Asce 7-22 CH 26com - For PC

You might also like

Download as pdf or txt
Download as pdf or txt
You are on page 1of 91

1 CHAPTER C26

2 WIND LOADS: GENERAL REQUIREMENTS


3

4 C26.1 PROCEDURES

5 Chapter 26 is the first of six chapters devoted to the wind load provisions. It provides the basic

LY
6 wind design parameters that are applicable to the various wind load determination methodologies
7 contained in Chapters 27 through 31. Specific items covered in Chapter 26 include definitions,

N
8 basic wind speed, exposure categories, internal pressures, elevation effects, enclosure

O
9 classification, gust effects, and topographic factors.

10 C26.1.1 Scope n
io
11 The procedures specified in this standard provide wind pressures and forces for the design of the
at

12 main wind force resisting system (MWFRS) and of components and cladding (C&C) of
m

13 buildings and other structures. The procedures involve the determination of wind directionality
14 and velocity pressure, the selection or determination of an appropriate gust-effect factor, and the
or

15 selection of appropriate pressure or force coefficients. The procedures account for the level of
nf

16 structural reliability required, the effects of differing wind exposures, the speed-up effects of
17 certain topographic features such as hills and escarpments, and the size and geometry of the
rI

18 building or other structure under consideration. The procedures differentiate between rigid and
Fo

19 flexible buildings and other structures, and the results generally envelope the most critical load
20 conditions for the design of the MWFRS, as well as C&C.
21 The pressure and force coefficients provided in Chapters 27, 28, 29, and 30 have been assembled
22 from the latest boundary-layer wind tunnel and full-scale tests and from previously available
23 literature. Because the boundary-layer wind tunnel results were obtained for specific types of
24 building, such as low-rise or high-rise buildings and buildings that have specific types of
25 structural framing systems, the designer is cautioned against indiscriminate interchange of values
26 among the figures and tables.

1
1 Tornado loads are treated separately from wind loads, as described in Section C32.1.
2 Storm Shelters. Storm shelters and safe rooms are designed with the intent of providing an
3 enhanced level of life-safety protection for a design storm event that is greater than typical for
4 conventional buildings and structures, including those designed to the wind loads in this chapter.
5 Since 2009, the model building and residential codes have required that structures designated as
6 storm shelters be designed according to ICC 500 (ICC 2020), the ICC/NSSA Standard for Design
7 and Construction of Storm Shelters, which applies to hurricane, tornado, and combined hurricane
8 and tornado shelters. The load provisions of ICC 500 directly reference ASCE 7, with some

LY
9 modifications including greater hazard levels.
10 FEMA also publishes guidance on the design and construction of structures for storm protection

N
11 in FEMA P-361, Safe Rooms for Tornadoes and Hurricanes: Guidance for Community and
12 Residential Safe Rooms (2021a), and FEMA P-320, Taking Shelter from the Storm: Building a

O
13 Safe Room for Your Home (2021b). FEMA P-361 provides comprehensive guidance for the
14
n
design of community and residential safe rooms, as well as operations and maintenance
io
15 guidance. Although the criteria in FEMA P-361 and ICC 500 are quite similar, “safe room” is
16 FEMA terminology for storm shelters that also meet the criteria in FEMA P-361. All safe room
at

17 criteria in FEMA P-361 meet the storm shelter requirements in ICC 500, but a few design and
m

18 performance criteria in FEMA P-361 are more restrictive than those in ICC 500 and are required
19 for safe room projects to be eligible for federal grant funding. FEMA P-320 provides prescriptive
or

20 solutions for residential and small business safe rooms.


nf

21 Storm shelters and safe rooms are designed for more extreme hazard levels than conventional
22 buildings. For example, where Chapter 26 uses wind speeds with return periods of 700, 1,700,
rI

23 and 3,000 years to determine the design wind loads for Risk Category II, III, and IV buildings,
Fo

24 respectively, the wind loads for hurricane shelters and safe rooms are based on hurricane wind
25 speed maps having a 10,000-year mean recurrence interval (MRI). “The design wind speeds
26 chosen by FEMA for safe room guidance were determined with the intent of specifying near-
27 absolute protection with an emphasis on life safety” (FEMA 2021a). Rain loads and flood loads
28 are also based on hazard levels with greater return periods than used for Risk Category II through
29 IV buildings. In addition, hurricane shelters and safe rooms have much more stringent wind-
30 borne debris requirements than ASCE 7. The cumulative effect of these and other differences
31 between ICC 500 and ASCE 7 results in storm shelters providing greater life-safety protection

2
1 from hurricanes than buildings designed to ASCE 7 Risk Category IV. Although ICC 500 does
2 not explicitly state a performance objective for shelters designed to that standard, safe rooms
3 constructed to the almost-identical FEMA criteria are intended to “provide near-absolute
4 protection from wind and wind-borne debris for occupants.” Since 1998, when FEMA first
5 published guidance for residential and community safe rooms, “thousands of safe rooms have
6 been built, and a growing number of these safe rooms have already saved lives in actual events.
7 There has not been a single reported failure of a safe room constructed to FEMA criteria”
8 (FEMA 2021a). Similarly, failures of hurricane shelters constructed to ICC 500 criteria (first

LY
9 published in 2008) have not been reported.

10 C26.1.2 Permitted Procedures

N
O
11 The wind load provisions provide several procedures (as illustrated in Figure 26.1-1) from which
12 the designer can choose.
13 n
io
14 For MWFRS:
at
15 1. Directional Procedure for buildings of all heights (Chapter 27).
16 2. Envelope Procedure for low-rise buildings (Chapter 28).
m

17 3. Directional Procedure for building appurtenances and other structures (Chapter 29).
or

18 4. Wind Tunnel Procedure for all buildings and other structures (Chapter 31).
19
nf

20 For C&C:
rI

21 1. Analytical Procedure for buildings and building appurtenances (Chapter 30).


22 2. Wind Tunnel Procedure for all buildings and other structures (Chapter 31).
Fo

23
24
25 The “simplified methods” for the determination of the MWFRS design pressures that were
26 contained in previous editions of this standard have been removed from this edition. The Wind
27 Load Subcommittee evaluated the need for the tabular methods and concluded that ASCE 7
28 should generally not tabulate calculated design pressures for specific buildings and that the
29 tabular methods belong in a guide or other resource.

3
1 Limitations. The provisions given under Section 26.1.2 apply to the majority of site locations
2 and buildings and other structures, but for some projects, these provisions may be inadequate.
3 Examples of site locations and buildings and other structures (or portions thereof) that may
4 require other approved standards, special studies using applicable recognized literature
5 pertaining to wind effects, or using the Wind Tunnel Procedure of Chapter 31, include
6
7 1. Site locations that have channeling effects or wakes from upwind obstructions.
8 Channeling effects can be caused by topographic features (e.g., a mountain gorge) or

LY
9 buildings (e.g., a neighboring tall building or a cluster of tall buildings). Wakes can be
10 caused by hills, buildings, or other structures. These effects can increase the local wind

N
11 speed and alter the action of wind on a subject building or structure. For flexible or
12 dynamically sensitive buildings, such interference effects can increase the resonant

O
13 dynamic wind loads and responses. Situations where interference effects may be
14 significant are listed in Item 4 below.
n
io
15 2. Buildings with unusual or irregular geometric shape, including barrel vaults, arched
16 roofs, and other buildings whose shape (in plan or vertical cross section) differs
at

17 significantly from the shapes in Figures 27.3-1, 27.3-2, 27.3-3, 27.3-7, 28.3-1, and 30.3-1
m

18 through 30.3-7. Unusual or irregular geometric shapes include buildings with multiple
19 setbacks, curved façades, or irregular plans resulting from significant indentations or
or

20 projections, openings through the building, or multitower buildings connected by bridges.


nf

21 3. Buildings or other structures with response characteristics that result in substantial


22 vortex-induced and/or torsional dynamic effects, or dynamic effects resulting from
rI

23 aeroelastic instabilities such as flutter or galloping. Such dynamic effects are difficult to
Fo

24 anticipate, being dependent on many factors, but should be considered when any one or
25 more of the following apply:
26  Height of the building or other structure is more than 400 ft (122 m).
27  Height of the building or other structure is greater than 4 times its minimum effective
28 width, Bmin , as defined following.

29  Lowest natural frequency of the building or other structure is less than n1  0.25 Hz .

4
1  Reduced velocity Vz / ( n1Bmin )  5 , where z  0.6h and Vz is the mean hourly

2 velocity in ft/s (m/s) at height z .

3 The minimum effective width Bmin is defined as the minimum value of h B / h


i i i

4 considering all wind directions. The summations are performed over the height of the
5 building or other structure for each wind direction, where hi is the height above grade of

6 level i and Bi is the width at level i normal to the wind direction.

7 4. Flexible or dynamically-sensitive buildings with one or more upwind building(s) for

LY
8 which interference effects, such as channeling effects or wakes, may be significant.
9 Interference effects can increase the resonant dynamic loads and response of a

N
10 dynamically-sensitive building, and can be significant when all of the following

O
11 characteristics apply to the building under study:
12  The height of the subject building is greater than 2 times the average height of
13
n
surrounding buildings within a distance of 10Bmin, where Bmin is defined in Item 3
io
14 above.
at
15  The presence of one or more upwind building(s) with a height comparable to or
16 greater than the height of the subject building, and located within a distance of less
m

17 than 10Bmin of the subject building.


or

18  The height of the building or other structure is greater than 4 times its minimum
19 effective width, and reduced velocity is greater than 5 ft/s (m/s), and the lowest
nf

20 natural frequency (n1) is less than 0.25 Hz (i.e., n1 < 0.25 Hz).
rI

21 Interference effects are generally more significant for site locations characterized by wind
22 with lower turbulence (e.g., relatively smooth upstream fetches with few obstructions
Fo

23 5. Bridges, cranes, electrical transmission lines, guyed masts, highway signs and lighting
24 structures, telecommunication towers, and flagpoles.
25
26 When undertaking detailed studies of the dynamic response to wind forces, the fundamental
27 frequencies of the building or other structure in each direction under consideration should be
28 established using the structural properties and deformational characteristics of the resisting
29 elements in a properly substantiated analysis, and not using approximate equations based on
30 height.

5
1 Shielding. Because of the lack of reliable analytical procedures for predicting the effects of
2 shielding provided by buildings and other structures or by topographic features, reductions in
3 velocity pressure caused by shielding are not permitted under the provisions of this chapter.
4 However, this does not preclude the determination of shielding effects and the corresponding
5 reductions in velocity pressure by means of the Wind Tunnel Procedure in Chapter 31.

6 C26.1.3 Performance-Based Procedures

7 Performance-Based Procedures for wind design can quantify service interruption or loss due to

LY
8 wind, improve wind resistance, or perform an enhanced evaluation of the MWFRS and/or
9 envelope. Performance-based procedures identify the source of loss or service interruption and

N
10 seek to mitigate undesirable outcomes due to design decisions informed by design objectives.

O
11 Selection of design objectives should consider the ramifications of structural and nonstructural
12 response for the building occupants, the ability of the building or structure to provide services
13 n
following and/or during a wind event, and maintenance of structural safety expressed by Table
io
14 1.3-1. The evaluation should consider strength-level wind effects, service-level wind effects, and
at
15 occupant-comfort wind effects, which all influence building or facility response and the resulting
16 ability to recover functionality following and/or during a wind event (also termed resilience).
m

17 Performance design objectives should consider that, historically, a large proportion of damage
or

18 from high winds has occurred through envelope component damage, including water penetration
19 through leakage and breaches in the building envelope. In addition, direct loss or interruption of
nf

20 service can occur due to structural motions, causing occupant discomfort and/or nonstructural
rI

21 damage. Using approved methods, the designer is permitted to evaluate the performance of a
22 building, facility, or structural system using performance-based wind engineering procedures
Fo

23 documented in the recognized literature, such as the ASCE/SEI Prestandard for Performance-
24 Based Wind Design (2019). The performance results should demonstrate consistency with
25 approved performance criteria. An independent peer review should be conducted as part of a
26 review of the performance-based design where the design takes exception to provisions of
27 Chapters 26 to 31.
28 Major advancements introduced by ASCE/SEI (2019) include limited inelasticity in the MWFRS
29 elements, nonlinear dynamic analysis for wind design, system-based performance criteria, and
30 enhanced design criteria for the building envelope.

6
1 The maps provided in Appendix F can be referenced where a performance-based wind
2 structural reliability assessment requires wind hazards with long return periods.

