Sie sind auf Seite 1von 40

C6-42 Recommendations for Loads on Buildings

A6.2 Wind force coefficients and wind pressure coefficients

A6.2.1 Procedure for estimating wind force coefficients


Wind force coefficients and wind pressure coefficients depend on building shape, building surface
condition, terrain condition and local topography at the construction site. Therefore, they should be
determined from wind tunnel experiments that properly simulate full-scale conditions. However, the
coefficients for buildings with regular shapes can be estimated from the procedure described in this
section. The coefficients are divided into two categories, one for the design of structural frames and
the other for the design of building components/cladding, because the wind effects on structural
frames and components/cladding are quite different from each other.
Wind force coefficients and wind pressure coefficients are generally defined in terms of the velocity
pressure qH evaluated at the reference height H . For lattice structures and members, the wind force
coefficients are defined in terms of the velocity pressure q Z evaluated at the height Z where the
members under consideration are placed.
The aspect ratio H / B is generally large for tall buildings, such as H > 45 m, for example, while
it is generally small, smaller than 1.0 in many cases, for lower buildings. The flow field around a
building changes with the aspect ratio, which results in a significant change in the wind force and
pressure coefficients. Therefore, two different procedures are provided for estimating the wind force
coefficients for buildings with H > 45 m and those with H 45 m.
The sign of the wind pressure coefficient indicates the direction of the pressure on the surface or
element; positive values indicate pressures acting towards the surface and negative values pressures
acting away from the surface (suction). In the case of curved roofs, the direction of wind pressure
varies with location, as shown in Fig. A6.2.1. The wind forces on buildings and structures are the
vector sum of the forces calculated from the pressures acting on surfaces such as walls and roofs or on
structural elements.
Wind force coefficients C D for estimating horizontal wind loads on structural frames are generally
given by the difference between the wind pressure coefficients, C pe1 and C pe2 , on the windward and
leeward faces, as shown in Eq.(A6.13); the exception is that for buildings with circular sections, where
the resultant wind force coefficients are provided. Similarly, the wind force coefficients C R for
estimating roof wind loads on structural frames are generally given by the difference between the
external and internal pressure coefficients, C pe and C pi , on the roof, as shown in Eq. (A6.14),
except for open roofs. The wind pressure coefficients are space- and time-averaged values where the
averaging duration is 10 minutes. The averaging area depends on the building shape. The wind force
coefficients C D for estimating horizontal wind loads on lattice structures are given as a function of
the solidity . The wind force can also be calculated by using the wind force coefficients for
individual members provided in A6.2.4(5).
The peak wind force coefficients C C for the design of components/cladding are generally given
*
by the difference between the peak external pressure coefficient C pe and the factor C pi for the
CHAPTER 6 WIND LOADS C6-43

effect of fluctuating internal pressures, except for open roofs, in which the value of C C is provided.
The values of C pe (and C C in the open roof case) are determined from the most critical positive
*
and negative peak values irrespective of wind direction. Note that the factor C pi for the effect of
fluctuating internal pressures is not the actual peak internal pressure coefficient C pi but an equivalent
value producing the peak wind force coefficient C C when combined with the peak external pressure
coefficient C pe .
The wind force coefficients and wind pressure coefficients given in this section are all for isolated
buildings and are obtained from the results of wind tunnel experiments. When nearby buildings are
expected to influence the wind forces and pressures, it is necessary to carry out wind tunnel
experiments or other special researches to determine the coefficients12).

Figure A6.2.1 External pressure on a building with a vaulted roof in a wind parallel to the gable
walls

A6.2.2 External pressure coefficient for structural frames


(1) External pressure coefficients C pe for buildings with rectangular sections and heights greater than
45m
External pressure coefficients on the windward and leeward walls of buildings with rectangular
sections have the following features:
1) External pressure coefficients on windward walls are nearly proportional to the velocity pressure of
the approach flow, except for areas near the top and bottom of the building. In the top and bottom
areas, the external pressure coefficient is almost independent of height.
2) External pressure coefficients on leeward walls are negative and almost independent of height.
Based on these features, the vertical distribution of external pressure coefficients on windward walls
are assumed to be proportional to the factor for vertical profile ( k Z ) provided in Table A6.8, while
those on leeward walls are assumed constant regardless of height. The external pressure coefficients
on leeward walls decrease with increase in side ratio D / B . This feature is related to the behavior of
the separated shear layer from the windward edge and is reflected in the value of C pe2 . The aspect
ratio H / B of high-rise buildings with H > 45 m is in the range from 1 to 8 in most cases. In this
range, the effect of H / B on the wind pressure coefficients is not significant. Therefore, the pressure
C6-44 Recommendations for Loads on Buildings

coefficients are provided independently of H / B .


The external pressure coefficients on roofs are determined from the results of various wind tunnel
experiments20), 28)
as well as on the provisions of international codes and standards. Although
flat-roofed buildings have parapets in many cases, their effect on the pressure coefficients is not
considered here. A reduction factor for external pressure coefficients on roofs with parapets is
provided in Eurocode28).
(2) External pressure coefficient C pe for buildings with rectangular sections and heights less than or
equal to 45m
1) Buildings with flat, gable and mono-sloped roofs
External pressure coefficients are influenced by many factors, such as roof shape, roof angle and
flow condition. The coefficients in this section are estimated from the results of wind tunnel
experiments on buildings with rectangular sections and reference heights less than or equal to 45m.
The roof shapes under consideration are flat, gable and mono-sloped. When the roof angle is less than
or equal to 10 degrees, the roof can be regarded as a flat roof.
The roof and walls are divided into several zones, and the external pressure coefficients for these
zones are provided in Table A.6.9(1) as a function of building configuration parameters ( B / H ,
D / H and ). The external pressure coefficient for each zone is estimated from the spatially
averaged pressure over the zone for a range of wind directions, the center of which is normal to the
wall. Both positive and negative values are provided for the external pressure coefficient for zone Ru ,
because the pressure coefficient becomes both positive and negative due to a small change in
experimental conditions. It is necessary to combine these values with those for the other zones when
the stresses in the members are calculated.
The net wind forces on windward eaves become very large, because negative pressures act on the
top surface and positive pressures on the bottom surface of the eaves. In this case, the external
pressure coefficient on the bottom surface is approximately equal to that on the windward wall just
bellow the eaves.
2) Buildings with vaulted roofs
The external pressure coefficient for a building with a curved surface generally depends on the
shape and surface roughness of the building, the flow conditions and the Reynolds number. Buildings
with vaulted roofs, however, are immersed in very turbulent flows. Furthermore, such buildings have
walls in most cases and therefore the flow tends to separate at the windward edge. These features
suggest that the external pressure coefficients on vaulted roofs are less sensitive to surface roughness
and the Reynolds number than those on circular cylindrical structures, as shown in A6.2.4(1). The
external pressure coefficients C pe in Table A6.9(2) are determined from the results of a wind tunnel
eperiment32), 33)
that focuses on medium-scale buildings in urban areas. The effects of surface
roughness are not considered in the experiment.
For a wind normal to the gable wall (wind direction W1 ), the building shape is represented by the
rise/width ratio f / B and the eaves-height/width ratio h / B . However, for a wind parallel to the
CHAPTER 6 WIND LOADS C6-45

gable wall (wind direction W2 ) it is represented by the rise/depth ratio f / D and the
eaves-height/depth ratio h / D . In both cases, the roof is divided into three zones. However, the zone
definitions vary because of the difference between the flow patterns of the two wind directions. For
wind direction W1 , the definition of zones is similar to that for flat, gable and mono-sloped roofs. For
wind direction W2 , however, the definition is similar to that for spherical domes.
The external pressure coefficient corresponds to the area-averaged value and the design wind load is
assumed constant over each zone. When h / B = 0 and f / B = 0 or when h / D = 0 and f / D = 0 ,
roof level coincides with ground level. The coefficients for these cases, which have no physical
meaning, are provided to make interpolation possible.
The external pressure coefficients on walls are determined in the same way as for buildings with flat,
gable and mono-sloped roofs.
3) Spherical domes
In the same manner as for buildings with vaulted roofs, the external pressure coefficients for
spherical domes are determined from the results of a wind tunnel experiment34). Since the counter lines
of mean pressure coefficients on a spherical dome are almost perpendicular to the wind direction, the
dome surface is divided into four zones ( Ra to Rd ), as shown in Table A6.10, and the external
pressure coefficient C pe for each zone is given by spatially averaging the mean external pressure
coefficient over the zone. The building shape is represented by the rise/span ratio f / D and the
eaves-height/span ratio h / D . The values of C pe for five f / D ratios and three h / D ratios are
provided in Table A6.10. Linear interpolation can be used for values of f / D and h / D other than
shown. Both positive and negative values of C pe are provided for zone Ra . The value for h / D = 0
and f / D = 0 are again provided for interpolation.
The wind force coefficients for walls can be obtained from Table A6.12 by substituting h for H .

A6.2.3 Internal pressure coefficients for structural frames


Internal pressures are significantly influenced by the following factors:
a) distribution of external pressures
b) openings and gaps in building envelope
c) internal volume of building
d) openings and gaps in internal partitions
e) operation of air-conditioners
f) distortion of walls and/or roofs
g) air temperature
h) damage to building envelope
In general, buildings have many gaps and openings, such as ventilating openings, etc., in their
envelopes. Air leaks through these gaps and openings due to differences between external and internal
pressures. The internal pressure is determined by applying the mass conservation principle to the air in
the internal volume. For instance, a dominant opening in the windward wall may produce positive
C6-46 Recommendations for Loads on Buildings

internal pressures, whereas one in a side or leeward wall may produce negative internal pressures.
Moreover, the internal pressure fluctuates and its characteristics depend on the relationship between
the size of the openings and the internal volume of the building. In this section, internal pressure
coefficients for buildings without dominant opening are provided based on the results of a series of
computations, in which it is assumed that the internal pressures are significantly influenced by factors
a) and b) mentioned above. That is, the values of Cpi in Table A6.11 are provided based on the
calculations of the mean internal pressures for various building configurations, assuming that the gaps
and openings are uniformly distributed over the external walls and the internal pressure is caused by
external pressures acting at the locations of the gaps and openings.
When the influence of other factors is assumed to be significant, it should be taken into account for
evaluating the internal pressure coefficient. For instance, when the internal volume is divided by
airtight partitions, the influence of factor d) is significant. When powerful air-conditioners are in
operation, the influence of factor e) is significant. In buildings with flexible roofs and/or walls, such as
membrane structures, the influence of factor f) is significant. When glass windows on the windward
face are broken by wind-borne debris in strong winds, the internal pressure is suddenly increased by
winds blowing into the building. This often results in failures of roof structures. In such cases, factor
h) should be considered appropriately.

A6.2.4 Wind force coefficients for design of structural frames


(1) Wind force coefficients C D for buildings with circular sections
Wind force coefficients for cylinders are affected by the Reynolds number, flow condition, aspect
ratio H / D , surface roughness of the cylinders, and other factors. Figure A6.2.236) shows the variation
of drag coefficient C D on a two dimensional smooth cylinder in a uniform flow with Reynolds
number Re ( = UD / , where U , D and are wind speed, cylinder diameter and kinematic
viscosity coefficient of flow, respectively). For wind, the Reynolds number is approximately given by
Re 7UD 10 4 , where U and D are expressed in units of m/s and m, respectively. It is found
from Fig. A6.2.2 that C D changes significantly with Re in the range from 2 105 to 5 106. The flow
around a cylinder is usually classified into four regimes, i.e. subcritical, critical, supercritical and
transcritical, as shown in Fig. A6.2.2. Since the Reynolds number of the flow around buildings in
strong winds is in the transcritical regime, the provision of C D in A6.2.4(1) is based on the values of
the drag coefficients in this regime.
The aspect ratio and surface roughness of the cylinder also affect the drag coefficient. In particular,
the effect of surface roughness is significant in the transcritical regime. In Table A6.12 the effects of
aspect ratio and surface roughness are represented by k1 and k 2 , respectively37).
The external pressure coefficients C pe on roofs are given in Table A6.10 assuming that f / D = 0
and h / D = 1 .
CHAPTER 6 WIND LOADS C6-47

Figure A6.2.2 Plots of drag coefficient C D on a two-dimensional cylinder with very smooth
surface as a function of Reynolds number Re 36)

