, the maximum

    instantaneous wind speed, and thus the wind load, is possibly overestimated in the separation region if

    only the topography factor of fluctuating wind speed is fitted to the experimental data. In order to

    reduce this possible overestimation, the actual topography factor for the fluctuating wind speed (

  • 8/10/2019 Chapter6 Com

    40/81

    C6-40 Recommendations for Loads on Buildings

    Eq.(A6.10) for the ridge-shaped topography because of the complexity of the change of fluctuating

    wind speed, but the coincidence is good where the topography factor of mean speed is larger than 1.

    Although Eq.(A6.10) is obtained from experiments carried out on a two-dimensional escarpment and

    ridge-shaped topography with the oncoming airflow passing at right angles, it can be applied to

    topography that extends a long distance in the transverse direction several times the height of the

    topography26)

    . However, if the construction site is in a complex terrain, it is necessary to investigate

    the topography factor for fluctuating wind speed by wind tunnel or CFD studies.

    (2) Power spectral density

    Power spectral density reflects the contribution to turbulence energy at each frequency. In the

    recommendations, a von Karman type power spectrum, expressed by Eq.(A6.1.3), is employed to

    express the power spectral density of fluctuating component of wind speed )(tu .

    6/52

    2

    uu

    })/(8.701{)/(4)(

    UfLULfF

    += (A6.1.3)

    where

    f : frequency

    u : standard deviation of fluctuating component of wind speed )(tu

    U: mean wind speed

    L : turbulence scale

    (3) Turbulence scale

    Equation (A6.11) is used as the turbulence scale ZL of the wind speed fluctuation )(tu at height

    Z.

    Turbulence scale is an important parameter in the power spectrum, expressed in Eq.(A6.1.3). It is

    the averaging length scale of the turbulence vortices. Figure A6.1.22 shows an example of a profile of

    turbulence scale, which can be expressed in Eq.(A6.11) independently of terrain category.

    eq.(A6.11)

    Figure A6.1.22 Observation of turbulence scale of wind speed fluctuation )(tu

  • 8/10/2019 Chapter6 Com

    41/81

    CHAPTER 6 WIND LOADS C6-41

    (4) Co-coherence

    Co-coherence of wind speed fluctuation ),,( yzu rrfR is evaluated using Eq.(A6.1.4). It expresses

    quantitatively the frequency-dependent spatial correlation of the wind speed fluctuation.

    +=

    U

    rkrkfrrfR

    2y

    2y

    2z

    2z

    yzu exp),,( (A6.1.4)

    where

    f : frequency

    yz , rr : distance between 2 points in the vertical and horizontal directions

    yz ,kk : decaying factors reflecting the degree of spatial correlation of wind speed in the

    vertical and horizontal directions

    U: mean wind speed averaged at two points

    It has been shown by observation that the decay factor is between 5-10.

    A6.1.7 Return period conversion factor

    Return period conversion factor rWk is defined as the ratio of the r-year-recurrence wind speed

    rU to the 100-year-recurrence basic wind speed 0U . In these recommendations, the maximum wind

    speed corresponding to an r-year return period should be estimated using Eq.(A6.1.5), assuming a

    Gumbel distribution for annual-maximum wind speeds.

    b

    r

    r

    a

    U +

    =

    1

    lnln1

    r (A6.1.5)

    where a and b are coefficients. Return period conversion factor krWis calculated approximately in

    Eq.(A6.12) by using the parameter U , which is the ratio of the 500-year-re