3 C26.2 DEFINITIONS

4 Several important definitions given in the standard are discussed in the following text. These
5 terms are used throughout the standard and are provided to clarify application of the standard
6 provisions.
7 BUILDING, ENCLOSED; BUILDING, OPEN; BUILDING, PARTIALLY ENCLOSED;

LY
8 BUILDING, PARTIALLY OPEN: These definitions relate to the proper selection of internal
9 pressure coefficients, ( GC pi ). “Enclosed,” “open,” and “partially enclosed” buildings are

N
10 specifically defined. All other buildings are considered to be “partially open” by definition,

O
11 although there may be large openings in two or more walls. An example of this would be a
12 parking garage through which the wind can easily pass but which does not meet the definition for
13 n
either an open or a partially enclosed building. The internal pressure coefficient for such a
io
14 building would be 0.18 , and the internal pressures would act on the solid areas of the walls and
at
15 roof. The standard also specifies that a building that meets both the “open” and “partially
16 enclosed” definitions should be considered “partially open.”
m

17 BUILDING OR OTHER STRUCTURE, FLEXIBLE: A building or other structure is


or

18 considered “flexible” if it contains a significant dynamic resonant response. Resonant response


nf

19 depends on the gust structure contained in the approaching wind, on wind loading pressures
20 generated by the wind flow about the building, and on the dynamic properties of the building or
rI

21 structure. Gust energy in the wind is smaller at frequencies above about 1 Hz. Therefore, the
Fo

22 resonant response of most buildings and structures with lowest natural frequency above 1 Hz are
23 sufficiently small that resonant response can often be ignored. The natural frequency of buildings
24 or other structures greater than 60 ft (18.3 m) in height is determined in accordance with Sections
25 26.11.1 and 26.11.2. When buildings or other structures have a height exceeding 4 times the least
26 horizontal dimension or when there is reason to believe that the natural frequency is less than
27 1 Hz (natural period greater than 1 s), the natural frequency of the structure should be
28 investigated. Approximate equations for natural frequency or period for various building and

7
1 structure types in addition to those given in Section 26.11.2 for buildings are contained in
2 Commentary Section C26.11.
3 BUILDING OR OTHER STRUCTURE, REGULAR-SHAPED: Defining the limits of
4 applicability of the various procedures in the standard requires a balance between the practical
5 need to use the provisions past the range for which data have been obtained and restricting use of
6 the provisions past the range of realistic application. Wind load provisions are based primarily on
7 wind tunnel tests on shapes shown in Figures 27.3-1, 27.3-2, 27.3-3, 27.3-7, 28.3-1, and 30.3-1
8 through 30.3-7. Extensive wind tunnel tests on actual structures under design show that relatively

LY
9 large changes from these shapes can, in many cases, have minor changes in wind load, while in
10 other cases seemingly small changes can have relatively large effects, particularly on cladding

N
11 pressures. Wind loads on complicated shapes are frequently smaller than those on the simpler
12 shapes of Figures 27.3-1, 27.3-2, 27.3-7, 28.3-1, and 30.3-1 through 30.3-7, and so wind loads

O
13 determined from these provisions are expected to envelop most structure shapes. Buildings or
14
n
other structures that are clearly unusual should be designed using the Wind Tunnel Procedure of
io
15 Chapter 31.
16 BUILDING OR OTHER STRUCTURE, RIGID: The defining criterion for “rigid,” in
at

17 comparison to “flexible,” is that the natural frequency is greater than or equal to 1 Hz. A general
m

18 guidance is that most rigid buildings and structures have height-to-minimum-width less than 4.
19 The provisions of Sections 26.11.1 and 26.11.2 provide methods for calculating natural
or

20 frequency (period = 1/natural frequency), and Commentary Section C26.11 provides additional
nf

21 guidance.
22 COMPONENTS AND CLADDING (C&C): Components receive wind loads directly or from
rI

23 cladding and transfer the load to the MWFRS. Cladding receives wind loads directly. Examples
Fo

24 of components include, but are not limited to, fasteners, purlins, girts, studs, sheathing, roof
25 decking, certain trusses, and elements of trusses receiving wind loads from cladding. Examples
26 of cladding include, but are not limited to, wall coverings, curtain walls, roof coverings,
27 sheathing, roof decking, exterior windows, and doors. Components can be part of the MWFRS
28 when they act as elements in shear walls or roof diaphragms, but they may also be loaded
29 directly by wind as individual elements. The designer should use appropriate loads for design of
30 components, which may require certain components to be designed for more than one type of
31 wind loading; for example, long-span roof trusses should be designed for loads associated with

8
1 MWFRS, and individual members of trusses should also be designed for C&C loads (Mehta and
2 Marshall 1998).
3 DIAPHRAGM: This definition for diaphragm in wind load applications, for the case of
4 untopped steel decks, differs somewhat from the definition used in Section 12.3 because
5 diaphragms under wind loads are expected to remain essentially elastic.
6 EFFECTIVE WIND AREA, A: Effective wind area is the area of the building surface used to
7 determine ( GCp ). This area does not necessarily correspond to the area of the building surface

8 contributing to the force being considered. Two cases arise. In the usual case, the effective wind

LY
9 area does correspond to the area tributary to the force component being considered. For example,
10 for a cladding panel, the effective wind area may be equal to the total area of the panel. For a

N
11 cladding fastener, the effective wind area is the area of cladding secured by a single fastener. A

O
12 mullion may receive wind from several cladding panels. In this case, the effective wind area is
13 the area associated with the wind load that is transferred to the mullion.
14
n
The second case arises where components such as roofing panels, wall studs, or roof trusses are
io
15 closely spaced. The area served by the component may become long and narrow. To better
at
16 approximate the actual load distribution in such cases, the width of the effective wind area used
17 to evaluate ( GCp ) need not be taken as less than one-third the length of the area. This increase in
m

18 effective wind area has the effect of reducing the average wind pressure acting on the
or

19 component. Note, however, that this effective wind area should only be used in determining (
GCp ) in Figures 30.3-1 through 30.3-6. The induced wind load should be applied over the actual
nf

20

21 area tributary to the component being considered.


rI

22 For membrane roof systems, the effective wind area is the area of an insulation board (or deck
Fo

23 panel if insulation is not used) if the boards are fully adhered (or the membrane is adhered
24 directly to the deck). If the insulation boards or membrane are mechanically attached or partially
25 adhered, the effective wind area is the area of the board or membrane secured by a single
26 fastener or individual spot or row of adhesive.
27 For windows, doors, and other fenestration assemblies, the effective wind area for typical single-
28 unit assemblies can be taken as the overall area of the assembly. For assemblies comprised of
29 more than one unit mulled together or for more complex fenestration systems, it is recommended

9
1 that the fenestration product manufacturer be consulted for guidance on the appropriate effective
2 wind area to use when calculating the design wind pressure for product specification purposes.
3 The definition of effective wind area for rooftop solar panels and arrays is similar to that for
4 C&C. As with C&C, the width of the effective wind area need not be less than one-third its
5 length (which is typically equal to the span of the framing element being considered). The
6 induced wind pressure is calculated per Figure 29.4-4 using this effective wind area, and the
7 wind pressure is then applied over the actual area tributary to the element.
8 Effective wind area is equal to the tributary area, except in cases where the exception is invoked

LY
9 that the width of the effective wind area need not be less than one-third its length. In such cases,
10 the effective wind area can be taken as larger than the tributary area.

N
11 Tributary area for a spanning structural member of a solar array depends on the span length of
12 that member times the perpendicular distances to adjacent parallel members. For a support point

O
13 or fastener, tributary area depends on the span of members framing into that support point.
14
n
Tributary area (and effective wind area) can depend on the characteristics of the solar array
io
15 support system and the load path. For a roof-bearing system that has different load paths for
16 upward, downward, and lateral forces, the appropriate effective wind area for each direction of
at

17 forces is used.
m

18 If the support system for the solar array has adequate strength, stiffness, and interconnectedness
19 to span across a support or ballast point that is subject to yielding or uplift, the effective wind
or

20 area can be correspondingly increased, provided strengths are not governed by brittle failure and
nf

21 the deformation of the array is evaluated and does not result in adverse performance. It should be
22 noted that effective wind areas for uplift are usually much smaller than for lateral (drag) forces
rI

23 for ballasted arrays.


Fo

24 MAIN WIND FORCE RESISTING SYSTEM (MWFRS): The MWFRS can consist of a
25 structural frame or an assemblage of structural elements that work together to transfer wind loads
26 acting on the entire building or structure to the ground. Structural elements such as cross-bracing,
27 shear walls, roof trusses, and roof diaphragms are part of the MWFRS when they assist in
28 transferring overall loads (Mehta and Marshall 1998).
29 WIND-BORNE DEBRIS REGIONS: Wind-borne debris regions are defined to alert the
30 designer to areas requiring consideration of missile impact design. These areas are located within

10
1 hurricane-prone regions where there is a high risk of glazing failure caused by the impact of
2 wind-borne debris.

3 C26.3 SYMBOLS

4 The following additional symbols and notation are used herein.


5 Aob = Average area of open ground surrounding each obstruction;
6 n = Reference period, years;
7 Pa = Annual probability of wind speed exceeding a given magnitude [Equation (C26.5-3)];

LY
8 Pn = Probability of exceeding design wind speed during n years [Equation (C26.5-3)];

N
9 Sob = Average frontal area presented to the wind by each obstruction;

O
10 Vt = Wind speed averaged over t seconds (see Figure C26.5-1), mi/h (m/s);

11 V3600 = Mean wind speed averaged over 1 h (see Figure C26.5-1), mi/h (m/s); and
n
io
12 β = Damping ratio (percentage of critical damping).
13
at

14
m
or
nf
rI
Fo

15
16 Figure C26.5-1. Maximum speed averaged over t (s) to hourly mean speed.
17

18 C26.4 GENERAL

19 C26.4.3 Wind Pressures Acting on Opposite Faces of Each Building Surface

11
1 Section 26.4.3 is included in the standard to ensure that internal and external pressures acting on
2 a building surface are taken into account by determining a net pressure from the algebraic sum of
3 those pressures. For additional information on the application of the net C&C wind pressure
4 acting across a multilayered building envelope system, including air-permeable cladding, refer to
5 Section C30.1.1.

6 C26.5 WIND HAZARD MAP

7 C26.5.1 Basic Wind Speed

LY
8 All the wind speed maps in ASCE 7-16 have been updated, based on (1) a new analysis of non-

N
9 hurricane wind data available through 2010, and (2) improvements to the hurricane simulation
10 model, which better account for the translation speed effects of fast-moving storms and the

O
11 transition from hurricanes to extratropical storms in the northern latitudes (i.e., transition from
12
n
warm-core to cold-core low-pressure systems). Separate wind speed maps are now provided for
io
13 Risk Category III and IV buildings and structures, recognizing the higher reliabilities required
14 for essential facilities and facilities whose failure could pose a substantial hazard to the
at

15 community. Location-specific basic wind speeds may be determined from the ASCE Wind
m

16 Design Geodatabase using the ASCE 7 Hazard Tool (https://asce7hazardtool.online/). This


17 website provides wind speeds to the nearest mile per hour based on a location defined using
or

18 either latitude/longitude or an address. The website results use the same data used to develop the
nf

19 paper maps currently in the standard, as well as wind speeds for Hawaii, Puerto Rico, and the US
20 Virgin Islands. Wind speeds are provided to the user for each of the Risk Categories and each of
rI

21 the serviceability periods.


Fo

22 In the 2016 edition, microzoned “effective” wind speed maps for Hawaii were added in
23 accordance with the strength design return periods, including the effect of topography. The
24 Hawaii effective wind speeds are algebraically formulated to include the macroscale and
25 mesoscale terrain-normalized values of K zt and Kd (Chock et al. 2005); that is, Veffective is the

26 basic windspeed V multiplied by  ( K zt  Kd / 0.85) , so that the engineer is permitted to more

27 conveniently use the standard values of K zt of 1.0 and Kd as given in Table 26.6-1. Note that

28 local site conditions of finer toposcale, such as ocean promontories and local escarpments,

12
1 should still be examined. Spatial resolution scales for digital modeling, including terrain effects,
2 are conventionally described in the recognized literature as follows:
Scale Spatial Resolution
Toposcale 32–656 ft (10–200 m)
Mesoscale 656 ft3.1 mi (200 m–5 km)
Macroscale 3.1–311 mi (5–500 km)
3
4 In the 2022 edition, microzoned “effective” wind speed maps for US Virgin Islands, Puerto Rico,
5 and the municipal islands of Culebra and Vieques were added in accordance with the strength

LY
6 design return periods. The effective wind speeds include the topographic effects which were
7 quantified by the factor of Kzt. Thus, the design professional is permitted to evaluate effective

N
8 wind speeds from these microzoned wind speed maps as a simplified method to determine wind

O
9 loads and pressures (using Kzt = 1.0 during calculation) on a building or structure.
10 To develop the microzoned wind maps, a wind speed-up model has been developed to evaluate
11
n
wind speed-ups in 16 directional sectors for 36,973 locations in the US Virgin Islands and
io
12 828,924 locations distributed among Puerto Rico and the municipal islands of Culebra and
13 Vieques (ARA 2019). These locations comprise a grid with a resolution of 0.00095 degrees
at

14 (about 100 m or 328 ft). A 1/3-arcsecond digital elevation model was used to create elevation
m

15 profiles for use in the wind speed-up calculations. The directional wind speed-up data were
or

16 combined with the directional hurricane winds from a hurricane simulation model to develop
17 effective wind speeds which include the effect of topographically induced wind speed-ups.
nf

18
rI

19 The decision in ASCE 7-10 to move to separate strength design wind speed maps for different
20 risk categories in conjunction with a wind load factor of 1.0, instead of using a single map in
Fo

21 conjunction with an Importance Factor and a load factor of 1.6, relied on several considerations:
22 1. A strength-level design wind speed map brings the wind loading approach in line with
23 that used for seismic loads, in that they both are aimed at achieving uniform risk rather
24 than uniform hazard and eliminate the use of a load factor for strength design.
25 2. Having separate maps removes inconsistencies that occurred with the use of Importance
26 Factors, which varied with location, and allows the geographical description of zones
27 affected by non-hurricane winds only and by both hurricane and non-hurricane winds as a
28 function of MRI.