(2) Wind force coefficients C R for free roofs with rectangular base
For free roofs where strong wind can flow under the roof, high fluctuating pressures act on both the
top and bottom surfaces. It is reasonable to evaluate the net wind force coefficients directly, not from
the wind pressure coefficients on the top and bottom surfaces, because the correlation between
fluctuating wind pressures on both surfaces is higher than that for enclosed buildings.
The wind force coefficients in Table A6.13 can be used for small-scale buildings, to which the
simplified method (A6.11) is applied, because the coefficients are determined from the results of wind
tunnel experiments on free roofs with H < 10 m. For gable ( > 0 ) and troughed roofs ( < 0 ),
previous studies have shown the most critical peak wind force coefficients on the windward and
leeward areas irrespective of wind direction. Since the tested roof angle is limited to the range of
| | 30 , the provision is also limited to that range.
The wind force coefficients are regulated for a clear flow case where there are no obstructions under
the roof. The flow pattern around a roof is significantly affected by obstructions under it. If there is
any obstruction whose blockage ratio is larger than approximately 50%, the wind pressure on the
bottom surface may increase significantly, resulting in a significant increase in the net wind force on
the roof. In such a case, it is necessary to evaluate the wind force coefficients from wind tunnel
experiments and so on.
(3) Wind force coefficients C D for lattice structures
The size of individual lattice structure members is generally much smaller than the width of the
structure, and they are arranged symmetrically. Therefore, it is assumed that the only wind force acting
on a plane of the structure is drag. Total drag can be estimated as the summation of the drags on each
member of the structure. Since the flow around a member depends only on the characteristics of the
local flow around it, drag is proportional to the velocity pressure at the height of the member. Based on
these features, the following two methods are often used for estimating the wind force on lattice
C6-48 Recommendations for Loads on Buildings

structures. One is to multiply the wind force coefficient, given as a function of the solidity of the
plane, by the projected area of the plane. The other method39) is to sum the wind forces on all members,
which is given by the product of the wind force coefficient C D of each member and its projected
area. For any method, the solidity should be small. In the Recommendations, the former method is
used and the wind force coefficient C D is provided only for 0.6 .
The wind force coefficient is represented as a function of the solidity , the plan of the structure
and the cross section of the member. The solidity is defined as the ratio of the projected area AF
of the plane to the whole plane area A0 = ( Bh) of the structure. The value of is calculated for
each panel of the lattice structure when the wind direction is normal to the plane. In the calculation,
the areas of the leeward lattice members and the appurtenances are not included. The wind forces on
the appurtenances can be estimated from the provision of C D for members (Table A6.16) or from
wind tunnel experiments and they are added to the wind force on the structure.
Table A6.14 provides the wind force coefficients C D for lattice structures with square and
triangular plan shapes, which consist of angles or circular pipes. The wind force coefficient C D for
the triangular shape in plan is the same for the two wind directions shown in the table. When the
members are circular pipes, the wind force coefficients C D for the members are affected by the
Reynolds number. The provisions are based on the value in the subcritical Reynolds number regime. In
strong winds, the value of C D may become smaller than that given in the provisions due to the effect
of the Reynolds number. However, this effect is not considered here.
When the plan of the structure and/or the cross section of the member are different from those in
Table A6.14, the wind loads on the structure can be estimated by using the wind force coefficients of
the members given in Table A6.16 together with the local velocity pressure. However, the solidity
of the structure is required to be less than 0.6.
(4) Wind force coefficients C D for fences on ground
Wind force coefficients C D for fences on the ground are defined as a function of the solidity
in the same manner as those for lattice structures. The value of C D for = 0 in Table A6.15 is
introduced to obtain intermediate values of C D for 0 < < 0.2 . Wind load for a fence can be
calculated according to the simplified procedure using C D and the projected area A, which is defined
as the whole area multiplied by .
(5) Wind force coefficients C for components
Wind force coefficients C for components are determined from wind tunnel experiments with
two-dimensional models in a smooth flow. The values of C can be applied to line-like members less
than approximately 50cm wide, but should not be applied to ordinary buildings. In some cases, the
value of C in the across-wind direction becomes relatively large when the wind direction deviates
only a little from the normal direction. In such cases, two values of C ( 0.6) are provided in Table
A6.16.
Wind force coefficients for components may also be used for calculating the wind loads on lattice
structures, together with the local velocity pressure q Z at height Z of the member under
CHAPTER 6 WIND LOADS C6-49

consideration. The wind load on a component is given by the product of q Z , C , (1 + 7 I Z ) and bl


( bl for nets), where I Z is the turbulence intensity at height Z (see Eq.(A6.7)).

A6.2.5 Peak external pressure coefficients for components/cladding


(1) Peak external pressure coefficients C pe for buildings with rectangular sections and heights
greater than 45 m
Peak external pressure coefficients for components/cladding correspond to the most critical positive
and negative peak pressure coefficients irrespective of wind direction. Positive pressures occur on
windward walls, and their characteristics are affected by the vertical profile of the approach flow. On
the other hand, negative pressures (suctions) occur on side and leeward walls, and their characteristics
are not significantly affected by the vertical profile of the approach flow; that is, the vertical
distribution is nearly uniform. Large negative pressures occur near the windward edges of sidewalls
due to flow separation from the edge. The peak external pressure coefficients provided in Table A6.17
are determined from the results of wind tunnel experiments41)-44). These coefficients are given by the
product of the external pressure coefficients influenced by the profile of the mean wind speed and the
gust effect factor influenced by the profile of the turbulence intensity. Therefore, the positive external
peak pressure coefficients are affected by the terrain category. However, negative external peak
pressures are almost independent of terrain category.
For tall buildings with recessed or chamfered corners, the negative peak pressures are influenced by
the size of the recess or chamfer. The values of C pe for such buildings are also determined from the
results of wind tunnel experiments42), 43). The values in Table A6.17 can also be used for buildings with
more than one recessed or chamfered corner.
Peak external pressure coefficients for roofs are provided only for flat roofs. For diagonal wind
directions, very large suctions are induced near windward corners due to the generation of conical
vortices. However, the large suction zone is limited to a relatively small area45). Therefore, the use of
such large peak pressure coefficients for large components may overestimate the design wind loads. In
order to consider the subject area of components/cladding in zone Rc , an area reduction factor kC for
roofs is introduced.
The provisions are applicable to buildings with aspect ratios H / B less than or equal to 8, because
the values are based on wind tunnel experiments on such buildings.
When a building is constructed on an escarpment or a ridge-shaped topography, the approach wind
is affected by the local topography, and therefore the positive peak pressure coefficients may change
significantly. Since wind speeds near the ground are increased by such local topography, the vertical
distribution of positive peak external pressure coefficients becomes nearly uniform. In such cases,
positive peak external pressures can be calculated by using the values of k Z and I Z at the reference
height H . This simplified method overestimates the wind loads to some degree in most cases.
However, for terrain category I, it may underestimate the positive peak external pressures. In this case,
investigations by wind tunnel experiments are recommended.
C6-50 Recommendations for Loads on Buildings

(2) Peak external pressure coefficient C pe for buildings with rectangular sections and heights less
than or equal to 45 m
1) Buildings with flat, gable and mono-sloped roofs
For estimating peak pressure coefficients for components/cladding of low-rise buildings, the subject
area is assumed to be 1 m2 as a typical value. Positive peak external pressure coefficients are given as
a function of the turbulence intensity, because the pressures depend significantly on the turbulence of
the approach flow. The positive peak external pressure coefficient on a roof is evaluated by using the
positive external pressure coefficient C pe for zone Ru in Table A 6.9(1). If no positive value of
C pe is provided for small roof angles, it is not necessary to evaluate the positive wind pressures.
Negative peak external pressure coefficients in the edge and corner regions are significantly influenced
by vortices related to flow separation at the edge. Negative peak pressure coefficients tend to increase
in magnitude as the turbulence intensity of the approach flow increases. However, the influence of
turbulence on negative peak pressure coefficients is smaller than that on positive peak pressure
coefficients on windward walls. Consequently, the provision of negative peak pressure coefficients is
determined from the values for terrain category IV and are independent of turbulence intensity. High
suctions are induced in the edge and corner regions of walls and roofs, whose widths are affected by
building dimensions such as height and width.
For gable roofs, very high suctions are induced near corners (zone Rb ) when the roof angle is
less than or equal to 10 and in the ridge corner (zones Rb and Rg ) when 20 . For
mono-sloped roofs, very high suctions are induced near the higher eaves corners (zone Rd ); the
suctions are larger and the high suction area is wider than that for gable roofs. Consequently, the peak
external pressure coefficient for zone Rd is larger than that for gable roofs. In such high suction
zones, the wind load can be reduced by using the area reduction factor k C when the subject area AC
of components/cladding is greater than 1 m2 (up to 5 m2).
2) Buildings with vaulted roofs
The peak external pressure coefficients C pe are determined from the results of wind-tunnel
experiments33), focusing on medium-scale buildings in urban areas, in which the h / B1 ratio is varied
from 0 to 0.7 and the f / B1 ratio from 0.1 to 0.4. When the f / B1 ratio is small, the corner and edge
regions of a roof are significantly affected by vortex generation as in the flat roof case. This results in
larger peak suctions in zones R a and Rd . When the f / B1 ratio is relatively large, large peak
suctions are induced in zone Rd for winds nearly perpendicular to the gable edge and in zone Rc
for winds nearly perpendicular to the eaves.
Taking these wind pressure features into account, the roof is divided into several zones and positive
and negative peak external pressure coefficients are provided for these zones, as shown in Table
A6.18(2). When the f / B1 ratio is lower than 0.1, the roof is subjected to higher suctions similar to
gable and mono-sloped roofs. Therefore, it is not necessary to evaluate the positive peak external
pressure coefficients. The values for walls can be determined from Table A6.18(1).
(3) Peak external pressure coefficients C pe for buildings with circular sections
CHAPTER 6 WIND LOADS C6-51

For buildings with circular sections, the maximum positive peak external pressure coefficient occurs
at the stagnation point on the windward face, whereas the maximum negative peak external pressure
coefficient occurs near the point of maximum negative mean external pressure. The vertical
distribution of positive peak pressure coefficients depends strongly on the mean velocity profile of the
approach flow in the same manner as that for buildings with rectangular sections. On the other hand,
negative peak external pressure coefficients are influenced by the aspect ratios H / D and surface
roughness of buildings. The factor k1 considers the effect of aspect ratio, and the factor k 2 the
effect of surface roughness in the transcritical Reynolds number regime. Negative peak external
pressure coefficients become larger in magnitude near the top of the building because of the flow
separation from the top (i.e. end effect). The factor k 3 considers this effect47). The values in Table
A6.19 are applicable to buildings with aspect ratios H / D less than or equal to 8, because the
provision is based on wind tunnel experiments using such models.
Only negative peak pressures are considered for roofs. The values of C pe for domes with
f / D = 0 provided in Table A6.20 can be used.
(4) Peak external pressure coefficients C pe for buildings with circular sections and spherical domes
Peak external pressure coefficients in Table A6.20 are determined from the results of wind tunnel
experiments34). External pressures on domes fluctuate significantly due to the effects of turbulence of
approach flow as well as of vortex generation. Therefore, both positive and negative peak pressure
coefficients are provided. Because the geometry of spherical domes is axisymmetric, they are divided
into three zones ( R a , Rb and Rc ) by coaxial circles. When the rise/span ratio ( f / D) is small,
negative peak external pressures become large in magnitude near the windward edge (zone R a ) due
to the flow separation at the windward edge. On the other hand, when the f / D ratio is large, large
positive peak external pressures are induced near the windward edge due to the direct influence of the
approach flow. Therefore, positive peak external pressure coefficients for zone R a are provided as a
function of the turbulence intensity I uH at the reference height H of the approach flow when
f / D 0.2 .

A6.2.6 Factor for effect of fluctuating internal pressures


Peak wind force coefficients for components/cladding shall be determined from the maximum
instantaneous values, both positive and negative, of the pressure difference between the exterior and
interior surfaces. However, there are few data on these pressure differences. In the Recommendations,
it is assumed that the peak wind force coefficient CC is represented by Eq.(A.6.15), because the peak
external pressure coefficients C pe are usually obtained from wind tunnel experiments and a large
amount of data is available.
Figure A6.2.3 shows a schematic illustration of fluctuating external and internal pressures. The
frequency of internal pressure fluctuations is lower than that of external pressure fluctuations, and the
*
peak external and internal pressures are not induced simultaneously. The factor C Pi for the effect of
fluctuating internal pressures in Eq.(A6.15) does not represent the peak internal pressure coefficient
C6-52 Recommendations for Loads on Buildings

itself but an equivalent value that provides the actual peak wind force when combined with the peak
external pressure coefficient C pe . The value of CC is evaluated from a series of computations for
the peak wind force coefficients using wind tunnel data on C pe for various building configurations.
The following assumptions are made in the computations:

1) Gaps and openings in the external walls are uniformly distributed, and the internal pressures are
generated from the external pressures at the locations of the gaps and openings.
2) The fluctuating internal and external pressures are independent of each other.

When the building has intentionally designed openings or when glass windows on the windward
face are broken by flying debris, the size of the openings may be very large compared with ordinary
gaps and openings. The values in Table A6.21 cannot be used for such cases. It is necessary to estimate
the peak wind force coefficients appropriately by using the data on the external and internal pressures
obtained from wind tunnel experiments. Some international codes and standards20), 50) provide internal
pressure coefficients for buildings with dominant openings.