13
1 3. Each map has the same MRI for design wind speeds in those two zones.
2 4. By providing the design wind speed directly, the maps more clearly inform owners and
3 their consultants about the storm intensities for which the buildings and other structures
4 are designed.
5
6 Selection of Return Periods. The methodology for selection of the return periods used in ASCE
7 7-10 (Vickery et al. 2010) has been modified for ASCE 7-16. To determine a return period for
8 each Risk Category consistent with the target reliabilities in Table C1.3-1, the ASCE 7 Load

LY
9 Combinations Subcommittee conducted a reliability analysis that incorporated new data on the
10 directionality factor. The nominal design value Kd  0.85 was based on a relatively simple

N
11 directional analysis conducted as part of the original ANSI A58/ASCE 7 load factor

O
12 development. One of the underlying assumptions of the original analysis was that the wind
13 directionality factor, Kd  0.85 , was unbiased because only limited data on the effects of wind

14 n
directionality were available at the time. More recent research by Isyumov et al. (2013),
io
15 simulating three building geometries at two different locations, indicates that the ASCE 7
at
16 nominal values of Kd are affected by a bias (defined as the ratio of the mean value, μKd , to the
m

17 nominal value, K dn ). The nominal value of Kd is conservative for both tropical and extratropical
or

18 winds because the mean value is less than the nominal value. Additional reliability analyses were
19 performed to examine the effect of Kd on the return period and associated reliability. The
nf

20 subcommittee found that the following return periods for each Risk Category are consistent with
rI

21 the target reliabilities in the first row of Table 1.3-1: Risk Category I, 300 years; Risk Category
22 II, 700 years; Risk Category III, 1,700 years; Risk Category IV, 3,000 years.
Fo

23 Wind Speed. The wind speed maps of Figure 26.5-1 present basic wind speeds for the
24 contiguous United States, Alaska, Hawaii, and other selected locations. In ASCE 7-22, wind
25 speeds are included for the first time for the US territory of the Northern Mariana Islands, which
26 match those of nearby Guam. The wind speeds correspond to 3-s gust speeds at 33 ft (10 m)
27 above ground for Exposure Category C. Because the wind speeds of Figures 26.5-1 and 26.5-2
28 reflect conditions at airports and similar open-country exposures, they do not account for the
29 effects of significant topographic features such as those described in Section 26.8. In ASCE 7-
30 16, wind speeds in non-hurricane-prone areas of the contiguous United States are mapped using

14
1 contours to better reflect regional variations in the extreme wind climate. Point values are
2 provided to aid interpolation, in a style similar to that used in the ASCE 7 seismic hazard maps.
3 Summaries of the data and methods used to estimate both the non-hurricane and hurricane wind
4 speeds are given below, along with a description of how these wind speeds are combined to
5 make the final maps. Detailed descriptions are provided by Pintar et al. (2015) and Lombardo
6 et al. (2016).
7 Non-hurricane Wind Speeds. The non-hurricane wind speeds for the contiguous United States
8 were estimated from peak gust speed data collected at 575 meteorological stations. The data at

LY
9 each station were extracted from the meteorological records and classified by storm type,
10 thunderstorm or non-thunderstorm after removal of gusts associated with tropical cyclones

N
11 (i.e., hurricanes and tropical storms). Recorded peak gusts from each station were corrected as
12 needed to standardize the observations to equivalent 3-s peak gusts at 33 ft (10 m) height over

O
13 open (Exposure C) terrain. At each station, there were at least 15 years of data, and there were
14
n
sufficient numbers of both thunderstorm and non-thunderstorm observations to account for their
io
15 potential differences when estimating wind speeds with specified MRIs. The estimation was
16 performed in two stages. In the first stage, a peaks-over-thresholds (POT) model was fitted to the
at

17 data from each station. The POT model used was the Poisson process model first described by
m

18 Pickands (1971) and extended by Smith (1989) to allow the parameters of the Poisson process to
19 be time-dependent. This model allowed differentiation between thunderstorm and non-
or

20 thunderstorm winds. The Poisson process has a tail length parameter that may be set to zero,
nf

21 leading to Gumbel-like tails for the distribution of wind speeds. Such distributional tails were
22 used in this work, consistent with past practice in wind engineering. The fitted POT models
rI

23 allowed the estimation of wind speeds for any required MRI at all stations. In the second stage,
Fo

24 local regression (Cleveland and Devlin 1988) was used to interpolate wind speeds at all points of
25 a fine regular grid covering the contiguous United States for all required MRIs. This had the
26 effect of spatially smoothing the noisy station estimates. The smoothed wind speed estimates
27 provided the basis for creating the isotach maps.
28 Limited data were available on the Washington and Oregon coast. In this region, a special wind
29 region was defined to permit local jurisdictions to select speeds based on local knowledge and
30 analysis. Speeds in the Aleutian Islands and in the interior of Alaska were established from gust
31 data. Insufficient data were available for a detailed coverage of the mountainous regions, so gust

15
1 data in Alaska were not corrected for potential terrain influence. It is possible that wind speeds in
2 parts of Alaska would be lower if the topographic wind speed-up effect on recorded wind speeds
3 were taken into account. In Alaska, the maps for each return period were determined by
4 multiplying the 50-year MRI contours given in ASCE 7-10 Figure CC-3 by a factor, FRA,
5 FRA = 0.45 + 0.085 ln(12T) (C26.5-1)
6 where T is the return period in years (Peterka and Shahid 1998). The resulting contours were
7 interpolated to the nearest 10 mi / h , except for the innermost and outermost contours, which

8 were rounded to the nearest 5 mi / h .

LY
9 Hurricane Wind Speeds. The hurricane wind speeds are based on the results of a Monte Carlo
10 simulation model generally described by ARA (2001), Vickery and Wadhera (2008a, b), and

N
11 Vickery et al. (2009a, b, 2010). The hurricane simulation model used to develop the wind speeds

O
12 in ASCE 7-16 includes two updates to the model used for ASCE 7-10. A reduced translation
13 speed effect for fast-moving storms (USNRC 2011) was incorporated, and a simple extratropical
14 n
transition model was also implemented, where the surface winds are reduced linearly by up to
io
15 10% over the latitude range 37 N to 45 N. This reduction approximates transitioning from a
at
16 hurricane boundary layer to an extratropical storm boundary layer. The effects of the model
17 revisions are to slightly reduce hurricane speeds in the northeast, extending from Maine to
m

18 Virginia.
or

19 In addition to the model changes noted above, since the 2010 Edition of ASCE 7 a number of
20 minor improvements have been made to the hurricane hazard and wind field models.
nf

21 The first change includes a reduction in the translation speed effect for storms translating at
rI

22 speeds of 10 m/s (4.5 m/s) or more (formerly the reduction in the translation speed effect was
23 used for storms translating at 15 m/s (6.7 m/s) or more, which reduced wind speeds in the New
Fo

24 York / New England area). This change affects wind speeds associated with storms in the Gulf of
25 Mexico that have recurved, or are recurving, to have a northerly through northeasterly heading,
26 typically affecting a portion of storms making landfall along the middle and eastern portions of
27 the Gulf of Mexico coastline. The translation speed change also affects storms moving in a north
28 to northeasterly direction along the Atlantic coast, particularly for storms making landfall along
29 Long Island, NY, and points further north.
30 The hurricane contours along the US coastline for locations south of the North Carolina–Virginia
31 border in the 2016 Edition of ASCE 7 are the same as those given in the 2010 Edition of ASCE 7

16
1 (except for the 3,000-year contours) and were developed using a hurricane climatology based on
2 the period 1900 through 2006. The hurricane contours were blended with the non-hurricane wind
3 speeds adjacent to the hurricane coastline, which, at the time of the 2010 edition of ASCE 7,
4 were a uniform 115 mi/h ( 51 m/s) (700-year). At the time of the 2016 Edition of ASCE 7 the
5 non-hurricane wind speeds adjacent to the hurricane coastline varied with location, but were
6 approximately 105 mi/h (47 m/s) (700-year). In the 2016 Edition of ASCE 7, the hurricane
7 contours south of the NC–VA border were left unchanged, and the transition between the new,
8 lower non-hurricane wind speeds was performed using judgment, likely resulting in an

LY
9 overestimate of wind speeds in parts of the transition region. For example, along the south Texas
10 coast the ASCE 7-22 contours increased by approximately 10 mi/h (4.5 m/s), from 150 mi/h (67

N
11 m/s) in ASCE 7-10 and ASCE 7-16 to 160 mi/h (72 m/s) in ASCE 7-22. However, the 130 mi/h
12 (58 m/s) contour did not move, and locations inland of the 130 mi/h (58 m/s) contour in ASCE 7-

O
13 22 are lower than those given in ASCE 7-16.
14
n
The wind field model used in the hurricane simulation model was modified where the inflow
io
15 angle was increased by 0.15 radian (8.6 degrees) to be consistent with the inflow angle used for
16 most of the hurricane validation studies that have been performed. The change in the inflow
at

17 angle changes how the rotational winds interact with the hurricane translation speed, resulting in
m

18 a small reduction of wind speeds on the right side of the hurricane and a small increase on the
19 left side of the hurricane. The net effect of the inflow angle change is about a 2% to 3%
or

20 reduction in wind speed for most return periods.


nf

21 The updated model uses historical track and landfall data for the period 1900 to 2018, a 13-year
22 increase of hurricane landfall data, resulting in an increase in the hurricane hazard (defined by
rI

23 landfalling hurricanes ranked by central pressure) along the Texas coast and the Florida
Fo

24 Panhandle, and a decrease along the mid-Gulf coastline.


25 The net effect of all the hurricane model updates yields increases in wind speeds in parts of
26 Texas and the Florida Panhandle, and decreases in wind speeds in parts of Louisiana, along the
27 Mississippi and Alabama coastlines, and in parts of the Northeast.
28 Wind speeds on the Florida peninsula were left unchanged from those given in the 2010 and
29 2016 editions of ASCE 7, making the ASCE 7-22 maps consistent with those used in the Florida
30 Building Code.

17
1 Combination of Non-hurricane and Hurricane Wind Speed Data. Non-hurricane wind
2 speeds and hurricane wind speeds were estimated for return periods ranging from 10 years to
3 100,000 years. The non-hurricane and hurricane winds were then combined as statistically
4 independent events using Equation (C26.5-2), the same general approach that has been used in
5 previous editions of ASCE 7:
6 Pa (v  V )  1  PNH (v  V )PH (v  V ) (C26.5-2)
7 where
8 Pa (v  V ) = Annual exceedance probability for the combined wind hazards,

LY
9 PNH (v  V ) = Annual nonexceedance probability for non-hurricane winds, and

PH (v  V ) = Annual nonexceedance probability for hurricane winds.

N
10

11 The combined winds were interpolated to yield the combined wind hazard curves for MRIs

O
12 associated with each of the wind speed maps. In cases where the hurricane contours are
13
n
unchanged from ASCE 7-10, the shape files from these previous maps were used to ensure
io
14 continuity between the maps.
15 Correlation of Basic Wind Speed Map with the Saffir-Simpson Hurricane Wind Scale.
at

16 Hurricane intensities are reported by the National Hurricane Center (NHC 2015) according to the
m

17 Saffir-Simpson Hurricane Wind Scale, shown in Table C26.5-1. This scale has found broad use
or

18 by hurricane forecasters and local and federal agencies responsible for short-range evacuation of
19 residents during hurricane alerts, as well as long-range disaster planners and the news media. The
nf

20 scale contains five categories of hurricanes and distinguishes them based on wind speed
rI

21 intensity.
22
Fo

23 Table C26.5-1. Saffir-Simpson Hurricane Wind Scale.


Hurricane Sustained Wind Speed, mi/h (m/s)* Types of Damage Due to Hurricane Winds
Category
1 74–95 (33–42) Very dangerous winds will produce some damage
2 96–110 (43–49) Extremely dangerous winds will cause extensive
damage
3 111–129 (50–57) Devastating damage will occur
4 130–156 (58–69) Catastrophic damage will occur
5 ≥157 (70) Highly catastrophic damage will occur
*
24 1 min average wind speed at 33 ft (10 m) above open water.

18
1
2 The wind speeds used in the Saffir-Simpson Hurricane Wind Scale are defined in terms of a
3 sustained wind speed with a 1-min averaging time at 33 ft (10 m) over open water. The ASCE 7
4 standard by comparison uses a 3-s gust speed at 33 ft (10 m) above ground in Exposure C
5 (defined as the basic wind speed, and shown in the wind speed map, Figure 26.5-1). The
6 sustained wind speed over water in Table C26.5-2 cannot be converted to a peak gust wind speed
7 using the Durst curve of Figure C26.5-1, which is only valid for wind blowing over open terrain
8 (Exposure C). An approximate relationship between the wind speeds in ASCE 7 and the Saffir-

LY
9 Simpson scale, based on recent data which indicate that the sea surface roughness remains
10 approximately constant for mean hourly speeds in excess of 67 mi/h (30 m/s), is shown in

N
11 Table C26.5-2. The table provides the sustained wind speeds of the Saffir-Simpson Hurricane
12 Wind Scale over water, equivalent-intensity gust wind speeds over water, and equivalent-

O
13 intensity gust wind speeds over land. For a storm of a given intensity, Table C26.5-2 takes into
14
n
consideration both the reduction in wind speed as the storm moves from over water to over land
io
15 because of changes in surface roughness, and the change in the gust factor as the storm moves
16 from over water to over land (Vickery et al. 2009a, Simiu et al. 2007).
at

17
m

18 Table C26.5-2. Approximate Relationship between Wind Speeds in ASCE 7 and Saffir-
19 Simpson Hurricane Wind Scale.
or

Sustained Wind Speed over Gust Wind Speed over Gust Wind Speed Over
nf

Water a Water b Land c


Saffir-Simpson Hurricane mi/h m/s mi/h m/s mi/h m/s
rI

Category
1 74–95 33–42 90–116 40–51 81–105 36–47
Fo

2 96–110 43–49 117–134 52–59 106–121 48–54


3 111–129 50–57 135–157 60–70 122–142 55–63
4 130–156 58–69 158–190 71–84 143–172 64–76
5 >157 >70 >191 >85 >173 >77
20
a
21 1-min average wind speed at 33 ft (10 m) above open water.
b
22 3-s gust wind speed at 33 ft (10 m) above open water.

19
c
1 3-s gust wind speed at 33 ft (10 m) above open ground in Exposure Category C. This column
2 has the same basis (averaging time, height, and exposure) as the basic wind speed from Figure
3 26.5-1.
4
5 Table C26.5-3 shows the design wind speed from the ASCE 7 basic wind speed maps (Figure
6 26.5-1) for various locations along the hurricane coastline from Maine to Texas, and for Hawaii,
7 Puerto Rico, and the Virgin Islands. Tables C26.5-4 through C26.5-6 show the basic wind speeds
8 for Risk Category II, III, and IV buildings and other structures in terms of the hurricane category

LY
9 equivalents on the Saffir-Simpson Hurricane Wind Scale. These wind speeds represent an
10 approximate limit state; structures designed to withstand the wind loads specified in this

N
11 standard, which are also appropriately constructed and maintained, should have a high
12 probability of surviving hurricanes of the intensities shown in Tables C26.5-4 through C26.5-6

O
13 without serious structural damage from wind pressure alone.
14
n
io
15 Table C26.5-3. Basic Wind Speeds at Selected Coastal Locations in Hurricane-Prone Areas.
Coordinates (decimal Basic Wind Speeds (mi/h)
at
degrees)
Location Latitude Longitude Risk Cat. II Risk Cat. III Risk Cat. IV
m

(700-year) (1,700-year) (3,000-year)


Bar Harbor, Maine 44.3813 68.1968 110 119 124
or

Hampton Beach, New 42.9107 70.8102 Table


116 C26.5-3124
needs to be updated
129
nf

Hampshire
Boston, Massachusetts 42.3578 71.0012 120 129 133
rI

Hyannis, Massachusetts 41.6359 70.2901 132 139 145


Newport, Rhode Island 41.453 71.3058 131 139 144
Fo

New Haven, Connecticut 41.2803 72.9327 121 129 134


Southhampton, New York 40.871 72.3844 130 138 145
Manhattan, New York 40.7005 74.0135 115 127 130
Atlantic City, New Jersey 39.3536 74.4336 126 135 140
Rehoboth Beach, 38.7167 75.0752 121 130 137
Delaware
Ocean City, Maryland 38.3314 75.0835 128 136 140
Virginia Beach, Virginia 36.8306 75.9691 123 132 137

20
Wrightsville Beach, North 34.1973 77.8014 146 156 160
Carolina
Folly Beach, South 32.6496 79.9512 149 158 166
Carolina
Sea Island, Georgia 31.179 81.3472 131 145 150
Jacksonville Beach, 30.2836 81.387 129 140 150
Florida
Melbourne Beach, Florida 28.0684 80.5564 151 162 170
Miami Beach, Florida 25.7643 80.1309 170 183 190
Key West, Florida 24.5477 81.7843 180 200 200

LY
Clearwater, Florida 27.9658 82.8042 146 153 158
Panama City Beach, 30.1558 85.7744 135 146 149
Florida

N
Gulf Shores, Alabama 30.2486 87.6808 160 172 177

88.8887

O
Biloxi, Mississippi 30.3924 160 176 183
Slidell, Louisiana 30.2174 89.824 142 152 158
Cameron, Louisiana 29.7761 93.2921
94.826
n 144 154 158
io
Galveston, Texas 29.2663 150 159 167
Port Aransas, Texas 27.8346 97.0446 150 157 163
at
m
or

1 Notes:
2 1. All wind speeds in Table C26.5-3 are 3-s gust wind speeds at 33 ft (10 m) above ground
nf

3 for Exposure Category C.


rI

4 2. Basic wind speed in hurricane–prone regions can vary significantly over a city or county;
5 the values shown are for randomly selected points along the coast for communities listed
Fo

6 in the tables. Wind speeds at other locations in those communities may be greater or less
7 than the values shown in the table.
8 3. Conversion of mi/h to m/s: mi/h × 0.44704 = m/s.
9

21
Table C26.5-4
needs to be
updated

LY
1

N
2 Table C26.5-4. Basic wind speed for Risk Category II buildings and other structures at selected

O
3 locations in hurricane-prone areas.

n
io
Table C26.5-5
needs to be
at

updated
m
or
nf
rI

4
Fo

5 Table C26.5-5. Basic wind speed for Risk Category III buildings and other structures at selected
6 locations in hurricane-prone areas.
7

22
Table C26.5-6
needs to be
updated

LY
N
1
2 Table C26.5-6. Basic wind speed for Risk Category IV buildings and other structures at selected

O
3 locations in hurricane-prone areas.
4
n
io
5 Tables C26.5-2 through C26.5-6 are intended to help users of the standard better understand
6 design wind speeds as used in this standard in relation to wind speeds reported by weather
at

7 forecasters and the news media, who commonly use the Saffir–Simpson Hurricane Wind Scale.
m

8 The Saffir-Simpson hurricane category equivalent Exposure C gust wind speed values given in
or

9 Tables C26.5-2 through C26.5-6, which are associated with a given sustained wind speed, should
10 be used as a guide only. The gust wind speeds associated with a given sustained wind speed may
nf

11 vary with storm size and intensity, as suggested by Vickery et al. (2009a).
rI

12 Wind Speeds for Long Return Periods. Wind hazard maps for longer return periods than the
13 3,000-year MRI map provided in Figure 26.5-1D are provided in AppendixF, which may be
Fo

14 needed for some performance-based wind designs and other applications.


15 Wind Speeds for Serviceability Design. For serviceability applications such as drift and
16 habitability, the Appendix C commentary presents maps of peak gust wind speeds at 33 ft (10 m)
17 above ground in Exposure C conditions for return periods of 10, 25, 50, and 100 years (Figures
18 CC.2-1 through CC.2-4).
19 The probability, Pn , that the wind speed associated with a certain annual probability, Pa , will be

20 equaled or exceeded at least once during an exposure period of n years is

23
1 Pn  1  (1  Pa )n (C26.5-3)

2 where
3 Pa  1  e( 1/MRI) (C26.5-4)

4 For MRIs of about 10 years or longer, Pa is very closely approximated by the reciprocal of the

5 MRI: Pa  1/ MRI .

6 As an example, if a wind speed is based on Pa = 0.02/year (50-year MRI), the probability that
7 this speed will be equaled or exceeded (at least once) during a 25-year period is 0.40 (i.e., 40%),

LY
8 and the probability of being equaled or exceeded in a 50-year period is 64%. Similarly, if a wind
9 speed is based on Pa  0.00143 (700-year MRI), the probability that this speed will be equaled

N
10 or exceeded during a 25-year period is 3.5%, and the probability of being equaled or exceeded in

O
11 a 50-year period is 6.9%.
12 Some products have been evaluated, and test methods have been developed, based on design
13 n
wind speeds that are consistent with the unfactored load effects typically used in allowable stress
io
14 design (ASD). Table C26.5-7 provides a comparison of the strength design-based wind speeds
at
15 used in the ASCE 7-10 and 7-16 basic wind speed maps and the ASCE 7-05 basic wind speeds
16 used in these product evaluation reports and test methods. A column of values is also provided to
m

17 allow comparison with ASCE 7-93 basic wind speeds.


or

18
19 Table C26.5-7. Basic Wind Speeds: ASCE 7-93 to ASCE 7-16.
nf

ASCE 7-16 and ASCE 7-10 Basic ASCE 7-05 through ASCE 7-95 Basic ASCE 7-93 and Prior Editions Basic Wind
rI

Wind Speed (3-s gust, mi/h) Wind Speed (3-s gust, mi/h) Speed (fastest mile, mi/h)
108 * 85 71
Fo

114 * 90 76
126 100 85
133 105 90
139 110 95
152 120 104
164 130 114
177 140 123
183 145 128
190 150 133
215 170 152
20

24
*
1 In ASCE 7-10 the wind speed values of 108 mi/h and 114 mi/h were rounded to 110 mi/h and
2 115 mi/h, respectively.
3 Note: Conversion of mi/h to m/s: mi/h × 0.44704 = m/s.
4
5 C&C Rating for Building-Envelope Products. Building-envelope products that have been
6 tested to air pressure standards (such as ASTM E330, CSA A123.21, or other standards that
7 incorporate a safety factor) are typically rated for an allowable stress design wind pressure (
8 0.6W ) rather than a strength design pressure ( 1.0W ) or wind speed. To properly select products

LY
9 tested and rated in this manner, the C&C pressures determined from Chapter 30 should be
10 adjusted for the allowable stress design load factor of 0.6W in Section 2.4.1.

N
11 C26.5.2 Special Wind Regions

O
12 Although the wind speed maps of Figure 26.5-1 are valid for most regions of the country, there
13 n
are special regions in which wind speed anomalies are known to exist. Some of these special
io
14 regions are noted in Figure 26.5-1. As site-specific studies within these Special Wind Regions
at
15 are performed, the information is being reviewed by the Committee and incorporated into the
16 ASCE 7 Hazard Tool with each edition of the Standard. In ASCE 7-16, the special wind regions
m

17 were restored to the areas originally designated in ANSI A58.1-1982 through ASCE 7-93. The
or

18 regions had inadvertently been changed in ASCE 7-95 because of a graphical error. The special
19 wind regions around the Great Lakes in the northeast United States and in the Puget Sound area
nf

20 near Seattle were omitted intentionally from ASCE 7-95 because of lack of meteorological data
rI

21 demonstrating the existence of wind speeds higher than could be explained by exposure. In these
22 special regions, winds blowing over mountain ranges or through gorges or river valleys can
Fo

23 develop speeds that are substantially higher than the values indicated on the map. When selecting
24 basic wind speeds in these special regions, use of regional climatic data and consultation with a
25 wind engineer or meteorologist is advised. For the special wind regions in northern Colorado and
26 Kern County, California, this analysis has been performed (Peterka 2006; Banks et al., 2019),
27 and recommended wind speeds are available from the ASCE 7 Wind Design Geodatabase, which
28 is available through the ASCE Hazard Tool.
29 It is also possible that anomalies in wind speeds exist on a micrometeorological scale. For
30 example, wind speed-up over hills and escarpments is addressed in Section 26.8. Wind speeds

25
1 over complex terrain may be better determined by wind tunnel studies, as described in
2 Chapter 31. Adjustments of wind speeds should be made at the micrometeorological scale on the
3 basis of wind engineering or meteorological advice and used in accordance with the provisions
4 of Section 26.5.3 when such adjustments are warranted. Because of the complexity of
5 mountainous terrain and valley gorges in Hawaii, there are topographic wind speed-up effects
6 that cannot be addressed solely by Figure 26.8-1 (ARA 2001). In the Hawaii special wind region,
7 research and analysis have established that there are special K zt topographic-effect adjustments

8 (Chock et al. 2005).

LY
9 The southernmost special wind region in California experiences Santa Ana winds (dry, mountain
10 downslope winds). The appropriate boundaries of this region are difficult to quantify because of

N
11 a lack of data (Searer et. al. 2010). The current southern boundary is based on analysis of

O
12 regional climatic data (SEAOC 2017). Similar work was undertaken in Kern County, California
13 (Kern County 2017), and northern Colorado (Peterka 2010). In each of these three cases, the
14
n
boundaries of the special wind regions changed from the original ANSI A58.1-1982 boundaries,
io
15 and it is likely that this will be the case for many special wind regions if they are examined
at
16 closely. Given the decrease in wind speeds for the western United States between ASCE 7-10
17 and ASCE 7-16, it is also possible that there are additional regions where higher localized wind
m

18 speeds exist that have not been identified as special wind regions. For example, where special
or

19 wind regions stop at county or state lines, this indicates where the study stopped, or data quality
20 was poor, rather than where the wind speeds are expected to suddenly decrease.
nf

21 C26.5.3 Estimation of Basic Wind Speeds from Regional Climatic Data


rI
Fo

22 When using regional climatic data in accordance with the provisions of Section 26.5.3 and in lieu
23 of the basic wind speeds given in Figure 26.5-1, the user is cautioned that the gust factors,
24 velocity pressure exposure coefficients, gust-effect factors, pressure coefficients, and force
25 coefficients of this standard are intended for use with the 3-s gust speed at 33 ft (10 m) above
26 ground in terrain with open exposure. It is therefore necessary to correct the data set, where
27 required, for anemometer height, upwind terrain, local topography, and/or averaging time. Care
28 should be taken to account for any historical changes in these factors within the data set and to
29 apply the appropriate corrections, noting that anemometer locations or heights may have changed
30 over time, as may the surroundings.
26
1 The results of statistical studies of wind speed records, reported by Durst (1960) for extratropical
2 winds and by Vickery et al. (2000) for hurricanes, are given in Figure C26.5-1, which defines the

3 relation between wind speed averaged over time t in seconds, Vt , and the hourly wind speed,

4 V3600 . The hurricane simulation model described in Section C26.5.1 uses the gust-factor curve
5 from ESDU (1982, 1993), which has been shown to be valid for hurricane winds (Vickery and
6 Skerlj 2005). Similar conclusions regarding hurricane gust factors were drawn by Jung and
7 Masters (2013). The relation between wind speed averaging times for extratropical winds in any
8 terrain exposure is given in Section 11.2.4.2 of Simiu (2011).