Wind force coefficient, wind pressure coefficient


^
C C
^ C*pi wind force coefficient
Cpe
external pressure coefficient
internal pressure coefficient
^
C peak wind force coefficient
C
^
C peak external pressure coefficient
pe
0
Time peak internal pressure coefficient
Cpi

Cpi

Fig.A6.2.3 Example of fluctuating external and internal pressures acting on components/cladding

A6.2.7 Peak wind force coefficient for components/cladding


For free roofs, it is necessary to directly evaluate the net wind force represented by the pressure
difference between the top and bottom surfaces. Regulation of peak wind force coefficients is based on
previous wind tunnel experiments for the most critical peak wind forces irrespective of wind
direction38). When the roof angle is relatively large, large peak wind forces are induced along the roof
edges as well as along the ridge, because large suctions are induced by conical vortices on either the
top or bottom surface of the roof. The roof is divided into two zones ( Ra and Rb ), and positive and
negative peak wind force coefficients are provided for each zone as a function of roof angle . Larger
net wind forces are induced in zone Rb .
When any obstruction whose blockage ratio is larger than approximately 50% is placed under the
roof, it is necessary to evaluate the peak wind force coefficients from an appropriate wind tunnel
experiment and so on.
CHAPTER 6 WIND LOADS C6-53

A6.3 Gust Effect Factors

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


(1) Fundamental consideration
In this recommendation, gust effect factor is based on overturning moment as described by the
following equation.
M M + g D MD g
GD = Dmax = Dmax = 1 + D MD (A6.3.1)
MD MD MD
where M Dmax , M D , MD are maximum value, mean value and rms of overturning moment at the
base of the building, respectively. M Dmax and MD involve load effect due to the dynamic
response of the building. If MD is expressed as composition of background component MDQ and
resonance component MDR , Eq.(A6.3.1) becomes as follows.
2 2
MDQ f D S MD ( f D )
GD = 1 + g D MDQ + MDR M D 1 + g D 1 + D2 2
(A6.3.2)
MD 4 D MDQ
where S MD ( f D ) is power spectrum density of overturning moment at natural frequency for the first
mode f D and D is the mode correction factor. MDR is considered for only the first mode
vibration, and MDR is inertia force by vibration as described in the following equation.
H H
MDR = a ( Z )m( Z ) ZdZ = a ( H ) ( Z )m( Z ) ZdZ (A6.3.3)
0 0
where a ( Z ) , m(Z ) and (Z ) are rms of acceleration at height Z , mass per unit height and
vibration mode, respectively.
The parameters of Eq.(A6.3.2) are expressed by aerodynamic force coefficients as follows.
M D = q H BH 2 C MD (A6.3.4)
2
MDQ = qH BH C ' MD (A6.3.5)
f D S MD ( f D ) f D* S CMD ( f D* )
2
= (A6.3.6)
MDQ C '2MD
'
where C MD is overturning moment coefficient, C MD is rms overturning moment coefficient and
S CMD ( f D* ) is power spectrum of overturning moment coefficient at non-dimensional frequency f D* .
If these equations are taken into consideration, Eq.(A6.3.2) becomes as follows.
C ' MD f D* S CMD ( f D* )
GD 1 + g D 1 + D2 (A6.3.7)
C MD 4 D C ' 2MD
Additionally, in this formula non-dimensional frequency is defined by turbulence scale,
f D* = f D LH / U H , but, in the wind tunnel test breadth of the building it is used usually f D* = f D B / U H .
(2) Model of wind force
The model of wind force is based on the assumption that wind velocity fluctuation is directly
changed into the wind pressure on the wall of the building. In this model, mean wind velocity,
turbulence intensity, power spectrum of wind velocity and co-coherence are described by Eqs.(A6.8),
C6-54 Recommendations for Loads on Buildings

(A6.11), (A6.1.3), (A6.1.4), respectively. Additionally, wind force coefficient is expressed by a


difference of the wind pressure coefficient of the windward side and the wind pressure coefficient
(constant) of a lee side as described by the following equation.
2
Z
CD = CPA CPB (A6.3.8)
H
'
C MD , C MD and S CMD ( f D* ) are expressed using the parameter of the recommendation equations as
follows.
CMD = CH Cg (A6.3.9)
'
C MD = C H Cg' (A6.3.10)
f D* S CMD ( f D* ) = C ' 2MD FD (A6.3.11)
where CH is wind force coefficient at the top of the building, Cg is a factor relevant to overturning
moment in the along-wind direction, Cg' is a factor relevant to rms overturning moment in the
along-wind direction and FD is a spectrum factor of windward force. Spectrum factor of wind
velocity F , size reduction factor S D , factor expressing correlation of wind pressure of a windward
side and a leeward side R are considered for FD .
Characteristics of overturning moment expressed by Eqs.(A6.3.9)(A6.3.11) are shown in
Fig.A6.3.1 in comparison with those obtained from wind tunnel tests. The recommendation values of
overturning moment and rms overturning moment are slightly greater than the test values, and the
spectrum is mostly in agreement with the test values.

1.5 1.5 10 -2
recommendation value/test value

recommendation value/test value

1.0 1.0 10 -3
fSCMD(f)

category
category

0.5 II 0.5 10 -4 test

II
III III recommendation

IV IV
0.0 0.0 10 -5 -3
0 1 2 3 0 1 2 3 10 10 -2 10 -1 10 0
side ratio D/B side ratio D/B fB/UH
(a) mean overturning moment (b) rms overturning moment (c) power spectrum density of
coefficient coefficient over turning moment
Figure A6.3.1 Along-wind force in comparison with those obtained from wind tunnel tests
( H / BD = 4 )52)

(3) Fluctuating component of overturning moment


When the vibration mode is = Z / H , the relation between spectrum of overturning moment due
to the wind force S MD ( f ) and spectrum of overturning moment due to the load effect by vibration
CHAPTER 6 WIND LOADS C6-55

'
S MD ( f ) is expressed by the following equation.
2
S 'MD ( f ) = m ( f ) S MD ( f ) (A6.3.12)
2
where m ( f ) is mechanical admittance as expressed by the following equation.
2 1
m ( f ) = (A6.3.13)
{
1 ( f / fD ) 2 2
}
+ 4 D2 ( f / f D ) 2
2
The variance of overturning moment due to the load effect by vibration MD is the integral of
2
Eq.(A6.3.12), and the variance consists of back ground component MDQ and resonance component
2
MDR as expressed by the following equation.

2
MD = S ' MD ( f )df MDQ
2 2
+ MDR
0
f D S MD ( f D )
0 0
2 2
= S MD ( f )df + S MD ( f D ) m ( f ) df = MDQ + (A6.3.14)
4 D
In this equation, resonance component is estimated approximately as a response to white noise
S MD ( f D ) .
Therefore, overturning moment for maximum load effect is expressed by following equation.
2 2
M Dmax = M D + g D MDQ + MDR (A6.3.15)
where g D is called peak factor, and is the ratio of maximum fluctuating component to standard
deviation. This is expressed by the following equation, based on the theory of stationary stochastic
process.
0.577
g D = 2 ln( DT ) + 2 ln( DT ) + 1.2 (A6.3.16)
2 ln( DT )
where T is time for evaluation and D is level crossing rate calculated from power spectrum density
as in the following equation.

D =
0 f 2 S ' MD ( f ) df
fD
RD
(A6.3.17)
1 + RD
0 S 'MD ( f )df
Additionally, in some foreign wind loading standards, M Dmax is expressed by the following formula.
In this equation, the background component and the resonance component are distinguished.
M Dmax = M D + g Q2 MDQ
2
+ g R2 MDR
2
(A6.3.18)
where g Q is peak factor of background component (=3.4) and g R is peak factor of resonance
component calculated from Eq.(A6.3.16) as D = f D .
(4) Vertical distribution of equivalent static wind load
In the gust effect factor method, the vertical distribution of wind load is given by mean wind load
multiplied by gust effect factor. This wind load is an approximate value based on the assumption that
vibration mode is close to mean wind load distribution and the building has uniform density. Actually,
the mean, background and resonance components of wind load distribution are different. The mean
component is expressed by Eq.(A6.3.8), and the resonance component is expressed by Eq.(A6.3.3).
Therefore, if the vertical distribution of building mass is remarkably uneven, the resonance component
C6-56 Recommendations for Loads on Buildings

should be estimated carefully. In that case, the distribution of resonance component for the
fundamental vibration mode could be estimated from the following equation.
2 2
WD = W D + WDQ + WDR (A6.3.19)
where
W D = qH CD A
C'g
WDQ = g DQ q H C D A
Cg
A
WDR = a Dmax ( Z )m( Z )
B
where

WD , WDQ , WDR (N): mean, background and resonance component of wind load, respectively
aDmax (cm/s2): maximum acceleration at top of building as defined in A6.10.2
g DQ : peak factor of background component
In this recommendation, it is assumed that the background component has a similar distribution to
mean component. The following methods may also be used.
1) Shear force or overturning moment at a certain building height may be obtained from the integral of
pressure on area over the height20).
2) Load distribution can be defined by LRC formula53).
(5) Example of calculation of gust effect factor
Figure A6.3.2 shows the variation of gust effect factor by terrain category and building height for
H / B = 4 , D / B = 1 and U 0 = 35 m/s. The gust effect factors become large with terrain category and
building height.

3.8
3.6 category
gust effect factor G D

3.4 V
IV
3.2 III
3.0 II
I
2.8
2.6
2.4
2.2
2.0
0 50 100 150 200 250 300
0 50 100 150 200 250 300
height of building (m)
Figure A6.3.2 Variation of gust effect factor with terrain category and building height

A6.3.2 Gust effect factor for roof wind loads on structural frames
Gust effect factor for roof wind loads on structural frames is influenced by external pressure and
internal pressure. It can be assumed that there is no correlation between fluctuation of external
CHAPTER 6 WIND LOADS C6-57

pressure and fluctuation of internal pressure for a building without dominant openings. Furthermore,
Helmholtz resonance, the phenomenon of varying internal pressure at a specific frequency by external
pressure, can be disregarded. Fluctuating internal pressure coefficient is derived from the theory for
buildings with uniform openings54). Therefore, external pressure fluctuation, which is slower than
response time of internal pressure, is transmitted as internal pressure, and it is assumed that quicker
pressure fluctuation is not transmitted as internal pressure. Furthermore, fluctuating internal pressures
act on all parts of a roof simultaneously for more safety. Generally, response time of internal pressure
is long enough, compared with the natural period for the first mode of the roof structure. Therefore,
resonance of the roof structure for internal pressure can be disregarded. Under these conditions, gust
effect factor for roof wind loads is given by the following equation.
2 2 2 2 2
g Re rRe (1 + RRe ) + g Ri rRi rc
GR = 1 (A6.3.21)
1 rc
where g Re and g Ri are peak factors for generalized external pressure and generalized internal
pressure, and these value are g Re = 3.5 , g Ri = 3 from the results of test and measurement. rRe and
rRi are the generalized fluctuating external and internal pressures divided by the generalized mean
wind pressure coefficient. rc is the generalized mean internal pressure divided by the generalized
mean external pressure coefficient. RRe is resonance factor, which is calculated from the
non-dimensional power spectrum density at the frequency of the first mode of the roof and the critical
damping ratio.
wind load

time

Figure A6.3.3 Fluctuation of roof wind loads when wind force coefficient is small

An equation of gust effect factor is expressed for two cases of internal pressure coefficient,
C pi = 0.4 and C pi = 0 , given by Table A6.11. If wind force coefficient is small, roof wind loads act
in the upward direction and in the downward direction as shown in Fig.A6.3.3. When combinations
with other loads are considered, downward wind load can be dominant even if the absolute value is
small. Therefore, downward wind load can be calculated. In Eq.(A6.17), GR for +corresponds to
load in the same direction as given by wind load coefficient, and GR for is opposite. The above is
the same for Eq.(A6.18) and Eq.(A6.19). However, wind force coefficients are given as positive or
negative in A6.2.2, and gust effect factor should be calculated from Eq.(A6.17) with +. Furthermore,
C6-58 Recommendations for Loads on Buildings

the equation, f R 0.57 ( is deformation at center due to weight), can approximately evaluate
the natural frequency for the first mode of the roof beam, and the document55) is useful for estimating
the critical damping ratio, R .
(1) Case for C pi = 0.4
Roof wind loads can be calculated for roof beams parallel to the wind direction and for roof beams
normal to the wind direction.
If external pressure coefficient C pe is 0.4 over the whole subject area as center beam shown in
Fig.A6.3.4(a), the wind force coefficient becomes C R = 0 . In this case, roof wind loads can be
calculated from Eq.(A6.18), which is the product C R G R of wind force coefficient C R and gust
effect factor GR . However, when the wind force coefficient becomes partially C R = 0 as shown in
Fig.A6.3.4(b), the wind loads can be calculated from Eq.(A6.17).
-0.2 -0.4
-1.0 -0.6
external
pressure
coefficient


-0.4
-0.8
internal pressure
-0.4 external
pressure coefficient

coefficient

-0.4
wind force
-0.4 internal
pressure
0 coefficient
+0.2 -0.2
coefficient

beam -0.6
=

0 -0.4
-0.4 wind force

coefficient
beam

beam

beam

wind direction

wind
direction

(a) beams normal to the wind direction (b) beams parallel to the wind direction
Figure A6.3.4 Relation between wind force coefficient and external or internal pressure coefficient
(for C pi = 0.4 )

(2) For C pi = 0
Wind force coefficient is equal to external pressure coefficient for C pi = 0 . In this case, gust effect
factor can be calculated from Eq.(A6.19). The equation considers the mean and fluctuating
components of external pressure, and the fluctuating component of internal pressure.

A6.4 Across-wind Vibration and Resulting Wind Load

A6.4.1 Scope of applications


The procedure described in this section applies to the equivalent static wind load with consideration
of across-wind forced vibration at a design wind speed lower than the non-dimensional critical wind
speed for vortex-induced vibration or aeroelastic instability. For a design wind speed expressed by
CHAPTER 6 WIND LOADS C6-59

U H /( f L BD) > 10 , aeroelastic instability may well occur and wind load will need to be calculated
from the wind force and the response in wind tunnel tests.
Along-wind vibration is caused by turbulence in natural wind, but across-wind vibration is caused
by wind turbulence as well as by the vortex in the wake of the building. Although there are many study
examples with regard to the behavior of a vortex in the wake of a building, unclear points remain.
Furthermore, since the behavior is greatly affected by building shape, it is difficult on the whole to
theoretically estimate across-wind vibrations in the same manner as for along-wind vibrations. With
consideration of the first mode, an estimation equation for across-wind load has been derived from
data of across-wind fluctuating overturning moment obtained from wind tunnel tests. Subjects for this
estimation equation are structures with rectangular planes (side ratio D / B = 0.2 ~ 5 ) from which
many experimental data have been obtained. Moreover, by taking into account the fact that
experimental data for buildings with an aspect ratio H / BD exceeding 6 are insufficient, and that
aeroelastic instability easily occurs in these buildings, the scope of application is limited to aspect
ratios of 6 or less.
Furthermore, data of across-wind fluctuating overturning moment for buildings with plane shapes
other than rectangular planes can be obtained from wind tunnel tests. Where it is unnecessary to
consider aeroelastic instability, across-wind wind loads can be calculated using the method indicated
in the recommendations.

A6.4.2 Procedure
(1) Concept of wind load estimation
Since a fundamental mode usually predominates in across-wind vibration, across-wind loads are
calculated using the spectral modal method considering only to the first translational mode, in the
same manner as for along-wind loads. For the non-resonance component, the profile of fluctuating
across-wind force is set to be vertically uniform and the magnitude of the fluctuating wind force is
decided to agree with the fluctuating overturning moment. The resonance component estimates the
inertia force due to vibration and the vertical profile is determined using L in Eq.(A6.33) so as to be
proportioned to the first translational mode.
It is recommended that the critical damping ratio be estimated with reference to Damping in
buildings 7).
(2) Modeling of overturning moment
The overturning moment varies with building shape and wind characteristics, but in the subjective
scope the breadth-depth ratio has the greatest effect on the overturning moment: the effects of other
parameters are slight. Therefore, in the recommendations, the fluctuating overturning moment is set as
a function of only the breadth-depth ratio of a building based on wind tunnel test data 52, 56).
(3) Buildings with circular planes
Across-wind responses of buildings with plane shapes other than rectangular planes can be
estimated with the same concept. This section details buildings with circular planes. The parameter
C6-60 Recommendations for Loads on Buildings

values used in Eq.(A6.20) need to be set to C L' = 0.06 , m = 1 , 1 = 0.9 , f S1 = 0.15U H / B ,


1 = 0.2 . These parameter values are in the transcritical critical region of Reynolds number
( U H D 6 (m2/s)).

A6.5 Torsional Vibration and Resulting Wind Load

A6.5.1 Scope of application


The procedure described in this section applies to the equivalent static wind load with consideration
of torsional vibration with a design wind speed lower than the non-dimensional critical wind speed for
vortex-induced vibration or aeroelastic instability. For the design wind speed expressed by
U H /( f T BD ) > 10 , aeroelastic instability may well occur and the wind load needs to be calculated
from the wind force or the response in wind tunnel tests.
Torsional vibration is caused by asymmetric wind pressure distribution on the windward face, side
faces and leeward face. This is due to both wind turbulence and the vortex in the buildings wake. The
torsional moment induced wind force is subject to the effects of building shape and wind behavior.
Therefore, the method for assessing the torsional wind load is derived from the fluctuating torsional
moment data obtained from wind tunnel tests as for the across-wind direction. Subjects for this
estimation equation are buildings with rectangular planes (side ratio D / B = 0.2 ~ 5 ) and aspect ratio
H / BD of 6 or less, from which many experiment data have been obtained.
Furthermore, data of torsional moment for buildings with plane shapes other than rectangular planes
can be obtained by carrying out wind tunnel tests. Where aeroelastic instability does not need to be
considered, torsional wind loads can be calculated using the method indicated in the
recommendations.

A6.5.2 Estimation equation


(1) Concept of wind load estimation
Since the effects of pressure acting on both sides on the torsional moment are complex, it is difficult
to formulate the power spectral density as a simple algebraic function. However, it is relatively easy to
collect experimental data of the response angle acceleration. Therefore, the equation for computing the
torsional wind load is based on the estimate of the response angle acceleration. With regard to the
non-resonant component, the profile of fluctuating torsional moment is set as vertically uniform and
the magnitude of the fluctuating torsional moment is decided to agree with the fluctuating torsional
moment at the base of the building. The resonant component estimates the inertia force due to
vibration and the vertical profile is determined using L in Eq.(A6.33) so as to be proportioned to the
first translational mode. Buildings with an eccentric factor (eccentric distance / radius of rotation) of
0.2 or less for which any effect of eccentricity can be ignored are subject to the formulation of the
estimation equation. The wind load on a building for which the eccentricity cannot be ignored needs to
be calculated by carrying out wind tunnel tests.
CHAPTER 6 WIND LOADS C6-61

It is recommended that the critical damping ratio be estimated with reference to Damping in
buildings 7).
(2) Modeling of torsional moment
The torsional moment varies according to building shape and wind characteristics, but in respect of
buildings in the subjective scope the breadth-depth ratio exerts the greatest effect on the torsional
moment and the effects of other parameters are slight. Therefore, in the recommendations, the
fluctuating torsional moment is set as a function of only the breadth-depth ratio of a building based on
wind tunnel test data 52, 56).

A6.6 Horizontal Wind Loads on Lattice Structural Frames

A6.6.1 Scope of application


This procedure has been prepared for estimating horizontal wind loads on lattice structures built
directly on the ground, and whose members all have small enough sections in comparison with the
width of the structure for the flow field around a member to be dominated by the local wind speed.
The procedure for estimating wind loads on lattice structures is basically the same as that described for
horizontal wind loads on buildings in Section 6.2, and can be applied to lattice structures of varying
widths and solidity ratios in the vertical direction. In addition, the effects of accessory ladders are
considered by the evaluation of wind force coefficients of those obtained from wind tunnel tests and so
on.

A6.6.2 Procedure for estimating wind loads


Horizontal wind loads are estimated by a gust effect factor method. The wind loads are calculated
from the local design velocity pressure because lattice structures often have varying widths and
solidity ratios in the vertical direction.
The projected area in Eq.(A6.22) is the total projected area of all elements on one face normal to the
wind. The area per panel is usually calculated.

A6.6.3 Gust effect factor


In deriving Eq.(A6.23), it is assumed as follow:
i) Solidity ratios in the vertical direction are uniform, that is to say, wind force coefficients
of each panel are uniform.
ii) A fundamental mode shape can be given by Eq.(A6.6.1) where = 2 , and vibration
modes higher than the fundamental one are neglected.

Z
= (A6.6.1)
H
According to the above assumptions, the peak response x max,Z at height Z is given as a function
of the generalized stiffness K of the fundamental mode by:
C6-62 Recommendations for Loads on Buildings

2I H
BD (1 + RD )
q H C D HB0
x max,Z = g D (A6.6.2)
K 0.95 + +
However, the mean response X Z at height Z is given by:
qH CD H B0 B BH
XZ = 0 (A6.6.3)
K 1 + 2 + 2 + 2 +
where qH , I H are the velocity pressure and the turbulence intensity, respectively, at H height,
and is the exponent of the power law in the wind speed profile. g D , RD and BD are the peak
factor, the resonance factor and the back ground excitation factor, respectively.
Gust effect factor is given by Eq.(A.6.23).

Figure A6.6.1 Definition of B0 , BH , H

A6.7 Vortex Induced Vibration

A6.7.1 Scope of application


This section describes vortex-induced vibration, which can occur in tall slender buildings, chimneys,
and structural components with circular sections.

A6.7.2 Vortex induced vibration and resulting wind load on buildings with circular sections
Shear layers separated from windward corners of both sides of buildings roll up alternately to shed
into wake and form Karman vortex streets behind the buildings. According to the alternate shedding,
the periodic fluctuating wind loads act on the buildings in the across-wind direction. When the natural
frequency of the building coincides with the vortex shedding frequency, the vibration of the building
can be resonant with the periodic fluctuating wind loads, causing the building to vibrate at large
amplitude in the across-wind direction. This is vortex-induced vibration, which is a problem for many
structures, particularly chimneys.
The critical wind speed of the resonance is larger than the design wind speed for most buildings, so
these phenomena are not normally important. However, as the critical wind speed is smaller than
CHAPTER 6 WIND LOADS C6-63

design wind speed for very slender buildings with small natural frequency and damping like steel
chimneys, tall buildings and building components, the effect of vortex induced vibration should be
checked carefully in the wind resistance design stage.
A lot of research has been done on vortex-induced vibration and a number of methods have been
developed in the past decade for estimating vibration amplitude and its equivalent static wind loads,
particularly for structures with circular sections. The equivalent wind loads described in the
recommendation are based on the spectral modal method in which the Strouhal number of vortex
shedding is 0.2, and the power spectrum of the fluctuating wind loads depends on the vibration
amplitude6) and the Reynolds number.
The effects of structural density, damping and Reynolds number are included in the resonant wind
force coefficient C r , which is shown in Table A6.2.3 for three categories of Reynolds number region
and for two types of structures with various density and damping. The rows in the table show the
effect of Reynolds number, that is, U r Dm < 3 is the subcritical region, 3 U r Dm < 6 is critical
region and 6 U r Dm is super/trance critical Reynolds number region. s L in Table A6.23
depends on the amplitude at the resonant condition. s L < 0.5 corresponds with the large
amplitude, and s L 0.5 corresponds with the small amplitude.

A6.7.3 Vortex induced vibration and resulting wind load on building components with circular
sections
Occurrence of vortex induced vibration of building components with circular section can be
checked by Eq.(A6.26). Most design wind speeds for components like members of truss towers are
larger than the critical wind speed, so the effect of vortex induced vibration should be checked
carefully. In particular, the vibration amplitude can be very large for components like steel pipes
whose mass and damping are small. The equivalent wind loads described in Eq.(A6.27) are introduced
in the sub-critical Reynolds number region based on wind tunnel tests59). The equation is applicable for
various boundary conditions at the ends of components.

A6.8 Combination of Wind Loads

A6.8.1 Scope of applications


This section defines the combination of horizontal wind loads and roof wind loads on structural
frames. These wind loads are evaluated separately, but this does not mean that each wind load acts on
the building independently. However, maximum wind loads do not occur at the same time. Therefore,
if they are applied to the building at the same time, the combination of wind loads overestimates actual
loads. This section shows the formula for combination of wind loads considering correlations of wind
force and response. The formula is divided in two ways: for buildings not satisfying the conditions of
Eq.(6.1) and for buildings satisfying the conditions of Eq.(6.1). Combination of horizontal wind loads
and roof wind loads is also described.
C6-64 Recommendations for Loads on Buildings

A6.8.2 Combination of horizontal wind loads for buildings not satisfying the conditions of Eq.(6.1)
Buildings not satisfying the conditions of Eq.(6.1) have a small resonance component. For such
cases, it is considered that wind load of times of the windward loads act in the across-wind
direction, as shown in figure 6.8.1. tends to increase with building height according to the stress
analysis for buildings with rectangular columns using wind load from wind tunnel tests. Therefore, an
approximate equation of for an 80m-high building is defined as per the recommendation.

wind quasi-static wind load


considered resonance
plan of building
Figure A6.8.1 Windward load and combined Figure A6.8.2 Relation between side ratio
load for across wind direction (D/B) and combination factor

A6.8.3 Combination of horizontal wind loads for buildings satisfying the conditions of Eq.(6.1)
Buildings satisfying the conditions of Eq.(6.1) have a large resonance component. For such cases, it
is assumed that response probability is expressed by a normal distribution. If the overturning moments
in two directions, M x , M y , are expressed by a 2-dimensional normal distribution, the equivalence
line of probability becomes an eliptical line using correlation coefficient of response, , as shown in
Figure A6.8.3. Every point on the eliptical line (solid line) can be considered as a load combination,
but it is not practical to consider a lot of them. Therefore, load combinations can be defined as the
apexes of an octagon enveloping the oval. In other words, y-direction overturning moment M yc ,
which should be combined with maximum x-direction overturning moment M xmax , is defined by the
following equation using mean y-direction overturning moment M y and maximum fluctuating
component of y-direction overturning moment m ymax .

M yc = M y + mymax ( 2 + 2 1 ) (A6.8.1)

Table A6.24 shows the combination of loads according to the upper equation considering following
characteristics of along-wind, across-wind and torsional wind loads.
Co-coherence (correlation coefficient for each frequency) is negligible between along-wind force
and across-wind force, and between along-wind force and torsional wind force. Therefore, = 0
as co-coherence of response is negligible.
Because the co-coherence between across-wind force and torsional wind force is not zero, the
absolute value of the correlation coefficient of response LT , shown in Table A6.25, is defined
by calculation based on wind tunnel tests.
LT is calculated by a statistical analysis method61) under the conditions that the critical damping
CHAPTER 6 WIND LOADS C6-65

ratios for across-wind vibration and torsional vibration are 0.02, and the building has no coupling
vibration mode. Therefore, if the critical damping ratio differs greatly from 0.02 or the buildings
vibration mode is significantly coupled, it is necessary to carry out special research.
My
mx max
point A
M y max

my max M yc
considered point of M y0 m y max ( 2 + 2 1)
combination load my max
My
m y max (1 2 2 )

Mx
Mx M x max
Figure A6.8.3 Schema of load combination in consideration of response correlation

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


Combination of horizontal wind loads and roof wind loads can be considered theoretically as in
A6.8.2 or A6.8.3. However, because the relation between horizontal wind loads and roof wind loads is
not well enough understood, it is defined that horizontal wind loads and roof wind loads act at the
same time.