LY
9 When local climatic records are used, the records should be screened to remove any obviously
10 erroneous data. If multiple locations are available, then the data from each should be compared

N
11 for consistency. In areas with mixed wind climates, where multiple storm types contribute to the

O
12 extreme wind climate, storm separation should be conducted on the data sets and individual
13 extreme value fits applied to each storm type before combining to determine the overall risk.
14 n
In using local data, it should be emphasized that sampling errors can lead to large uncertainties in
io
15 specification of wind speed. Sampling errors are the errors associated with the limited size of the
at

16 climatological data samples (e.g., years of record of extreme speeds or number of storms). When
m

17 short local records are used to estimate extreme wind speeds, care and conservatism should be
18 exercised. Different methods, such as periodic independent maxima or the Method of
or

19 Independent Storms, may be used to increase the effective size of the data set in comparison with
nf

20 just the use of annual maxima, as may the use of superstations where multiple data sets are
21 combined.
rI

22 An examination of statistical sampling error can be used to ensure the reliability of the statistical
Fo

23 fits. It is recommended that design wind speeds for use in strength design be based on the best-
24 estimate wind speed plus a minimum of two to three times the standard deviation of the sampling
25 error (Simiu and Scanlan 1996), equating to 95% and 99% confidence intervals, respectively.
26 This is particularly important if justifying a wind speed lower than the basic wind speed from
27 Figure 26.5-1. The above analyses should only be conducted by professionals experienced in
28 extreme wind climate analysis for wind loading applications. Where values significantly below
29 those from Figure 26.5-1 are being proposed, an independent peer review by a suitably
30 experienced professional may be advisable.

27
1 C26.6 Wind Directionality

2 The wind load factor 1.3 in ASCE 7-95 included a “wind directionality factor” with a nominal
3 value of 0.85 (Ellingwood 1981; Ellingwood et al. 1982). This factor accounts for two effects:
4 (1) The reduced probability of maximum winds coming from any given direction, and (2) the
5 reduced probability of the maximum pressure coefficient occurring for any given wind direction.

6 The nominal wind directionality factor (denoted by Kd in the standard) is tabulated in


7 Table 26.6-1 for different structure types. As new research becomes available, this factor can be

LY
8 directly modified. Nominal values for the factor were established from references in the literature
9 and collective committee judgment. A value of 0.85 might be more appropriate if a triangular

N
10 trussed frame is shrouded in a round cover. A value of 0.95 might be more appropriate for a
11 round structure that has a nonaxisymmetrical lateral load resistance system.

O
12 C26.7 Exposure
n
io
13 The descriptions of the surface roughness categories and exposure categories in Section 26.7
at
14 have been expressed as far as possible in easily understood verbal terms that are sufficiently
15 precise for most practical applications. Upwind surface roughness conditions required for
m

16 Exposures B and D are shown schematically in Figures C26.7-1 and C26.7-2, respectively.
or

17 Aerial photographs showing examples of Exposures B, C, and D are shown in Figures C26.7-5
18 through C26.7-7. For cases where the designer wishes to make a more detailed assessment of the
nf

19 surface roughness category and exposure category, the following more mathematical description
rI

20 is offered for guidance (Irwin 2006). The ground surface roughness is best measured in terms of

a roughness length parameter called z0 . Each of the surface roughness categories B through D
Fo

21

22 corresponds to a range of values of this parameter, as does the even rougher category A used in
23 older versions of the standard in heavily built-up urban areas but removed in the recent editions.

24 The range of z0 for each terrain category is given in Table C26.7-1. Exposure A has been
25 included in the table as a reference that may be useful when using the Wind Tunnel Procedure.

26 Further information on values of z0 in different types of terrain can be found in Simiu and Yeo
27 (2019) and in Table C26.7-2, based on Davenport et al. (2000) and Wieringa et al. (2001).
28
28
1
2 Figure C26.7-1. Upwind surface roughness conditions required for Exposure B.
3

LY
N
O
n
io
at
m

4
or

5 Figure C26.7-2. Upwind surface roughness conditions required for Exposure D, for the cases
6 with (a) Surface Roughness D immediately upwind of the building, and (b) Surface Roughness B
nf

7 and/or C immediately upwind of the building.


rI

8
z0 by Exposure Category.
Fo

9 Table C26.7-1. Range of

Lower Limit of Typical Value of Upper Limit of


Exposure
z0 Inherent in Tabulated Kz Values
Category
z0 , ft (m) z0 , ft (m) z0 , ft (m) in Table 26.10-2, ft (m)
A 2.3 (0.7) ≤ z0 6.6 (2) —
B 0.5 (0.15) ≤ z0 1.0 (0.3) z0 < 2.3 (0.7) 1.0 (0.3)
C 0.033 (0.01) ≤ z0 0.1 (0.03) z0 < 0.5 (0.15) 0.1 (0.03)
D — 0.016 (0.005) z0 < 0.033 (0.01) 0.016 (0.005)
10
11 Table C26.7-2. Davenport Classification of Effective Terrain Roughness.

29
Class αb zg’, ft zd (ft or m)c Wind Flow and Landscape Descriptiond
z0 , ft
(m)b
(m)a
1 0.0007 12.9 509
zd  0 Sea: Open sea or lake (irrespective of wave size), tidal flat,
(0.0002) (155) snow-covered flat plain, featureless desert, tarmac, and concrete,
with a free fetch of several kilometers.
2 0.016
(0.005)
11.4 760
(232)
zd  0 Smooth: Featureless land surface without any noticeable
obstacles and with negligible vegetation, e.g., beaches, pack ice
without large ridges, marsh, and snow-covered or fallow open
country.
3 0.1 (0.03) 9.0 952
(290)
zd  0 Open: Level country with low vegetation (e.g., grass) and
isolated obstacles with separations of at least 50 obstacle heights,

LY
e.g., grazing land without windbreaks, heather, moor, and tundra,
runway area of airports. Ice with ridges across-wind.
4 0.33 (0.10) 7.7 1,107
zd  0 Roughly open: Cultivated or natural area with low crops or plant

N
(337) covers, or moderately open country with occasional obstacles
(e.g., low hedges, isolated low buildings, or trees) at relative

O
horizontal distances of at least 20 obstacle heights.
5 0.82 (0.25) 6.8 1,241
(378)
zd  0.2zH Rough: Cultivated or natural area with high crops or crops of
varying height and scattered obstacles at relative distances of 12
n
to 15 obstacle heights for porous objects (e.g., shelterbelts) or 8
io
to 12 obstacle heights for low solid objects (e.g., buildings).
6 1.64 (0.5) 6.2 1,354
zd 0.5zH Very rough: Intensely cultivated landscape with many rather
at
(413) large obstacle groups (large farms, clumps of forest) separated by
open spaces of about 8 obstacle heights. Low, densely planted
m

major vegetation like bushland, orchards, young forest. Also,


area moderately covered by low buildings with interspaces of 3
to 7 building heights and no high trees.
or

7 3.3 (1.0) 5.7 1,476


(450)
zd  0.7zH Skimming: Landscape regularly covered with similar-size large
obstacles, with open spaces of the same order of magnitude as
nf

obstacle heights, e.g., mature regular forests, densely built-up


area without much building height variation.
rI

8 ≥matu 5.2 1,610 Analysis by Chaotic: City centers with mixture of low-rise and high-rise
(≥mat) (490) wind tunnel buildings, or large forests of irregular height with many
Fo

advised clearings. (Analysis by wind tunnel advised.)

1 a
The surface roughness length, z0 , represents the physical effect that roughness objects
2 (obstacles to wind flow) on the Earth’s surface have on the shape of the atmospheric boundary-
3 layer wind velocity profile as determined by the logarithmic law and used in the ESDU model.
b
4 The power law uses the exponent α and is valid up to the height at which geostrophic wind flow

5 begins to occur (see Section C26.10.1). The values provided in this table are based on z0 values
6 originally proposed by Davenport (1960), and the equations (α = c1z0-0.133 and z′g = c2z00.125) can
7 be used to calculate α and historial gradient height z′g, respectively, where c1 = 6.62 and c2 =
30
1 1,273 when the units of z0 and z′g are ft, and c1 = 5.65 and c2 = 450 when the units of z0 and z′g
2 are m, as described in the commentary of ASCE 7-16.

3 c
The zero plane displacement height, zd , is the elevation above ground to which the base of the
4 logarithmic law (and power law) wind profile must be elevated to accurately depict the

5 boundary-layer wind flow. Below zd and less than some fraction of the typical height, zH, of
6 obstacles causing roughness, the near-ground wind flow is characterized as a turbulent exchange
7 with the boundary-layer wind flow above, resulting in significant shielding effects under uniform

LY
8 to moderately uniform roughness conditions (e.g., Classes 5 through 7 in this table). In this

9 condition, the effective mean roof height, heff , may then be determined as h – zd [but not less

N
10 than 15 ft (4.6 m)] for the purpose of determining MWFRS wind loads acting on a building

O
11 structure located within such a roughness class. Appropriate values of zd for a given site may
12 vary widely, and those shown in this table should be used with professional judgment. Because
n
io
13 of the presence of highly turbulent flow at elevations near or below zd (except perhaps structures
at
14 embedded in uniform Class 7 roughness), use of an effective mean roof height should not be
15 applied for the determination of C&C wind loads. In Class 8 roughness, where wind flow
m

16 disruptions can be highly nonuniform, channeling effects and otherwise “chaotic” wind flow
or

17 patterns can develop between and below the height of obstacles to wind flow. For this reason, a
18 wind tunnel study is generally advised.
nf

d
19 Use of these wind flow and landscape descriptions should result in no greater than one
rI

20 roughness class error, corresponding to a maximum +/− of these wind qh .


Fo

21 The roughness classifications in Table C26.7-2 are not intended to replace the use of exposure
22 categories as required in the standard for structural design purposes. However, the terrain
23 roughness classifications in Table C26.7-2 may be related to exposure categories by comparing

24 z0 values between Tables C26.7-1 and C26.7-2. For example, the z0 values for Classes 3 and 4

25 in Table C26.7-2 fall within the range of z0 values for Exposure C in Table C26.7-1. Similarly,

26 the z0 values for Classes 5 and 6 in Table C26.7-2 fall within the range of z0 values for
27 Exposure B in Table C26.7-1.

31
1
2 Research described by Powell et al. (2003), Donelan et al. (2004), and Vickery et al. (2009a)
3 showed that the drag coefficient over the ocean in high winds in hurricanes does not continue to
4 increase with increasing wind speed, as was previously believed (e.g., Powell 1980). These
5 studies showed that the sea surface drag coefficient, and hence the aerodynamic roughness of the
6 ocean, reached a maximum at mean wind speeds of about 67 mi/h (30 m/s). There is some
7 evidence that the drag coefficient actually decreases (i.e., the sea surface becomes
8 aerodynamically smoother) as the wind speed increases further (Powell et al. 2003) or as the

LY
9 hurricane radius decreases (Vickery et al. 2009a). The consequences of these studies are that the
10 surface roughness over the ocean in a hurricane is consistent with that of Exposure D rather than

N
11 Exposure C. Consequently, the use of Exposure D along the hurricane coastline is required.
12

O
13 For Exposure B, previous editions of ASCE 7, beginning with 7-02, used tabulated values of Kz

14 corresponding to n
z0  0.66 ft (0.2 m), which is below the typical value of 1 ft (0.3 m), whereas
io
for Exposures C and D they have always corresponded to the typical values of z0 . The reason for
at
15

16 the difference in Exposure B is that this category of terrain, which is applicable to suburban
m

17 areas, often contains open patches, such as highways, parking lots, and playing fields. These
or

18 open patches cause local increases in the wind speeds at their edges. By using an exposure

coefficient corresponding to a lower-than-typical value of z0 , some allowance is made for this.


nf

19

20 The alternative would be to introduce a number of exceptions to the use of Exposure B in


rI

21 suburban areas, which would add an undesirable level of complexity. ASCE 7-22 uses z0 = 1.0 ft
Fo

22 (0.3 m) for Exposure B, corresponding to the typical value of z0. Although resulting Kz values for
23 Exposure B are generally reduced by using the typical value of z0, the updated ASCE 7-22 power
24 law equation results in Kz values comparable to those determined in earlier editions of ASCE 7
25 up to 60 ft (18.3 m). Thus, the design intent of considering open patches applicable to suburban
26 areas in Exposure B has been maintained for the majority of low-rise building heights where
27 these terrain features will have the largest impact on wind speed. For taller buildings and other
28 structures taller than 60 ft (18.3 m), the influence of suburban open patches on wind speed will

32
1 decrease with height, which is reflected in corresponding values of Kz based on the typical value
2 of z0 and the updated ASCE 7-22 power law.
3 For Exposure C, ASCE 7-22 updated the typical value of z0 from 0.066 ft (0.02 m) to 1 ft
4 (0.3 m), to maintain consistency with the value of z0 used in the hazard maps for both hurricane
5 and non-hurricane winds. The tabulated values of Kz also correspond to the updated typical value
6 of z0.
7

8 The value of z0 for a particular terrain can be estimated from the typical dimensions of surface

LY
9 roughness elements and their spacing on the ground area using an empirical relationship found
10 by Lettau (1969):

N
Sob
11 z0  0.5Hob (C26.7-1)

O
Aob
12 where

13 Hob n
= Average height of the roughness in the upwind terrain,
io
Sob = Average vertical frontal area per obstruction presented to the wind, and
at
14

Aob
m

15 = Average area of ground occupied by each obstruction, including the open area

16 surrounding it.
or

17 Vertical frontal area is defined as the area of the projection of the obstruction onto a vertical
nf

18 plane normal to the wind direction. The area Sob may be estimated by summing the approximate
rI

19 vertical frontal areas of all obstructions within a selected area of upwind fetch and dividing the
Fo

20 sum by the number of obstructions in the area. The average height, Hob , may be estimated in a
21 similar way by averaging the individual heights rather than using the frontal areas. Likewise,

22 Aob may be estimated by dividing the size of the selected area of upwind fetch by the number of
23 obstructions in it.
24
25 As an example, if the upwind fetch consists primarily of single-family homes with typical height

26 Hob  20 ft (6 m), vertical frontal area (including some trees on each lot) of 1,000 ft2 (100 m2),

33
1 and ground area per home of 10,000 ft2 (1,000 m2), then z0 is calculated to be 0.5 × 20 ×
2 1,000/10,000 ft = 1 ft (0.3 m), which falls into Exposure Category B, according to Table C26.7-
3 1.
4
5 Trees and bushes are porous and are deformed by strong winds, which reduce their effective
6 frontal areas (ESDU 1993). For conifers and other evergreens, no more than 50%, and for
7 deciduous trees and bushes, no more than 15%, of their gross frontal area can be taken to be
8 effective in obstructing the wind. Gross frontal area is defined in this context as the projection

LY
9 onto a vertical plane (normal to the wind) of the area enclosed by the envelope of the tree or
10 bush.

N
11

O
12 Ho (1992) estimated that the majority of buildings (perhaps as much as 60% to 80%) have an
13 exposure category corresponding to Exposure B. While the relatively simple definition in the
14 n
standard normally suffices for most practical applications, the designer is often in need of
io
15 additional information, particularly with regard to the effect of large openings or clearings
at
16 (e.g., large parking lots, freeways, or tree clearings) in the otherwise “normal” ground Surface
17 Roughness B. The following is offered as guidance for these situations.
m

18
or

19 1. The simple definition of Exposure B given in the body of the standard, using the surface
20 roughness category definition, is shown pictorially in Figure C26.7-1. This definition
nf

21 applies for the Surface Roughness B condition prevailing 2,630 ft (800 m) upwind with
rI

22 insufficient “open patches,” as defined in the following procedure to disqualify the use of
23 Exposure B. This procedure on the net effect of these open patches applies where the
Fo

24 prevailing exposure beyond 2,600 ft (792 m) is Exposure B.


25 2. An open area in the Surface Roughness B large enough to have a significant effect on the
26 exposure category determination is defined as an “open patch.” To be considered an open
27 patch, an open area meets the following:
28 a. Open areas should be greater than the minimum areas given by Figure C26.7-4.
29 Interpolation shall be used between the reference distances of 500 ft, 1,500 ft, and
30 2,600 ft (152 m, 457 m, and 790 m) to determine the intermediate minimum open
31 patch area criteria.

34
1 b. The open area shall have minimum dimensions given by conditions i, ii, or iii
2 below and have length-to-width ratios between 0.5 and 2.0.
3 i. Within 500 ft (152 m) of the building or structure, an open area greater than
4 or equal to approximately 164 ft (50 m) in length or width.
5 ii. At 1,500 ft (457 m) upwind from the building or structure, an open area
6 greater than or equal to approximately 328 ft (100 m) in length or width.
7 iii. At 2,600 ft (790 m) upwind from the building or structure, an open area
8 greater than or equal to approximately 500 ft (152 m) in length or width.

LY
9 3. Open patches separated by less than the along-wind dimension of the larger patch shall be
10 treated as equivalent to a single open patch with length equal to the sum of the individual

N
11 patch along-wind dimensions and width determined to provide an area equal to the sum
12 of the individual open patch areas.

O
13 4. A circular sector is an area defined by a limiting radius from the center and an arc, in this
14
n
case, 45 degrees, per Section 26.7.4. If the proportion of open patch within any of the
io
15 sectors defined by the three radii above is less than 25% of the sector area, the sector is
16 considered to meet the requirements for Exposure Category B. Where the proportion of
at

17 open patches within any 45-degree sector within any of the three radii of 500 ft (152 m),
m

18 1,500 ft (457 m), or the greater of 2,600 ft (790 m) or 20 times the height of the building

Kz
or

19 or structure exceeds 25% of the sector area but is not greater than 50%, the values of

20 are taken as the average of the Exposure B and C values within 100 ft (31 m) height
nf

21 above grade. Above 100 ft (31 m), Exposure B values of Kz shall still apply. Where the
rI

22 proportion of open patches within any of the sectors defined by the three radii of the
Fo

23 building or structure exceeds 50%, the values of Kz shall be based on Exposure C.


24 5. The procedure for evaluation of the net effect of open patches of Surface Roughness C or
25 D on the use of Exposure Category B is shown pictorially in Figures C26.7-3 and C26.7-
26 4. Note that the plan location of any open patch may have a different effect for different
27 wind directions.
28
29 This above procedure is a simplification derived from a boundary-layer model, and therefore
30 more exact results for the velocity profile may be achieved through direct use of an accepted

35
1 boundary-layer model that is capable of addressing the effects of open areas within a regime
2 defined by the surface roughness parameters given in Table C26.7-2.
3
4 Aerial photographs, representative of each exposure type, are included in Figures C26.7-5 to
5 C26.7-7 to aid the user in establishing the proper exposure for a given site. Obviously, the proper
6 assessment of the exposure is a matter of good engineering judgment. This fact is particularly
7 true in light of the possibility that the exposure could change in one or more wind directions due
8 to future demolition and/or development.

LY
9

N
O
n
io
at
m
or

10
11 Figure C26.7-3. Sector analysis for Exposure B with upwind open patches.
nf

12 Notes:
rI

13 1. For each selected wind direction at which the wind loads are to be determined, the
Fo

14 exposure of the building or structure shall be determined for the two upwind sectors
15 extending 45 degrees to either side of the selected wind direction.
16 2. Consider open patches of sizes equal to or greater than the areas given in Figure C26.7-4
17 per Commentary Section C26.7.
18 3. Determine the proportion of open patches in any 45-degree sector within radii of 500 ft
19 (152 m), 1,500 ft (457 m), or the greater of 2,600 ft (790 m) or 20 times the height of the
20 structure.
21 4. If the proportion of open patch within any of the three radii is less than 25% of the sector
22 area, the sector is considered to meet the requirements for Exposure B. Where the
36
1 proportion within any of the three radii exceeds 25% of the sector area but is not greater

2 than 50%, the values of Kz are taken as the average of the Exposure B and C values
3 within 100 ft (31 m) height above grade. Above 100 ft (31 m), Exposure B values shall
4 still apply. Where the proportion of open patches within any of the three radii of the

5 structure exceeds 50%, the values of Kz shall be based on Exposure C.


6 5. Apply the exposure requirements of Section 26.7.4 once the directional exposures are
7 determined for each sector. See Commentary Section C26.7.4.

LY
8

N
O
n
io
at
m
or
nf

9
rI

10 Figure C26.7-4. Minimum area of individual open patches affecting qualification of Exposure B.
11
Fo

12 C26.7.4 Exposure Requirements

13 Section 26.5.1 of the standard requires that a structure be designed for winds from all directions.
14 A rational procedure to determine directional wind loads is described here. Wind load for
15 buildings using Section 27.3.1 and Figures 27.3-1, 27.3-2, or 27.3-3 are determined for eight
16 wind directions at 45-degree intervals, with four falling along primary building axes, as shown in
17 Figure C26.7-8. For each of the eight directions, upwind exposure is determined for each of two
18 45-degree sectors, one on each side of the wind direction axis. The sector with the exposure
37
1 giving the highest loads is used to define wind loads for that direction. For example, for winds
2 from the north, the exposure from Sector 1 or 8, whichever gives the higher load, is used. For
3 wind from the east, the exposure from Sector 2 or 3, whichever gives the highest load, is used.
4 For wind coming from the northeast, the more exposed of Sectors 1 or 2 is used to determine full
5 x and y loading individually, and then 75% of these loads is to be applied in each direction at the
6 same time, according to the requirements of Section 27.3.5 and Figure 27.3-8. The procedure
7 defined in this section for determining wind loads in each design direction is not to be confused

8 with the determination of the wind directionality factor, Kd. The Kd factor determined from

LY
9 Section 26.6 and Table 26.6-1 applies for all design wind directions. See Section C26.6.
10

N
11 C&C loads for all buildings and MWFRS loads for low-rise buildings are determined using the

O
12 upwind exposure for the single surface roughness in one of the eight sectors of Figure C26.7-8
13 that gives the highest wind loads.
n
io
14 C26.8 TOPOGRAPHIC EFFECTS
at

15 This section specifies when topographic effects need to be applied to a particular structure
m

16 (Means et al. 1996). In some wind loading guidelines, hills of sufficient size have been assumed
17 to eliminate topographic effects for up to 2 miles downwind, Research by Almeida et al. (1993)
or

18 indicates that very little sheltering is expected from nearby (3H upwind) hills of comparable size.
nf

19 Some sheltering is expected downwind of significantly larger hills, but evidence from studies or
20 literature is needed to quantify this.
rI

21
Fo

22 Condition 3 specifies a lower height, H , for consideration of topographic effects in Exposure C


23 and D than for Exposure B (Means et al. 1996), and observation of actual wind damage has
24 shown that the affected height, H , is less in Exposures C and D.
25
26 Buildings sited on the upper half of an isolated hill or escarpment may experience significantly
27 higher wind speeds than buildings on level ground. The topographic feature (2D ridge or

28 escarpment, or 3D axisymmetrical hill) is described by two parameters, H and Lh. H is the height

38
1 of the hill or difference in elevation between the crest and that of the upwind terrain. Lhis the
2 distance upwind of the crest to where the ground elevation is equal to half the height of the hill.

3 K zt is determined from three multipliers, K1, K2, and K3, which are obtained from Figure 26.8-1.

4 K1is related to the shape of the topographic feature; the maximum speed-up near the crest, K2,

5 accounts for the reduction in speed-up with distance upwind or downwind of the crest; and K3
6 accounts for the reduction in speed-up with height above the local ground surface.

LY
7
8

N
O
n
io
at
m

9
or

10 Figure C26.7-5(a). Exposure B: suburban residential area with mostly single-family dwellings.
11 Low-rise structures, less than 30 ft (9.1 m) high, in the center of the photograph have sites
nf

12 designated as Exposure B with Surface Roughness Category B terrain around the site for a
rI

13 distance greater than 1,500 ft (457 m) in any wind direction.