A6.9 Mode Shape Correction Factor

A6.9.1 Scope of application


The mode shape correction factor can be used in calculating the gust effect factor, the across-wind
load and the torsional wind load for a conventional building, as described in A.6.3.1, A6.4.2 and
A6.5.2, respectively, if the first translation mode shape function is different from = Z H and the
vertical distribution of mass per height of a building over the ground is not regarded as almost constant.
The mode shape correction factor can be used in calculating the gust effect factor for a lattice structure,
as described in A.6.3.3, if the first mode shape function is different from = (Z / H )2 and the mass
per height of a lattice structure is not regarded as almost constant.
The mode shape correction factor can be applied with ranging from 0.2 to 4 for a conventional
building, and with ranging from 1 to 3.5 for a lattice structure when the mode shape function can
be approximated by the function = (Z / H ) .

A6.9.2 Procedure
The mode shape correction factor is specified by Eq.(A6.32). This corrects the gust effect factor for
C6-66 Recommendations for Loads on Buildings

an along-wind load on a building according to its vibration mode.


The vibration mode shape correction factors for the resonance components of across-wind load and
torsional wind load are specified by Eqs.(A6.33) and (A.6.34), respectively.
given by Eq.(A6.35) is the ratio of the resonance component of the generalized wind force
for its first vibration mode to 1 for the reference vibration mode shape (the power index of a first
vibration mode = 1 for a conventional building and = 2 for a lattice structure).

= (A6.9.1)
1
The values of for a conventional building in Eq.(A6.35) are approximations that fit the results
obtained from a wind tunnel test for rectangular cross section buildings, in which the power index
indicating the vibration mode shape between 0.2 and 4 are taken into consideration.
The mode shape correction factor can be derived by multiplying the correct factor of the
generalized wind force by the correct factor of the generalized mass or the generalized inertial moment
of the building.
The vertical distributions of the along-wind load are taken into consideration by the vertical
distribution of the mean wind load, but the resonance component of the across-wind load or the
torsional wind load is proportional to the vibration mode, because the mean load is not considered in
the recommendation. As a result, the mode shape correction factor of the across-wind load or the
torsional wind load involves a variable for height.
The mode shape correction factor for a lattice structure is derived from the buffeting theory. This is
to deal with the lattices of varying widths in the vertical direction.
The mode shape correction factor can be set to 1 if the vibration mode shape agrees with the
reference vibration mode shape and the vertical distribution of mass per unit height of a building over
the ground is regarded as almost constant. If the vibration mode shape agrees with the reference
vibration mode shape and the vertical distribution of mass per unit height can not be regarded as
constant, the mode shape correction factor can be replaced by the ratio of the generalized mass or the
generalized inertia moment of a building to that with a uniform mass distribution in the vertical
direction. Furthermore, if the vertical distribution of mass per unit height of a building over the ground
is regarded as almost constant, the mode shape correction factors for the along-wind load, the
across-wind load and the torsional wind load can be simplified by Eqs.(A6.9.2), (A6.9.3) and (A6.9.4),
respectively.

1.1 0.1 conventional building



D = (A6.9.2)
(0.16 + 0.4 ) 0.5 B 0.3( 2 ) + 1.4
BH
lattice structure
0
1
Z
L = (0.27 + 0.73) (A6.9.3)
H
CHAPTER 6 WIND LOADS C6-67

1
Z
T = (0.27 + 0.73)
(A6.9.4)
H
In addition, the generalized mass M D , M L and the generalized inertial moment I T of a
building can be calculated according to Eqs.(A6.9.5) and (A6.9.5), respectively.
2
H Z
M D (L ) = 0 mZ
H
dZ (A6.9.5)

2
HZ
IT =
0I Z dZ
H
(A6.9.6)

where mZ and I Z are the mass and the inertial moment at height Z , respectively.

A6.10 Response Acceleration

A6.10.1 Scope of application


This section defines the maximum along-wind response acceleration for ordinary buildings, the
maximum across-wind response acceleration for buildings with rectangular plan satisfying the
conditions of A6.4.1 and the maximum torsional response acceleration for buildings with rectangular
plan satisfying the conditions of A6.5.1.
Each formula considers only the first vibration mode. If a building has a large dynamic response in
higher modes or partial vibration, other special research should be carried out.

A6.10.2 Maximum along-wind response acceleration


Rms of generalized response acceleration aD is given by the following equation.
2
m( f )
0 S g ( f )(2f )
2 4
aD = df (A6.10.1)
K g2
where aD is rms of generalized acceleration, S g ( f ) is power spectrum density of generalized
wind force, m ( f ) 2 is mechanical admittance as described in Eq.(6.3.13), f is frequency and K g
is generalized stiffness as described in the following equation.
K g = M D (2f D ) 2 (A6.10.2)
where M D is generalized mass. Because the resonant component is dominant in acceleration,
S g ( f ) can be substituted by white noise having power spectrum density at natural frequency f D , as
described in the following equation.
FD
S g ( f D ) = ( qH BHCH C 'g ) 2 (A6.10.3)
fD
where FD is along-wind force spectrum factor, as shown in A6.3.1.
If Eqs.(A6.10.2) and (A6.10.3) are incorporated in Eq.(A6.10.1), the equation become the
following.
RD
aD = qH BHCH C 'g (A6.10.4)
MD
C6-68 Recommendations for Loads on Buildings

Furthermore, aD is multiplied by the peak factor in the recommended equation for the
acceleration at the top of the building. Because the resonant component is dominant in acceleration,
level crossing rate D for calculating peak factor is approximated by the natural frequency f D .

A6.10.3 Maximum across-wind response acceleration


The equation consists of coefficients according to across-wind direction as a development in the
along-wind direction, A6.10.2.

A6.10.4 Maximum torsional response acceleration


admax = aTmax d (A6.10.5)

A6.11 Simplified Procedure

A6.11.1 Scope of application


A simplified procedure is used for estimating wind load for small buildings. This procedure can be
applied to buildings that have regular shapes and structural systems, such as detached houses. The
reference height and the projected breadth shall be less than 15m and 30m, respectively.

A.6.11.2 Procedure
The simplified procedures are derived from the results of calculation for buildings with reference
heights of 5 - 15m and projected breadths of 5 - 30m, assuming that the wind directionality factor K D
is 1.0 and the terrain category is III. Therefore, this procedure can be applied to terrain categories IV
and V with some overestimates in wind loads. For terrain categories less than III, the exposure factor
Ce is introduced. When wind speed is expected to increase due to local topography, the wind loads
shall be increased appropriately, for example, by multiplying by the square of the topography factor
Eg .

A6.12 Effects of Neighboring Tall Buildings

When groups of two or more tall buildings are constructed in proximity, the fluid flow through the
group may be significantly deformed and have a much more complex nature than is usually
acknowledged, resulting in enhanced dynamic pressures and motions especially on neighboring
downstream structures. Therefore, study of mutual interference among closely-located tall buildings is
an important problems not only in wind resistant structure design but even in minimizing wind-motion
discomfort to building occupants. Wake-induced oscillation in the downstream structure is considered
to be affected by interference from upstream buildings of various sizes placed in various locations and
also by the turbulence of incident flows.
Figure A6.12.1 shows contours of the increase or decrease ratios for the maximum along/across
CHAPTER 6 WIND LOADS C6-69

wind responses of the downstream building exposed to interference from an upstream building at
various locations to those of an isolated building where the maximum responses including mean
deflection are estimated at near the design wind speeds of 4060m/s by a modal-spectrum method
(1,2). The contours are illustrated for an identical pair of square tall buildings with aspect ratio
H / BD = 4 where two coordinate axes are normalized by the non-dimensional distance using the
reference building breadth BD .
The response ratios in the across-wind direction are usually larger than those in the along-wind
direction. Interfering positions producing response ratio contours higher than 1.2 are generally
restricted to regions of 12 BD in the x-direction and 6 BD in the y-direction, whereas interfering
positions higher than 1.1 exceed the regions indicated in the figure.
When the flat terrain subcategories increase from Category II to Category IV, the dynamic responses
of the downstream building are relatively independent of mutual interference effect. This is closely
related to the fact that when turbulence is added to an incident flow, shedding vortices from an
upstream building and the alternately deformed wake surrounding the vortices are not clearly formed
in the wake owing to increased entrainment and diffusive action, and the production of additional
turbulence by the introduction of the upstream building is unlikely because of the sufficiently high
turbulence in the incident flow (3).

y y
6 BD 6 BD
1.1 1.2
4 4
1.1 1.1 1.2
1.3

1.2 1.3
2 2
1.1 1.0
x 1.0 1.0 x 1.2 1.1
12 BD 6 4 2 12 BD 6 4 2 1.3
0.8
(a) Terrain category II, along-wind direction (b) Terrain category II, across-wind direction
1.2
y
6 BD

1.2 1.1
1.2 4
1.2 1.4
1.2
1.1
1.2 2
1.0
1.0
1.0
x 1.1 1.0 1.1
12 BD 6 4 2 0.8

(c) Terrain category IV, across-wind direction


Figure A6.12.1 Contours of response ratios
C6-70 Recommendations for Loads on Buildings

A6.13 1-Year-Recurrence Wind Speed

1-year-recurrence wind speed U 1H is used to calculate the acceleration of wind response for the
evaluation of the habitability, defined in Eq.(A6.41).
Figure A6.5 is smoothing of the wind speed map based on the 1-year-recurrence wind speed at the
metrological offices, from which the wind speed U1 at any locations can be estimated. The
1-year-recurrence wind speeds at the metrological offices are established based on the daily-maximum
wind speed data regardless of wind directions collected from 1991 to 2000. On the other hand, because
the wind response characteristic is not the same for the wind direction, the wind speed, which becomes
the same acceleration is also different for the wind direction. Therefore, if the wind direction
characteristic, that is, the frequency of exceedance of each wind speed can be understood, a reasonable
design becomes possible. This wind direction characteristic in the range of the wind speed to evaluate
the habitability is generally clarified.
When the maximum acceleration a max is approximated as a function of wind speed U shown in
Eq.(A6.13.1), the return period t a max for maximum acceleration a max is calculated by
Eq.(A6.13.2). The probability at the right side of Eq.(A6.13.2) is expressed as the total sum of the
occurrence probability of the wind speed in every 16 azimuths shown in Eq.(A6.13.3).
amax = f (U ) (A6.13.1)
1
t a max = (A6.13.2)
1 Fa ( amax )

Fa ( amax ) = pi FU { }
16
f i 1 (amax ) (A6.13.3)
i =1
where
Fa ( a max ) : probability that maximum acceleration does not exceed a max
pi : occurrence frequency for wind direction i
{ }
FU f i 1 (amax ) : probability that the wind speed does not exceed the wind speed that the
maximum acceleration is equal to a max for wind direction i
The occurrence frequency at each wind direction pi , parameters ai and bi in Eq.(A6.13.4),
which are the parameters to calculate the right side of Eq.(A6.13.3), are shown in Table A6.13.1.
These parameters are estimated based on the daily maximum wind speed at 30 cities, with the least
square method applied for the data at Naha where typhoon is dominant, and the Gumbels moment
method for other cities. These parameters ai and bi should be used for the return period less than 1
year.
FU ( U i ) = exp[ exp{ ai (U i bi )}] (A6.13.4)
where
U i : 10-minute mean wind speed at 10m above ground over a flat and open terrain for wind
direction i
ai , bi : parameters estimated based on the daily maximum speed for wind direction i
CHAPTER 6 WIND LOADS C6-71

In addition, the wind direction factor in A6.1.4 should be used for 100-year-recurrence wind
speed, and it is not possible to use it here.