14
Fo

39
LY
1

N
2 Figure C26.7-5(b). Exposure B: urban area with numerous closely spaced obstructions having
3 the size of single-family dwellings or larger. For all structures shown, terrain representative of

O
4 Surface Roughness Category B extends more than 20 times the height of the structure or 2,600 ft
5 (792 m), whichever is greater, in the upwind direction.
n
io
6
at
m
or
nf
rI
Fo

7
8 Figure C26.7-5(c). Exposure B: buildings and other structures in the foreground are located in
9 Exposure B. Buildings and other structures in the center top of the photograph adjacent to the
10 clearing to the left, which meets the “open patch” criteria defined in Section C26.7, are located in
11 Exposure C when wind comes from the left over the clearing.
12
13 The multipliers listed in Figure 26.8-1 are based on the assumption that the wind approaches the
14 hill along the direction of maximum slope, causing the greatest speed-up near the crest. The

40
1 average maximum upwind slope of the hill is approximately H/2Lh , and measurements have

2 shown that hills with slopes of less than about 0.10 ( H / Lh  0.20) are unlikely to produce

3 significant speed-up of the wind. For values of H / Lh  0.5, the speed-up effect is assumed to be
4 independent of slope. The speed-up principally affects the mean wind speed rather than the

5 amplitude of the turbulent fluctuations, and this fact has been accounted for in the values of K1,

6 K2, and K3 given in Figure 26.8-1. Therefore, values of K zt obtained from Figure 26.8-1 are

LY
7 intended for use with velocity pressure exposure coefficients, Kh and Kz, which are based on gust

N
8 speeds.
9

O
n
io
at
m
or

10
11 Figure C26.7-6(a). Exposure C: flat open grassland with scattered obstructions having heights
nf

12 generally less than 30 ft (9.1 m).


rI

13
Fo

14
15 Figure C26.7-6(b). Exposure C: open terrain with scattered obstructions having heights generally
16 less than 30 ft (9.1 m). For most wind directions, all structures with a mean roof height less than

41
1 30 ft (9.1 m) in the photograph are less than 1,500 ft (457 m) from an open field, which prevents
2 the use of Exposure B.
3

LY
4

N
5 Figure C26.7-7. Exposure D: a building at the shoreline with wind flowing over open water for a

O
6 distance of at least 5,000 ft (1,524 m) or 20 times the building or structure height, whichever is
7 greater.
8 n
io
9 It is not the intent of Section 26.8 to address the general case of wind flow over hilly or complex
at

10 terrain, for which engineering judgment, expert advice, or the Wind Tunnel Procedure as
m

11 described in Chapter 31 may be required. Background material on topographic speed-up effects


12 may be found in the literature (Jackson and Hunt 1975, Lemelin et al. 1988, Walmsley et al.
or

13 1986).
nf

14
rI
Fo

15
16 Figure C26.7-8. Determination of wind loads from different directions.
17

42
1 The designer is cautioned that, at present, the standard contains no provision for vertical wind
2 speed-up because of a topographic effect, even though this phenomenon is known to exist and
3 can cause additional uplift on roofs. Additional research is required to quantify this effect before
4 it can be incorporated into the standard.
5 As illustrated in the example (in US Customary Units) that follows, there is a minimal difference
6 in the calculated Kzt for Exposure B, C, or D. For this reason, the table of topographic multipliers
7 shall be used for any exposure unless there is a need for increased accuracy in the topographic
8 multiplier.

LY
9 An example using the topographic provisions follows for Exposures B, C, and D. Using the
10 graphic of a hill, escarpment, or 3D escarpment shown in Figure 26.8-1, the topographic

N
11 parameters are
12 H = 500 ft,

O
13 Lh = 1,500 ft,
14 x = 300 ft, and
n
io
15 z = 50 ft.
16
at

17 And then evaluation downwind of the crest results in the following:


m

Feature Exp. H (ft.) Lh (ft) H/Lh K1 x (ft) µ (dn) K2 γ z (ft) K3 Kzt


2D escarp. B 500 1500 0.33 0.25 300 4 0.95 2.5 50 0.92 1.48
or

2D escarp. C 500 1500 0.33 0.28 300 4 0.95 2.5 50 0.92 1.56
2D escarp. D 500 1500 0.33 0.32 300 4 0.95 2.5 50 0.92 1.63
nf

3D axi. B 500 1500 0.33 0.32 300 1.5 0.87 4 50 0.88 1.54
3D axi. C 500 1500 0.33 0.35 300 1.5 0.87 4 50 0.88 1.60
rI

3D axi. D 500 1500 0.33 0.38 300 1.5 0.87 4 50 0.88 1.67
2D ridge B 500 1500 0.33 0.43 300 1.5 0.87 3 50 0.90 1.80
Fo

2D ridge C 500 1500 0.33 0.48 300 1.5 0.87 3 50 0.90 1.90
2D ridge D 500 1500 0.33 0.52 300 1.5 0.87 3 50 0.90 1.97
18

19 C26.9 GROUND ELEVATION FACTOR

20 The ratio of air pressure and density at elevation z relative to the standard values at z  0 , with
21 constant temperature, is given by the barometric formula

22 pz / p0  ρz /ρ0  egz/RT

43
1 where
2 g = Acceleration of gravity = 32.174 ft/s2 (9.807 m/s2),

3 R = Gas constant of air = 1,718 lb-ft/slug/°R (287 N-m/kg/K)), and


4 T = Absolute temperature = 518 °R (288 K).
5 With these values and elevation z (elevation of ground above sea level), the ratio is determined

6 from the formulas given in Table 26.9-1, where Ke  ρz /ρ0 . For reference, a more complete
7 version of Table 26.9-1 including air density values is provided in Table C26.9-1.
8

LY
9 Table C 26.9-1. Gound Elevation Factor including Air Density.
Ground Elevation, Air Density, ρ
Ke

N
𝒛𝒆 Ratio,

ft m Slug/ft3 kg/m3

O
ρ/ ρ0
–10,000 –305 0.000247 1.269 1.04
0 0 0.000238 1.224 1.00
n
io
1,000 305 0.000229 1.180 0.96
2,000 610 0.000221 1.138 0.93
at
3,000 914 0.000213 1.098 0.90
4,000 1,219 0.000206 1.059 0.86
m

5,000 1,524 0.000198 1.021 0.83


6,000 1,829 0.000191 0.985 0.80
or

7,000 2,134 0.000185 0.950 0.78


8,000 2,438 0.000178 0.916 0.75
nf

9,000 2,743 0.000172 0.883 0.72


rI

10,000 3,048 0.000166 0.852 0.70


10
Fo

11 Ke 1.0 is permitted in all cases. While this is somewhat unconservative for elevations below
12 sea level, the committee believes it is reasonable to permit this since the effect is very small for
13 all areas below sea level in the United States (0 to −300 ft (-91 m) in Death Valley, for a
14 maximum of 1% increase in air density), and is likely to be reduced even more because of higher
15 average temperatures.

16 C26.10 Velocity Pressure

44
1 C26.10.1 Velocity Pressure Exposure Coefficient

2 The velocity pressure exposure coefficient Kz can be obtained using these equations:
/∝
3 𝐾 2.41 for z < 15 ft (C26.10-1)

/∝
.
4 𝐾 2.41 for z < 4.6 m (C26.10-1.SI)

/∝
5 𝐾 2.41 for 15 ft (4.6 m) ≤ z ≤ zg (C26.10-2)

LY
6 𝐾 2.41 for zg < z ≤ 3,280 ft (1,000 m) (C26.10-3)

7 in which values of α and zg are given in Table 26.11-1. These formulas are also given in

N
8 Table 26.10-1 to aid the user.

O
9 Previous editions of the standard, from ASCE 7-95 to ASCE 7-16, effectively established a
10
n
maximum value for Kz of 2.01 by limiting the height z to the nominal height of the atmospheric
io
11 boundary layer, also known as gradient height zg. The boundary-layer height was based on
12 research of the 1950s and 1960s (e.g., Davenport 1960). Since the 1970s, studies by Harris and
at

13 Deaves (1981) and Zilitinkevich and Esau (2002), have shown that atmospheric boundary-layer
m

14 heights in large synoptic storms are much greater (e.g., 6,700 ft to 13,000 ft (2 km to 4 km)) than
15 the zg values from previous editions of ASCE 7, and wind speeds in such storms are known to
or

16 increase above the previous values of zg (Irwin 2006; Simiu et al. 2016). Recent studies of
nf

17 tropical cyclones (Tse et al. 2013) have also shown boundary-layer heights greater than the zg
18 values in ASCE 7-16 and previous editions. As the power-law equation in use from ASCE 7-95
rI

19 to 7-16 deviates significantly at high elevations from wind speed profiles based on atmospheric
Fo

20 boundary-layer theory and measurements (Panofsky and Dutton 1984), and as wind speed
21 profiles at elevations of up to 3,280 ft (1,000 m) are known to be reasonably accurate (Irwin
22 2006), the value of zg in ASCE 7-22 was increased to 3,280 ft (1,000 m), 2,460 ft (750 m), and
23 1,935 ft (590 m) for Exposures B, C, and D, respectively. The power law exponents were
24 recalibrated accordingly using more recent wind speed profile models (Harris and Deaves 1981,
25 Kelly et al. 2019). At height zg the value of Kz is limited to a maximum of 2.41 for the exposure
26 being considered. Above this height, and up to 3,280 ft (1,000 m), Kz is assumed to remain
27 constant; however, for buildings and other structures with height greater than zg, or for types of

45
1 wind that differ significantly from straight (synoptic) winds, special studies are warranted. As the
2 profiles for Exposures B, C, and D are simplified to work within the constraints of the standard,
3 it is acceptable to use site-specific profiles determined by a rational analysis method defined in
4 the recognized literature, such as the ESDU model, in their place.
5 Research by Vickery et al. (2009a) suggests that in hurricanes, specifically near the eyewall, the
6 boundary layer height is lower than for synoptic winds. Where hurricane wind speeds govern the
7 design, the velocity pressure exposure coefficients in this standard, which are based on synoptic
8 winds, may conservatively model wind loads at elevations exceeding a few hundred meters

LY
9 (about 600-700 ft) above the ground, especially in Exposure D. With proper documentation and
10 approval by the Authority Having Jurisdiction, the use of alternative exposure coefficients that

N
11 appropriately reflect this lower boundary layer in hurricane-prone regions is acceptable.
12 Table 26.10-1 was adjusted in ASCE 7-22 according to the updated Kz formulas presented above;

O
13 however, where updated values of Kz were within 0.01 of the tabular values presented in
14
n
previous editions of ASCE 7, the previous values were retained. Variances in Kz of 0.01 or less
io
15 occurred at heights of up to 30 ft (9.1 m) for Exposure B, up to 120 ft (36.6 m) for Exposure C,
16 and up to the table limit of 500 ft (152.4 m) for Exposure D. The intent was to dismiss
at

17 insignificant changes to Kz values in Table 26.10-1 at heights covering the majority of low-rise
m

18 buildings and other structures for which tabular values of Kz are most often used.
19 The values of α given in Table 26.11-1 (7.5, 9.8, and 11.5 for Exposures B, C, and D,
or

20 respectively) define gust profiles used to calculate Kz, and were updated in ASCE 7-22 for the
nf

21 first time since 1995. The updated  values of mean hourly profiles in Table 26.11-1 based on
rI

22 4.5, 6.4, and 8.0 are used only to calculate the gust factor, Gf , for flexible structures in Section
23 26.11.5.
Fo

24

25 Changes were implemented in ASCE 7-98, including truncation of Kz values for Exposures A
26 and B below heights of 100 ft (30.5 m) and 30 ft (9.1 m), respectively, applicable to C&C and
27 the Envelope Procedure. Exposure A was eliminated in the 2002 edition.
28

46
1 In the ASCE 7-05 standard, the Kz expressions were unchanged from ASCE 7-98. However, the
2 possibility of interpolating between the standard exposures using a rational method was added in
3 the ASCE 7-05 edition. One rational method is provided in the following text.
4

5 To a reasonable approximation, the empirical exponent α and gradient height zg in the preceding

6 expressions [Equations (C26.10-1) through (C26.10-3)] for exposure coefficient Kz may be

7 related to the roughness length z0 (defined in Section C26.7) by the relations

LY
.
8 ∝ 𝑐𝑧 (C26.10-4)

N
9 and
.
10 𝑧 𝑐 𝑧 (C26.10-5)

O
11 where c1 is 6.69 and c2 is 1,166 when the units of z0 and zg are m, and c1 is 7.59 and c2 is 3,284
z0 and zg are ft. n
io
12 when the units of

13
at

14 The preceding relationships are based on matching the empirical boundary-layer models
m

15 proposed by Harris and Deaves (1981) and Kelly at al. (2019) with the power law relationship in
or

16 Equations (C26.10-1) through (C26.10-3). The models correspond to a gradient wind of 168 mi/h

17 (75 m/s) at latitude 40 degrees. If z0 has been determined for a particular upwind fetch,
nf

18 Equations (C26.10-1) through (C26.10-5) can be used to evaluate Kz. The correspondence
rI

19 between z0 and the parameters α and zg implied by these relationships does not align exactly
Fo

20 with that described in Table 26.11-1; however, the differences are relatively small and not of

21 practical consequence. The following simplified method (Irwin 2006) allows evaluation of Kz
22 following a transition from one surface roughness to another.
23
24 In uniform terrain, the wind travels a sufficient distance over the terrain for the planetary
25 boundary layer to reach an equilibrium state. The exposure coefficient values in Table 26.10-1
26 are intended for this condition. Suppose that the site is x miles downwind of a change in terrain.