Table A6.13.1 parameters ai , bi and occurrence frequency pi for each wind direction at 30 cities
Asahikawa Sapporo Aomori Akita Sendai
ai bi pi(%) ai bi pi(%) ai bi pi(%) ai bi pi(%) ai bi pi(%)
NNE 0.58 4.26 3.3 1.58 3.59 0.4 1.22 3.57 4.6 0.73 4.79 0.4 0.82 4.40 1.7
NE 0.52 4.32 0.8 1.23 3.76 0.5 0.82 4.03 4.0 0.56 5.88 0.1 0.61 3.72 1.1
ENE 0.54 3.63 0.2 1.30 3.80 1.6 0.76 5.84 3.7 0.73 3.17 0.2 0.56 5.86 0.8
E 1.45 2.28 0.7 0.94 4.93 4.0 0.90 5.46 7.9 0.63 4.77 0.3 0.66 5.09 0.7
ESE 1.05 2.74 0.7 0.72 5.52 6.2 0.60 6.09 1.1 0.65 6.57 7.2 0.74 4.97 0.8
SE 0.76 4.44 6.4 0.59 7.49 8.1 0.64 7.71 0.7 0.63 6.25 17.0 1.14 3.89 16.7
SSE 0.55 5.61 17.1 0.47 8.86 13.5 0.58 4.61 0.5 1.25 4.70 0.2 0.76 4.58 13.3
S 0.54 4.34 3.9 0.43 7.94 3.2 1.58 2.79 0.2 0.72 6.27 0.1 0.75 4.66 6.4
SSW 0.48 6.37 3.3 0.45 7.31 1.8 0.54 5.27 2.8 0.38 8.91 2.2 0.87 5.06 1.8
SW 0.59 6.72 1.2 0.44 7.96 2.1 0.47 6.47 12.6 0.46 7.44 9.9 0.77 5.55 0.9
WSW 0.49 7.58 10.1 0.47 8.41 3.9 0.48 7.56 10.2 0.38 6.98 12.5 0.46 8.02 1.6
W 0.63 6.45 17.6 0.53 8.53 5.1 0.50 9.11 14.0 0.36 7.91 17.7 0.42 8.75 7.7
WNW 0.65 5.80 19.4 0.45 9.28 5.2 0.55 8.43 15.7 0.37 9.68 11.9 0.39 9.42 16.9
NW 0.68 4.78 8.8 0.46 8.63 19.5 0.66 6.00 8.6 0.45 9.56 9.2 0.43 8.62 10.1
NNW 0.86 5.66 4.7 0.59 7.05 23.3 1.00 4.20 6.7 0.60 7.87 3.8 0.56 5.44 7.9
N 0.83 4.92 1.8 0.83 4.78 1.6 0.84 3.49 6.7 0.75 5.86 7.3 0.75 5.17 11.6
Niigata Kanazawa Utsunomiya Maebashi Tokyo
ai bi pi(%) ai bi pi(%) ai bi pi(%) ai bi pi(%) ai bi pi(%)
NNE 1.07 4.61 14.4 0.76 5.11 5.8 0.70 4.47 18.5 0.28 9.56 0.1 0.87 5.46 4.2
NE 1.78 3.66 6.6 0.81 5.38 3.0 0.88 4.43 8.8 0.0 1.04 5.43 6.2
ENE 0.85 3.72 0.3 0.99 5.02 10.5 1.01 4.52 2.3 1.20 4.78 0.1 1.05 5.58 6.6
E 1.24 4.17 0.2 0.93 4.62 9.9 1.13 3.81 3.5 0.71 4.32 1.4 1.10 5.47 3.3
ESE 0.64 7.39 0.2 0.87 3.56 0.9 1.37 3.85 7.1 0.81 5.19 22.7 1.22 5.67 3.6
SE 0.69 8.05 6.6 1.35 3.32 1.2 1.33 3.76 9.2 0.99 4.83 8.7 1.37 6.02 1.9
SSE 0.98 5.69 4.4 2.15 3.26 1.6 1.00 4.26 9.1 1.21 3.84 2.0 0.94 4.99 0.3
S 1.65 4.38 1.7 0.17 4.11 0.5 0.78 4.68 6.0 1.17 3.76 1.2 0.83 6.53 20.2
SSW 1.19 4.78 3.0 0.42 7.62 8.2 0.75 4.48 6.4 0.57 2.88 0.3 0.56 7.61 2.0
SW 0.45 6.84 3.2 0.43 9.16 8.5 0.81 4.12 2.8 0.49 3.30 0.3 0.53 7.80 9.8
WSW 0.40 8.65 14.4 0.45 8.65 9.7 0.68 4.79 1.5 0.50 4.65 1.3 0.64 5.84 0.4
W 0.44 7.29 18.8 0.36 7.49 12.1 0.59 6.99 2.2 0.44 6.78 1.5 0.45 7.83 0.2
WNW 0.38 8.39 7.7 0.37 6.62 9.1 0.53 7.17 3.1 0.34 6.86 3.5 0.50 7.23 0.2
NW 0.48 8.40 6.9 0.44 5.36 7.4 0.47 5.14 1.6 0.45 6.82 26.0 0.45 8.28 4.9
NNW 0.52 7.43 6.7 0.38 5.82 3.7 0.50 5.67 3.6 0.47 8.43 26.7 0.47 7.37 25.8
N 0.66 5.56 4.9 0.66 4.89 7.9 0.58 4.49 14.3 0.51 11.0 4.2 0.64 5.85 10.4
Chiba Yokohama Shizuoka Hamamatsu Nagoya
ai bi pi(%) ai bi pi(%) ai bi pi(%) ai bi pi(%) ai bi pi(%)
NNE 0.73 6.38 6.2 0.58 7.52 1.8 0.83 4.75 2.9 0.80 4.59 0.4 1.10 3.61 1.6
NE 0.89 6.10 6.1 0.81 6.50 0.1 0.77 5.74 9.5 1.25 3.20 3.0 1.80 2.92 0.7
ENE 0.97 5.46 6.7 0.64 7.58 1.4 0.94 5.74 23.7 0.56 6.29 7.1 2.25 2.29 0.8
E 1.02 4.85 2.5 1.04 5.62 9.1 0.88 5.51 1.8 0.59 6.69 7.8 2.46 3.82 0.1
ESE 1.41 4.08 6.9 1.19 5.00 1.9 0.80 4.50 1.8 0.74 6.59 2.5 0.47 6.21 0.4
SE 1.27 4.26 9.9 0.71 5.81 0.9 0.75 3.95 1.3 1.03 5.35 5.5 0.47 6.03 3.8
SSE 0.68 4.96 3.8 0.76 5.48 7.5 0.92 4.96 6.1 0.80 5.06 6.7 0.51 6.54 13.5
S 0.77 4.62 2.2 0.63 5.85 4.2 0.81 5.46 17.6 0.93 4.93 2.5 1.16 4.85 11.0
SSW 0.35 9.90 4.8 0.40 9.60 5.8 0.50 6.50 4.1 0.67 5.94 0.6 1.26 4.67 2.0
SW 0.45 7.69 13.4 0.38 8.74 15.8 0.42 8.63 12.4 0.58 5.77 3.2 1.14 4.78 1.1
WSW 0.74 4.95 8.8 0.40 8.42 3.4 0.57 9.01 4.0 0.68 6.00 13.8 0.88 3.81 1.3
W 0.59 4.84 0.7 0.48 8.68 0.2 0.51 10.1 5.3 0.66 7.30 16.7 0.59 5.61 1.4
WNW 0.42 8.33 0.6 0.65 8.27 0.3 0.39 7.05 3.8 0.49 9.15 23.6 0.70 6.49 18.8
NW 0.46 7.68 7.0 0.36 8.39 0.5 0.55 4.96 1.8 0.39 8.27 5.9 0.53 7.29 19.0
NNW 0.48 6.26 15.1 0.32 6.95 2.5 1.28 4.07 2.9 0.55 4.18 0.5 0.50 5.49 16.2
N 0.61 5.46 5.3 0.46 7.21 44.5 0.59 3.81 1.0 1.69 4.88 0.2 0.89 4.10 8.3
C6-72 Recommendations for Loads on Buildings

Table A6.13.1(continued) parameters ai , bi and occurrence frequency pi for each wind


direction at 30 cities
Kyoto Osaka Kobe Wakayama Okayama
ai bi pi(%) ai bi pi(%) ai bi pi(%) ai bi pi(%) ai bi pi(%)
NNE 0.75 4.05 5.6 0.88 4.24 17.6 0.57 6.22 1.4 0.82 5.92 8.9 0.47 5.61 4.2
NE 0.73 4.01 5.1 0.72 5.23 4.4 0.32 5.41 1.2 1.28 4.44 5.3 0.89 3.61 2.3
ENE 0.93 4.97 6.1 0.56 5.95 4.4 0.48 6.50 12.1 1.90 3.99 10.3 1.13 4.17 7.6
E 0.85 5.15 6.6 0.63 5.80 1.7 0.59 5.67 3.3 1.63 4.19 1.4 0.67 5.30 5.2
ESE 0.78 4.80 3.9 0.89 4.80 1.3 1.01 4.06 0.5 0.29 4.96 0.3 0.56 5.13 4.3
SE 0.84 4.64 1.5 0.57 5.16 0.5 1.71 3.33 0.7 0.64 7.04 0.5 0.73 4.27 7.9
SSE 0.95 4.06 3.3 0.68 4.96 0.5 0.95 3.56 0.3 0.39 8.26 1.0 1.26 3.72 1.8
S 0.92 4.99 8.9 0.51 6.42 0.5 1.56 3.84 1.6 0.35 7.97 5.3 1.35 3.43 5.3
SSW 0.80 5.76 10.3 0.31 7.86 1.8 0.60 5.27 6.9 0.43 7.32 9.0 1.01 4.37 5.7
SW 0.58 6.03 2.5 0.64 5.75 14.4 1.05 5.54 5.9 0.94 4.71 10.7 0.60 5.24 14.1
WSW 0.72 6.56 4.6 0.80 5.57 19.0 0.65 5.89 13.7 0.61 4.44 9.0 0.62 5.34 6.0
W 0.69 6.83 3.8 0.46 6.47 12.7 0.50 7.15 11.8 0.38 6.58 4.0 0.45 7.01 5.9
WNW 0.61 6.74 4.7 0.39 6.80 1.9 0.51 7.17 7.2 0.54 8.19 5.3 0.40 8.07 5.8
NW 0.58 7.53 3.1 0.63 6.32 3.6 0.72 5.91 3.9 0.61 6.99 7.7 0.58 6.09 6.5
NNW 0.85 6.55 10.0 0.77 6.66 7.4 0.56 6.72 13.2 0.77 5.48 9.3 0.62 5.24 8.6
N 0.75 5.49 20.0 0.77 5.73 8.3 0.72 5.78 16.3 0.80 5.61 12.0 0.51 5.60 8.8
Matsue Hiroshima Takamatsu Kochi Matsuyama
ai bi pi(%) ai bi pi(%) ai bi pi(%) ai bi pi(%) ai bi pi(%)
NNE 0.72 5.15 1.6 1.26 4.40 31.5 1.29 3.65 2.5 0.82 7.11 6.7 0.91 6.61 3.4
NE 0.64 5.77 8.2 0.58 5.18 0.9 0.79 3.39 1.2 1.04 6.95 1.0 0.74 6.15 2.6
ENE 0.59 6.07 7.1 0.42 7.43 0.3 1.17 3.90 12.7 0.96 4.78 2.4 0.96 4.74 3.6
E 0.78 5.25 10.0 1.48 5.93 0.4 1.10 3.86 9.8 0.75 3.85 4.2 1.57 3.26 4.5
ESE 1.35 3.90 8.0 0.68 5.19 0.5 0.70 4.24 6.2 0.57 4.99 6.9 0.91 3.50 4.7
SE 0.61 5.22 1.7 1.39 5.13 0.2 0.87 3.71 2.2 1.38 4.02 21.7 0.87 4.45 2.7
SSE 0.45 5.62 0.5 0.23 5.84 0.6 0.93 3.88 0.2 1.93 4.19 6.0 0.47 6.18 2.6
S 0.68 5.75 0.2 0.49 4.66 5.7 0.77 2.89 0.4 1.11 4.50 8.5 0.42 6.40 2.0
SSW 0.40 8.19 1.8 0.84 5.28 8.2 0.69 5.39 0.6 0.71 5.18 2.4 0.89 4.59 2.1
SW 0.55 6.61 1.5 1.05 4.60 10.8 0.61 4.48 3.2 0.80 6.32 1.2 0.79 5.48 2.2
WSW 0.45 6.91 12.1 0.67 5.11 0.5 0.57 6.10 13.0 0.78 4.55 4.0 0.91 5.31 6.8
W 0.41 7.72 28.3 0.69 7.13 3.3 0.57 6.90 19.4 0.55 4.62 19.9 0.66 5.92 20.7
WNW 0.42 7.10 5.9 0.64 7.44 3.2 0.49 7.19 5.6 0.55 4.30 7.2 0.66 5.99 20.1
NW 0.59 5.91 8.6 0.70 6.54 0.9 0.50 6.53 2.9 0.59 7.03 1.7 0.99 5.09 13.0
NNW 0.67 5.93 3.4 0.79 6.65 3.7 0.87 4.32 11.4 0.56 7.08 1.4 0.92 5.04 6.6
N 0.64 6.08 1.1 0.99 4.54 29.3 0.98 4.41 8.7 0.66 7.22 4.8 0.69 6.99 2.4
Fukuoka Oita Kumamoto Kagoshima Naha
ai bi pi(%) ai bi pi(%) ai bi pi(%) ai bi pi(%) ai bi pi(%)
NNE 0.67 5.37 3.2 0.78 4.67 8.7 1.52 3.06 1.5 0.56 5.50 9.7 0.54 4.98 11.2
NE 0.95 5.61 1.6 0.95 4.27 7.3 0.56 5.42 2.1 0.52 6.03 7.0 0.63 5.75 2.6
ENE 0.87 5.82 0.5 1.54 3.91 11.3 0.60 6.28 3.7 0.36 5.95 0.9 0.79 6.17 7.2
E 1.18 5.42 0.2 0.80 3.71 3.3 0.50 6.42 2.9 0.42 6.55 0.9 0.30 -0.27 8.9
ESE 1.03 5.01 0.7 0.47 5.92 0.6 0.59 5.74 2.5 0.31 5.67 1.7 0.43 3.63 8.6
SE 0.77 5.41 9.7 0.37 5.99 2.1 0.34 3.88 1.2 0.46 5.17 5.0 0.20 -4.53 6.7
SSE 0.40 7.33 4.5 0.60 5.39 6.4 0.72 3.69 1.5 0.50 4.48 7.5 0.20 3.93 4.1
S 0.37 7.29 2.2 0.97 3.42 7.9 0.38 3.71 3.3 0.43 4.28 4.7 0.20 -1.73 5.2
SSW 0.63 8.48 3.1 0.52 4.11 4.1 0.57 5.09 6.0 0.57 6.88 0.9 0.39 2.01 9.8
SW 0.65 6.43 0.7 0.56 6.41 2.0 0.83 5.15 17.7 0.39 7.90 2.0 0.16 -9.17 4.4
WSW 0.62 8.56 0.6 0.56 6.86 2.6 0.65 6.35 9.8 0.52 7.37 4.0 0.09 -25.88 2.5
W 0.55 7.92 3.1 0.57 8.41 7.1 0.52 5.52 4.9 1.03 5.41 11.6 0.39 5.35 1.1
WNW 0.56 8.60 6.3 0.63 8.08 2.7 0.52 6.14 7.6 0.67 6.23 18.1 0.46 4.52 1.3
NW 0.63 6.15 3.1 0.65 7.36 10.3 0.64 5.23 18.5 0.56 7.25 8.5 0.22 -5.84 1.9
NNW 0.63 5.84 31.7 0.71 6.66 17.8 0.70 4.89 13.8 0.71 5.52 13.2 0.52 7.13 7.3
N 0.69 5.82 28.8 1.62 3.64 5.8 0.77 4.86 3.0 0.75 5.23 4.3 0.72 7.89 17.2
CHAPTER 6 WIND LOADS C6-73

Appendix 6.6 Dispersion of Wind Load

1. Factors influencing wind loads


The horizontal wind load for structural frames is obtained from Eq.(6.4), and the roof wind load for
structural frames is based on this equation.
WD = qH C D GD A (6.4)
where qH is velocity pressure, C D is wind force coefficient, GD is gust effect factor for
along-wind load and A is projected area at height Z .
The wind load for components/cladding is obtained form Eq.(6.6).
W = q C A
C H C C (6.6)
where qH is velocity pressure, CC is peak wind force coefficient and A is subject area.
The velocity pressure qH is expressed as Eq.(Appendix 6.6.1) form Eq.(A6.1) and Eq.(A6.2)
1 1
qH =U H2 = (U 0 K D EH k rW ) 2 (Appendix 6.6.1)
2 2
where is air density, U H is design wind speed, U 0 is basic wind speed, K D is wind
directionality factor, EH is wind speed profile factor at the reference height H and k rW is return
period conversion factor.
The factors influencing dispersion of horizontal wind load for structural frames WD and wind load
for components/cladding WC are air density , basic wind speed U 0 , wind directionality factor
K D , wind speed profile factor EH at reference height H according to the surface roughness, return
period conversion factor k rW , wind force coefficient C D and gust effect factor GD or peak wind
force coefficient CC .
The gust effect factor GD is influenced by design wind speed U H , turbulence intensity I H ,
turbulence scale LH , reference height H , building breadth B , building natural frequency f D ,
building critical damping ratio D and so on. The dispersion of these factors must be evaluated when
estimating the wind load on the frame for limit state design.

2. Dispersion of each factor


(1) Air density
The air density varies with temperature, atmospheric pressure and humidity, but the influence of
humidity can usually be ignored. In these recommendations, = 1.22 (kg/m3) at 15 and 1013hPa
can be used. The difference between this value and that for the range of 0, 1013hPa to 25, 960hPa
is within 10%.
(2) Basic wind speed U 0 and return period conversion factor k rW
For allowable stress design, the wind load can be obtained from Eq.(6.4) or Eq.(6.6) and
Eq.(Appendix 6.6.1) based on basic wind speed U 0 , wind directionality factor K D , wind speed
profile factor E and return period conversion factor k rW . For limit state design, however, the
maximum wind speed occurs during the buildings service life T years (T -year maximum value)
and its coefficient of variation is required. These recommendations provide maps for
C6-74 Recommendations for Loads on Buildings

100-year-recurrence basic wind speed U 0 and 500-year-recurrence wind speed U 500 based on the
annual maximum wind speed approximated by a Gumbel distribution. The mean value and the
standard deviation of the T -year maximum value can be obtained from these values based on the
method described in chapter 2. A calculated example for the mean value, the standard deviation and
the coefficient of variation of 50-year maximum values is shown in appendix Table 6.6.1. The
difference between U 500 and U 0 is 4m/s and the coefficient of variation is about 0.08 to 0.11 in
most areas other than the Okinawa Islands.

Appendix Table 6.6.1 Mean value, standard deviation and coefficient of variation for 50-year
maximum values of wind speed
50-year maximum value
coefficient of
city U 0 (m/s) U 500 (m/s) standard deviation
mean (m/s) variation
(m/s)
Sapporo 30.5 34.5 30.2 3.2 0.11
Aomori 31.0 35.0 30.7 3.2 0.10
Sendai 30.5 34.5 30.2 3.2 0.11
Niigata 37.0 41.0 36.7 3.2 0.09
Tokyo 36.0 40.0 35.7 3.2 0.09
Nagoya 32.5 36.5 32.2 3.2 0.10
Osaka 34.5 38.5 34.2 3.2 0.09
Hiroshima 30.0 34.0 29.7 3.2 0.11
Kochi 39.0 43.0 38.7 3.2 0.08
Fukuoka 33.5 37.5 33.2 3.2 0.10
Kagoshima 42.0 46.0 41.7 3.2 0.08

(3) Wind directionality factor


The wind directionality factor is decided in order to make the load effect using the wind
directionality factor equivalent to the load effect considering the wind direction. When the wind
directionality factor is considered, the standard deviation of the design wind speed is about 1m/s to
2m/s and its coefficient of variation is about 0.03 to 0.05 for each wind direction. However,
considering phenomena such as down-bursts, which cannot be caught enough, its lower limit of 0.85
and pitch of 0.05 are adopted. Furthermore, considering various uncertain parts, the maximum wind
directionality factor for adjacent wind directions is employed.
(4) Wind speed profile factor
Five flat terrain subcategories and wind speed profile factor EH corresponding to these flat terrain
subcategories are prescribed based on the observed data and the results calculated from computational
fluid dynamics. It is difficult to estimate differences between the actual values and the prescribed
values in consideration of the condition for the flat terrain subcategories of used data. When the flat
terrain subcategories entrusted to designer's judgment varies by one classification, the value of wind
CHAPTER 6 WIND LOADS C6-75

speed profile factor EH deviates 25% at H = 5 m, 15% at H = 100 m and 10% at H = 200 m, and
the coefficients of variation can be estimated as half their values as follows;
0.13 at H = 5 m
0.08 at H = 100 m
0.05 at H = 200 m
(5) Wind force coefficient, wind pressure coefficient
The case for a rectangular plan building is introduced here as an example for wind force coefficients
of horizontal wind load for structural frames of a building whose reference height is greater than 45m.
Wind tunnel test results obtained from reference papers and so on vary with aspect ratio and side ratio
of the building, and the wind force coefficients shown in Table A6.8 are their mean. For the vertical
distribution of wind force coefficient, test values at heights from 0.2 H to 0.9 H are mostly within
the range of 10% of these recommendation values. For the overturning moment coefficient at the
building base, most test results are within the range of 20% of these recommendation values. If a
building has a corner recess, the wind force coefficient generally takes a safe value. Therefore, if these
recommendations are adopted for such a building, its design is generally safe.
Horizontal wind force coefficients for structural frames of a rectangular plan building whose
reference height is 45m or less are influenced not only by building shape but also by many other
parameters such as wind characteristics. The values shown in Table A6.9(1) are simplified so that they
represent the results under various conditions. Therefore, their values are 10-30% greater than actual
ones, and 50% greater in some parts. They exceed 30% in part Lb when the roof slope is 30 or less,
but about 10-20% in parts WU and La . Furthermore, they may exceed 30% in part RLb when the
roof slope is less than 30 but about 10-20% in part RU on negative pressure parts and positive
pressure parts.
For the external pressure coefficient C pe , to calculate the roof wind load on structural frames
around the leading edge of the eave, for example, for B / H 6 and D / H > 1 , the spatial mean
value of the test results deviates within the range of 30% of these recommendation values of -1.0.
The positive and negative peak external pressure coefficients of the roof wind load for
components/claddings are determined from the maximum and minimum peak external pressures on
each part of the building for all wind directions. These values vary with wind profile, wind tunnel test
condition (such as sampling frequency, measuring position), side ratio and size reduction rate of the
test model and so on. Their coefficients of variation are about 0.2.
(6) Gust effect factor GD
The parameters that influence the gust effect factor GD of the horizontal wind load for structural
frames, excluding the height and the width of the building, are the natural frequency f D of the first
translational mode in the along-wind direction, the critical damping ratio D of the first translational
mode in along-wind direction, the design wind speed U H , turbulence scale LH , turbulence intensity
I H and the exponent of the power law in the wind speed profile. The influence of these
parameters on the gust effect factor varies with the flat terrain subcategory, the assumed building
C6-76 Recommendations for Loads on Buildings

shape and so on. Here, the reference height H = 80 m, the width B = 40 m, the natural frequency for
the first translational mode f D = 0.5 Hz, the critical damping ratio for the first translational mode
D = 2 %, the basic wind speed U 0 = 39 m/s and the flat terrain subcategory III are assumed. The
increase of the gust effect factor GD when each parameter is increased by 1% individually is
shown in appendix Table 6.6.2.

Appendix Table 6.6.2 Increase of gust effect factor GD when value of each parameter is
increased by 1% individually
parameter increase of gust effect factor GD
natural frequency f D 0.29%
critical damping ratio D 0.16%
design wind speed U H 0.34%
turbulence intensity I H 0.55%
turbulence scale LH 0.07%
exponent of power law 0.02%

For example, if the coefficient of variation of the critical damping ratio is 20%, that for the gust
effect factor caused by the critical damping ratio is estimated as 0.160.20=0.032.
Although the gust effect factor of the roof wind load for structural frames is influenced by various
parameters, the difference between the maximum loading effect for roof structural frames obtained
from these recommendations and the wind tunnel test results is within 15% and mostly around 30%.
(7) Natural frequency and critical damping ratio of first mode
Damping in Buildings7) proposed an estimation formula for the natural frequency and the critical
damping ratio of the first mode. When the dispersion of the values calculated from these proposed
formula is evaluated as the coefficient of variation of the difference between these recommendation
values and the field measurement values, the coefficient of variation of the natural frequency for the
first mode is about 0.1-0.5 for reinforced concrete structures, steel reinforced concrete structures and
steel structures, and that of the critical damping ratio for the first mode is about 0.2 for reinforced
concrete structure and steel reinforced concrete structures, about 0.3 for steel structures.
(8) Turbulence intensity I H
Fig.A6.1.17 compares the turbulence intensities of these recommendations and field measurements.
The coefficient of variation of the difference between these values can be estimated as about 0.2 for
flat terrain subcategory III where many field measurement data have been obtained.
(9) Turbulence scale LH
Fig.A6.1.21 compares the turbulence scales of these recommendations and field measurements. The
coefficient of variation of the difference between these values can be estimated as about 0.5.
CHAPTER 6 WIND LOADS C6-77

3. Coefficient of variation of wind load


The coefficient of variation of horizontal wind load for structural frames and of wind load for
components/claddings can be obtained from Eq.(Appendix 6.6.2) or Eq.(Appendix 6.6.3).

Horizontal wind load for structural frames: VWD = V2 + 4VU2H + VC2D + VG2D (Appendix 6.6.2)

Wind load for components/cladding : VWC = V2 + 4VU2H + VC2 (Appendix6.6.3)


C

where
VWD : coefficient of variation of horizontal wind load for structural frames WD
VWC : coefficient of variation of wind load for components/cladding WC
V : coefficient of variation of air density
VU H : coefficient of variation of design wind speed U H
VCD : coefficient of variation of wind force coefficient C D
VG D : coefficient of variation of gust effect factor GD
VC : coefficient of variation of peak wind force coefficient CC
C

When a building with reference height H = 80 m, width B = 40 m, natural frequency for first
translational mode f D = 0.5 Hz, and critical damping ratio for first translational mode D = 2 % is
constructed in a region of flat terrain subcategory III in each city of appendix Table 6.6.1, the
coefficient of variation VWD can be estimated as around 0.3 to 0.33 for wind load on structural frames
and the coefficient of variation VWC can be estimated as around 0.32 to 0.35 for wind load on
components/claddings.

References

1) Nishimura, H. and Taniike, Y: Aeroelastic Instability of High-Rise Building in a Turbulent


Boundary Layer, Journal of Structural and Construction Engineering AIJ, No.456, pp.31-.37, 1994
No.482, pp.27-32, 1996 (in Japanese)
2) Katagiri, J., Ohkuma, T., Marukawa, H. and Shimomura, S: Motion-induced wind forces acting on
rectangular high-rise buildings with side ratio of 2 Characteristics of motion-induced wind
forces during across-wind and torsional vibrations , Journal of Structural and Construction
Engineering AIJ, No.534, pp.25-32, 2000 (in Japanese)
3) Ohkuma, T., Katagiri, J., Marukawa, H. and Shimomura, S: Motion-induced wind forces and
coupled across-wind torsional unstable aerodynamic vibrations of rectangular high-rise buildings
with various side-ratio, Journal of Structural and Construction Engineering AIJ, No.560, pp.43-50,
2002 (in Japanese)
4) Fujimoto, M., Ohkuma, T: A theoretical study on evaluation of wind-induced vibrations of towers
with a circular cross section, Part I, Transactions of Architectural Institute of Japan, No.185,
C6-78 Recommendations for Loads on Buildings

pp.37-44, 1971 (in Japanese)


5) Ohkuma, T: A theoretical study on evaluation of wind-induced vibrations of towers with a circular
cross section, Part II, Transactions of Architectural Institute of Japan, No.187, pp59-67, 1971 (in
Japanese)
6) Ohkuma, T: A theoretical study on evaluation of wind-induced vibrations of towers with a circular
cross section, Part III, Transactions of Architectural Institute of Japan, No.188, pp25-32, 1971 (in
Japanese)
7) Damping in Buildings, AIJ, 2000 (in Japanese)
8) Davenport, A. G.: The application of statistical concepts to the wind loading of structures,
Proceeding of Institution of Civil Engineers, Vol.19, pp.449-472, 1961
9) Zhou, Y., Kareem, A: Gust loading factor: new model, Journal of Structural Engineering, ASCE,
pp.168-178, 2001
10) Ohkuma, T. , Marukawa, H: Mechanism of aero-elastically unstable vibration of large span roof,
Journal of Wind Engineering, No.42, January, pp.35-48, 1980
11) Matumoto, T: Wind tunnel study of self-excited oscillation of one-way type suspension roofs in a
smooth flow, Journal of Structural and Construction Engineering (Transactions of AIJ), No.384,
pp.90-96, 1988 (in Japanese)
12) Guide book for engineers on wind tunnel test of buildings: The building center of Japan, 1994
13) Ishihara, T., Hibi, K., Kato, H., Ohtake, K. and Matsui, M: A Database of Annual Maximum Wind
Speed and Corrections for Anemometers in Japan, Wind Engineers, No. 92, July, 2002
14) Ohtake, K. and Tamura, Y: Study on estimation of flat terrain for each wind direction, Summaries
of Technical Papers of Annual Meeting Architectural Institute of Japan 2002, B-1, pp. 119-120,
2002 (in Japanese)
15) Gomes, L., Vickery, B. J: Extreme wind speeds in mixed climates, Journal of Wind Engineering
and Industrial Aerodynamics, Vol.2, pp.331-334, 1978
16) Cook, N. J: Improving the Gumbel analysis by using M-th highest extremes, Revision of
ANSI/ASCE 7-95, 2000
17) ASCE 7-98: Minimum design loads for buildings and other structures, Revision of ANSI/ASCE
7-95, 2000
18) Matsui, M., Tamura, Y. and Tanaka, S: Wind load of Tall Buildings Considering Wind
Directionality Effects, Proceedings of 17th National Symposium on Wind Engineering, pp.
499-504, 2002 (in Japanese)
19) BS6399-2: British Standard, Loading for buildings, Part 2. Code of practice for wind loads, 1997
20) AS/NZS 1170.2: Australian/New Zealand Standard, Structural design actions, Part 2 : Wind
actions, 2002
21) Melbourne, W. H: Designing for directionality, 1st Workshop on Wind Engineering and Industrial
Aerodynamics, Highett, Victoria, pp.1-11, 1984
22) Suda, K., Sasaki, A., Ishibashi, R., Fujii, K., Hibi, K., Maruyama, T., Iwatani, Y., Tamura, Y:
CHAPTER 6 WIND LOADS C6-79

Observation of wind speed profiles in Tokyo city area using doppler soda, Proceedings of 16th
National Symposium on Wind Engineering, pp.13-18, 2000 (in Japanese)
23) Kondo, K., Kawai, H., Kawaguchi, A: Topographic multipliers for mean and fluctuating wind
velocities around up-slope cliffs, Summaries of Technical Papers of Annual Meeting, Architectural
Institute of Japan, pp.105-106, 2001 (in Japanese)
24) Kawai, H., Kondo, K: Topographic multipliers around micro-topography, Summaries of Technical
Papers of Annual Meeting, Architectural Institute of Japan, pp.103-104, 2003 (in Japanese)
25) Tsuchiya, M., Kondo, K., Kawai, H., Sanada, S: Effect of micro-topography on design wind
velocity, Summaries of Technical Papers of Annual Meeting, Architectural Institute of Japan,
pp.119-120, 1999 (in Japanese)
26) Meng, Y., Hibi, K: An experimental study of turbulent boundary layer over steep hills, Proceedings
of 15th National Symposium on Wind Engineering, pp.61-66, 1998
27) Goto, S., Suda, K. Miyashita, K: Profiles of turbulence intensity on the basis of full scale
measurements, Summaries of Technical Papers of Annual Meeting Architectural Institute of Japan
B-1, pp.111-112, 2002 (in Japanese)
28) Eurocode ENV 1991-2-4: 1997
29) Kamei, I. and Maruta, E: Wind Tunnel Test for evaluating wind pressure coefficients of buildings
with a gable roof -Part 2-, Summaries of Technical Papers of Annual Meeting Architectural
Institute of Japan, Structures, 1981, pp.1041-1042 (in Japanese)
30) Kanda, M. and Maruta; E: Study on design wind pressure coefficient of low rise buildings with a
flat roof or a gable roof -Part 3- Averaging wind pressure coefficient and wind direction,
Summaries of Technical Papers of Annual Meeting Architectural Institute of Japan, Structures ,
1992, pp.119-120 (in Japanese)
31) Ueda, H., Tamura, Y. and Fujii; K: Effect of turbulence of approaching wind on mean wind
pressures acting on flat roofs -Part 1- Study on characteristics of wind pressure acting on flat roofs,
Journal of Structural and Construction Engineering, No. 425, pp.91-99, 1991 (in Japanese)
32) Ueda, H., Hagura, H. and Oda, H: Characteristics of stress generated by wind pressures and wind
loads acting on stiff two-dimensional arches supporting a barrel roof, Journal of Structural and
Construction Engineering, AIJ, No. 496pp.29-35, 1997 (in Japanese)
33) Kikuchi, T., Ueda, H. and Hibi, K: Characteristics of wind pressures acting on the curved roofs,
Summaries of Technical Papers of Annual Meeting Architectural Institute of Japan, Structures ,
2003, pp.147-148 (in Japanese)
34) Nogichi, M., Uematsu, Y: Design wind pressure coefficients for spherical domes, Journal of Wind
Engineering, JAWE, No.95, 2003, pp.177-178 (in Japanese)
35) Chino, N. and Okada, H: Wind-induced internal pressures in buildings, Part 1 mean internal
pressures, Journal of Wind Engineering, JAWE, No.56, 1993, pp.11-20 (in Japanese)
36) Schewe, G: On the force fluctuations acting on a circular cylinder in crossflow from subcritical up
to transcritical Rynolds numbers, Journal of Fluid Mechanics, Vol.133, pp.265-285, 1983
C6-80 Recommendations for Loads on Buildings

37) Uematsu, Y., Yamada, M: Aerodynamic forces on circular cylinders of finite height, Journal of
Wind Engineering and Industrial Aerodynamics, Vol.51, pp.249-265, 1994
38) Uematsu, Y: Design wind force coefficients for free-standing canopy roofsJournal of Wind
Engineering, JAWE, No.95, 2003, pp.181-182 (in Japanese)
39) JEC-127, The Institute of Electrical Engineers of Japan 1979 (in Japanese)
40) Nishimura, H: Aerodynamic Characteristics of Shapes yielding Stagnated Flow, GBRC, No.106,
2002, pp.19-24 (in Japanese)
41) Nishimura, H, Asami, Y, Takamori K. and Okeya M: A wind tunnel study of fluctuating pressures
on buildings Part4 Pressure coefficient and drag coefficient, Summaries of Technical Papers of
Annual Meeting Architectural Institute of JAPAN, Structures I, 1992, pp.57-58 (in Japanese)
42) Katagiri, J, Kawabata, S, Niihori, Y, and Nakamura, O: Pressure Characteristics of Rectangular
cylinders with Cut Corner, Summaries of Technical Papers of Annual Meeting Architectural
Institute of Japan, Structures , 1992, pp.47-48 (in Japanese)
43) Ohtake, K: Peak wind pressure coefficients for cladding of a tall building - Part 1 Characteristics
of peak pressure, Summaries of Technical Papers of Annual Meeting Architectural Institute of
Japan, Structures , 2000, pp.193-194 (in Japanese)
44) Ohtake, : Peak wind pressure coefficients for cladding of a tall building - Part 2 Stretch of Peak
Wind Pressure, Summaries of Technical Papers of Annual Meeting Architectural Institute of Japan,
Structures , 2001, pp.143-144 (in Japanese)
45) Maruta, E., Ueda, H. and Kanda, M: Local wind pressure on gable roofs, Summaries of Technical
Papers of Annual Meeting College of Industrial Technology Nihon University, 1991, pp.33-36 (in
Japanese)
46). Uematsu, Y: Peak gust pressures acting on low-rise building roofs, Proceedings of the 8th East
Asia-Pacific Conference on Structural Engineering and Construction, Singapore, 2001
47) Uematsu, Y, Yamada, M: Fluctuating wind pressures on buildings and structures of circular
cross-section at high Reynolds numbers, Proceedings of the 9th International Conference on Wind
Engineering, New Delhi, 1995, pp.358-368
48) Okada, H. and Chino, N: Wind-induced internal pressures in buildings, Part 2 gust response factor
of internal pressure, Journal of Wind Engineering, JAWE, No.58, 1994, pp.43-53 (in Japanese)
49) Chino, N, Okada, H. and Kikitsu, H: On a new way to estimate wind load on cladding considering
correlation between external and internal pressures, Summaries of Technical Papers of Annual
Meeting Architectural Institute of Japan, Structures , 2000, pp.197-198 (in Japanese)
50). International Standard ISO4354, 1997
51) Asami, Y, Nakamura, O: A proposal for alongwind load model, Summaries of Technical Papers of
Annual Meeting, AIJ, Structures I, pp.195-196, 2002 (in Japanese)
52) Asami, Y, Kondo, K, Hibi, K: Experimental research of aerodynamic force on rectangular prism,
Journal of Wind Engineering, JAWE, 91, pp.83-88, 2002 (in Japanese)
53) Holmes, J. D: Effective static load distributions in wind engineering, Journal of Wind Engineering
CHAPTER 6 WIND LOADS C6-81

and Industrial Aerodynamics, Vol.90, pp.91-109, 2002


54) Harris, R. I: The propagation of internal pressures in buildings, Journal of Wind Engineering and
Industrial Aerodynamics, Vol.34, pp.169-184, 1990
55) Uematsu, Y, Sone, T, Noguchi, M: Geometric and structural characteristics and wind resistant
design method of spatial structures constructed in Japan, Journal of Wind Engineering, JAWE, 96,
pp.107-116, 2003 (in Japanese)
56) Marukawa, H. and Ohkuma, T: Formula of fluctuating wind forces for estimation of across-wind
and torsional responses of prismatic high rise buildingsJournal of Structural and Construction
Engineering AIJ, No.482, pp.33-42, 1996 (in Japanese)
57) Holmes, J. D: Along-wind response of lattice towers: Part I Derivation of expressions for gust
response factors, Engineering Structures, Vol.16, No.4, pp.33-42, 1996
58) Marukawa, H., Tamura, Y., Sanada, S., Nakamura, O: Lift and across-wind response of q 200m
concrete chimney, Journal of Wind Engineering JAWE, pp.37-52, 1984 (in Japanese)
59) Tamura, Y., Amano, A: Vortex induced vibration of circular cylinder Part III, Transactions of AIJ,
337, pp.65-72, 1984 (in Japanese)
60) Hibi, K., Tamura, Y., Kikuchi, H: Peak normal stresses and wind load combinations of middle-rise
buildings, Summaries of Technical Papers of Annual Meeting, AIJ, Structures I, pp.113-114, 2003
(in Japanese)
61) Asami, Y: Combination method for wind loads on high-rise buildings, Proceedings of the 16th
National Symposium on Wind Engineering, pp.531-534, 2000 (in Japanese)
62) Marukawa, H., Sasaki, A., Itou, J: Mode correction factor for modal wind forces, Summaries of
Technical Papers of Annual Meeting, AIJ, Structures I, pp.251-252, 1999 (in Japanese)
63) Takamori, K., Nishimura, H., Taniike, Y., Okazaki, M. Taniguchi, T: Aerodynamic interference
effect between twin tall buildings Part1 Square sectioned buildings, Summaries of Technical
Papers of Annual Meeting, Architectural Institute of Japan, pp.249-250, 2000 (in Japanese)
64) Takamori, K., Nishimura, H., Taniike, Y., Okazaki, M: Aerodynamic interference effect between
twin tall buildings Part2 Study of influences on the wind load, Summaries of Technical Papers of
Annual Meeting, Architectural Institute of Japan, pp.173-174, 2001 (in Japanese)
65) Taniike, Y: Interference effect between tall buildings with square section in a high-turbulent
boundary layer, Proceedings of 10th National Symposium on Wind Engineering, pp.247-252, 1988
(in Japanese)
66) Structural Design Concepts for Earthquake and Wind, Architectural Institute of Japan, 1999 (in
Japanese)

Das könnte Ihnen auch gefallen