47
1 The equilibrium value of the exposure coefficient at height z for the terrain roughness downwind

2 of the change will be denoted by Kzd , and the equilibrium value for the terrain roughness

3 upwind of the change will be denoted by Kzu . The effect of the change in terrain roughness on

4 the exposure coefficient at the site can be represented by adjusting Kzd by an increment Δ K ,

5 thus arriving at a corrected value Kz for the site:

6 Kz  Kzd ΔK (C26.10-6)

LY
7 In this expression, Δ K is calculated using
K zd
Δ K  ( K 33,u  K 33, d ) FΔ K ( x )

N
8 K 33, d (C26.10-7)
| Δ K || K zu  K zd |

O
9 where K33,d and K33,u are the downwind and upwind equilibrium values, respectively, of the
n FΔK(x) is given by
io
10 exposure coefficient at 33 ft (10 m) height, and the function

x
at
log10 ( 1 )
11 FΔK ( x)  x (C26.10-8)
x1
m

log10 ( )
x0
or

12 For x0 xx1:
nf

13 FΔk (x) 1 for x  x0


rI

14 FΔk (x)  0 for x  x1


Fo

15
16 In the preceding relationships,
( K33,d K33,u )2 2.3
17 x0  c3 10 (C26.10-9)

18

48
1 The constant c3  0.621 mi (1.0 km). The length x1  6.21 mi (10 km) for K33,d  K33,u (wind

2 going from smoother terrain upwind to rougher terrain downwind) or x1  62.1 mi (100 km) for
3 K33,d  K33,u (wind going from rougher terrain upwind to smoother terrain downwind).
4
5 The above description is in terms of a single roughness change. The method can be extended to
6 multiple roughness changes. The extension of the method is best described by an example.
7 Figure C26.10-1 shows wind with an initial profile characteristic of Exposure D encountering an

LY
8 expanse of B roughness, followed by a further expanse of D roughness, and then some more B
9 roughness again, before it arrives at the building site. This situation is representative of wind

N
10 from the sea flowing over an outer strip of land, then a coastal waterway, and then some

O
11 suburban roughness before arriving at the building site. The above method for a single roughness

12 change is first used to compute the profile of Kz at Station 1 in Figure C26.10-1. Call this profile
n
io
13 Kz(1) . The value of Δ K for the transition between Stations 1 and 2 is then determined using the
at

14 equilibrium value of K33,u for the roughness immediately upwind of Station 1, that is, as though
m

15 the roughness upwind of Station 1 extended to infinity. This value of Δ K is then added to the

Kzd(2)
or

16 equilibrium value of the exposure coefficient for the roughness between Stations 1 and 2 to

Kz at Station 2, which we will call Kz(2) . Note, however, that the value of
nf

17 obtain the profile of


rI

18 Kz(2) obtained in this way cannot be any lower than Kz(1) . The process is then repeated for the
19 transition between Stations 2 and 3. Thus, Δ K for the transition from Station 2 to Station 3 is
Fo

20 calculated using the value of K33,u for the equilibrium profile of the roughness immediately

21 upwind of Station 2, and the value of K33,d for the equilibrium profile of the roughness

22 downwind of Station 2. This value of Δ K is then added to Kzd(2) to obtain the profile Kz(3) at

23 Station 3, with the limitation that the value of Kz(3) cannot be any higher than Kz(2) .
24

49
1
2 Figure C26.10-1. Multiple roughness changes caused by coastal waterway.
3

LY
4 Example 1 (US Customary Units): Single Roughness Change. Suppose that the building is
5 66 ft high, and its local surroundings are suburban, with a roughness length z0 = 1 ft. However,
6 the site is 0.37 mi downwind of the edge of the suburbs, beyond which the terrain is

N
7 characteristic of open country, with z0 = 0.1 ft. From Equations (C26.10-4) and (C26.10-5), for

O
8 the open terrain,
9
∝ 𝑐 𝑧 .
7.59 x 0.1 .
9.6
n
io
10
. .
11 𝑧 𝑐 𝑧 3.284 x 0.1 2,451 ft
at

12
m

13 Therefore, applying Equation (C26.10-2) at 66 ft (20 m) and 33 ft (10 m) heights,


14
or

/ .
15 𝐾 2.41 1.13
nf

16 and
rI

/ .
17 𝐾 , 2.41 0.98
,
Fo

18
19 Similarly, for the suburban terrain,
20
. .
21 ∝ 𝑐 𝑧 7.59 x 0.1 7.6
. .
22 𝑧 𝑐 𝑧 3.284 x 0.1 3,284 ft
23
24 Therefore,
25

50
/ .
1 𝐾 2.41 0.86
,

2 and
/ .
3 𝐾 , 2.41 0.72
,

4
5 From Equation (C26.10-9),
6
. . . .
7 𝑥 𝑐 𝑥10 , , 0.621𝑥10 0.00266 mi

LY
8
9 From Equation (C26.10-8),

N
10

O
. / .
11 𝐹∆ 𝑥 0.36
. / .

12
13 Therefore, from Equation (C26.10-7), n
io
14
at
.
15 ∆𝐾 0.98 0.72 0.36 0.11
.
m

16

is 0.11, which is less than the 0.26 value of Δ| K33,u  K33,d |, 0.11 is
or

17 Note that because |ΔK |

Kz is
nf

18 retained. Finally, from Equation (C26.10-6), the value of

19
rI

20 Kz = Kzd + ΔK = 0.86 + 0.11 = 0.97


Fo

21

22 Because the value 0.97 for Kz lies between the values 0.85 and 1.15, which would be derived
23 from Table 26.10-1 for Exposures B and C, respectively, it is an acceptable interpolation. If it

24 falls below the Exposure B value, then the Exposure B value of Kz is to be used. The value

25 Kz 0.97 may be compared with the value 1.15 that would be required by the simple 2,600 ft
26 fetch length requirement of Section 26.7.3.
27

51
1 The most common case of a single roughness change where an interpolated value of

2 Kz is needed is for the transition from Exposure C to Exposure B, as in the example just

3 described. For this particular transition, using the typical values of z0 of 0.1 ft and 1.0 ft, the
4 preceding formulas can be simplified to
5
.
6 𝐾 𝐾 1 0.107 log (C26.10-10)

7 KzB ≤ Kz ≤ KzC

LY
8 where x is in miles, and Kzd is computed using α = 7.6.

N
KzB KzC

O
10 and are the exposure coefficients in the standard Exposures B and C, respectively.

11 Figure C26.10-2 illustrates the transition from Terrain Roughness C to Terrain Roughness B

12 n
from this expression. Note that it is acceptable to use the typical z0 rather than the lower limit for
io
13 Exposure B in deriving this formula because the rate of transition of the wind profiles depends
at

14 on average roughness over significant distances, not local roughness anomalies. The potential
m

15 effects of local roughness anomalies, such as parking lots and playing fields, are covered by

KzB , as a lower limit to the


or

16 using the standard Exposure B value of exposure coefficient,

Kz.
nf

17 calculated value of

18
rI
Fo

52
LY
N
1

O
2
3 Figure C26.10-2. Transition from Terrain Roughness C to Terrain Roughness B,
4 Equation (C26.10-10).
n
io
5
at

6 Example 2: Multiple Roughness Change. Suppose we have a coastal waterway situation, as


m

7 illustrated in Figure C26.10-1, where the wind comes from open sea with roughness type D, for

z0  0.01ft , and passes over a strip of land 1 mi wide, which is covered in


or

8 which we assume

buildings that produce typical B-type roughness, that is, z0 1ft . It then passes over a 2-mile-
nf

10 wide strip of coastal waterway, where the roughness is again characterized by the open-water
rI

11 value, z0  0.01ft . It then travels over 0.1 mi of roughness type B ( z0 1ft ) before arriving at
Fo

12 the building site, Station 3 in Figure C26.10-1, where the exposure coefficient is required at the
13 50 ft height. The exposure coefficient at Station 3 at a 50 ft height is calculated as shown in
14 Table C26.10-1.
15
16 Table C26.10-1. Tabulated Exposure Coefficients

Transition from sea to Station 1


K33,u K33,d K50,d FΔK ΔK50 K50(1)
1.251 0.717 0.800 0.221 0.132 0.932

53
Transition from Station 1 to Station 2
K33,u K33,d K50,d FΔK ΔK50 K50(2)
0.835 1.251 1.339 0.334 –0.149 1.191
Transition from Station 2 to Station 3
K33,u K33,d K50,d FΔK ΔK50 K50(3)
1.113 0.717 0.800 0.519 0.229 1.029

1 Note: The equilibrium values of the exposure coefficients, K33,u , K33,d and K50,d (downwind

2 value of Kz at 50 ft), were calculated from Equation (C26.10-2) using α and zg values obtained

3 from Equations (C26.10-4) and (C26.10-5) with the roughness values given. Then FΔK is

LY
4 calculated using Equations (C26.10-8) and (C26.10-9), and then the value of Δ K at 50 ft height,

N
5 ΔK50 , is calculated from Equation (C26.10-7). K33,u, K33,d, and K50,d from Station 1 to Station 3

O
6 include the computed Exposure Kz determined in the calculation rows upwind. Finally, the

7 exposure coefficient at 50 ft at Station i , K50(i ) , is obtained at each transition as well as at the site
8 from Equation (C26.10-6). n
io
9
at

10 The value of the exposure coefficient at 50 ft at Station 3 is seen from the table to be 1.029. This
m

11 is above the value in Table 26.10-1 for Exposure B, which would be 0.79, but well below that for
12 Exposure D, which would be 1.27, and somewhat below that for Exposure C, which would be
or

13 1.09.
nf

14 C26.10.2 Velocity Pressure


rI

qz in lb/ft2 (N/m2) at height z by the use


Fo

15 The basic wind speed is converted to velocity pressure

16 of Equation (26.10-1).
17
18 The constant in this equation reflects the mass density of air for the standard atmosphere, that is,
19 temperature of 59 °F (15 °C) and sea level pressure of 29.92 in. (101.325 kPa) of mercury, equal
20 to 0.0765 lbm/ft3 or 0.002378 slug/ft3 or 0.002378 lb-s2/ft4 (1.225 kg/m3), and dimensions
21 associated with wind speed in mi/h (m/s). The constant is obtained as follows.
22

54
1 Dynamic pressure from Bernoulli’s law:

2 p ½ρV2
3 with V in mi/h,
p  ½(0.002378 lb s 2 / ft 4 )[V mi / h (88 ft / s / 60 mi / h) 2 ]
4
 0.00256V 2 lb / ft 2

5 with V in m/s,
p  ½(1.225 kg / m3 )(1 N /1 kg m / s 2 )(V m / s) 2
6 (SI)
 0.613V 2 N / m2

LY
7
8 Values of air density other than the standard-atmosphere values above may be adjusted using the

N
9 factor Ke as described in Section C26.9.

O
10 In ASCE 7-22 the wind directionality factor, Kd, was removed from the velocity pressure
11
n
equation and placed in the individual wind pressure equations and force equations. Prior to
io
12 ASCE 7-98 the effect of Kd had been incorporated into the wind load factor. When Kd was
13 removed from the load factor it was explicitly introduced into the standard in ASCE 7-98
at

14 through the calculation of the velocity pressure. Kd is a function of (1) the directionality of the
m

15 extreme wind climate and (2) the shape of the building or other structure and the directional
or

16 characteristics of the structural system, so it is more appropriate to incorporate Kd into the


17 individual design wind pressure and force equations. The effect of building or other structure
nf

18 shape on Kd is evident in Table 26.6-1. In the velocity pressure equation used before ASCE 7-22,
rI

19 any building or other structure that comprises multiple shapes will have velocity design pressures
20 which is not consistent with the physics.
Fo

21 C26.11 Gust Effects

22 This standard specifies a single, conservative, gust-effect factor of 0.85 for rigid buildings. As an
23 option, the designer can incorporate specific features of the wind environment and building size
24 to more accurately calculate an alternate but more accurate gust-effect factor that accounts for
25 the decorrelation of wind gusts over the size of the structure. One such procedure is in the body
26 of the standard (Solari and Kareem 1998). Neither of these factors accounts for dynamic
27 amplification caused by vibration of the structure, but they are considered acceptable for rigid

55
1 structures as defined in the standard. The alternate calculated gust factor is 5% to 10% lower
2 than the value of 0.85 permitted in the standard without calculation.
3

4 A third gust-effect factor, Gf , is provided for flexible buildings and structures that do not meet

5 the requirements that the fundamental natural frequency, n1, is greater than or equal to 1 Hz.
6 This factor accounts for the building size and gust size in the same manner as the alternate
7 calculated factor for rigid buildings, but it also accounts for dynamic amplification caused by the

LY
8 design wind speed, the fundamental natural frequency of vibration, and the damping ratio. The
9 natural frequency should account for the minimum expected structural stiffness under the design

N
10 wind event V, be it the ultimate wind speed or any desired serviceability speed, and the mean
11 building mass expected under the design wind event, including self-weight, dead load, and

O
12 sustained live load. The damping ratio should account for inherent structural damping under the
13
n
design wind event (see subsection following). It may also include damping added by
io
14 supplementary devices if they are diagnosed to be operational under the design wind event. The
15 in-service reliability of such devices should be ensured through an established and documented
at

16 maintenance program.
m

17
18 Example: Calculation of the gust-effect factors for a subject building is demonstrated in
or

19 Table C26.11-1. The frequency-dependent relationship among all factors is illustrated in the
nf

20 graph at the end of this table. The flexiblily factor, Gf , may be used for all cases but is required
rI

21 when n1 1. This factor gradually approaches the alternate calculated factor for rigid cases, G ,
Fo

22 as the natural frequency exceeds 1, especially for higher levels of damping, but it always exceeds

23 G . The difference is deemed negligible for n1 > 1, so G , which is considerably simpler to

24 calculate, is offered as an acceptable alternative. The default value of G  0.85 , which requires no

25 calculation, is offered as an even more convenient alternative when n1 1, if the greater
26 conservatism is acceptable to the designer. In addition, the default value results in a large abrupt
27 change in the gust-effect value for cases that have a natural frequency close to 1, which may be

56
1 awkward for a designer to reconcile. A designer is free to use any other rational procedure in the
2 approved literature, as stated in Section 26.11.5.
3
4 Table C26.11-1. Gust-Effect Factor, Example.
Example Calculation—Gust-Effect Factors
Item Value Source
Default factor for rigid building (requires
n1 1)
G Gust-effect factor 0.85 26.11.1

LY
Alternate calculated factor for rigid

building (requires n1 1)

N
h Mean roof height 600 ft (183 m) User spec

O
B Width normal to wind 100 ft (30 m) User spec
D Depth parallel to wind 100 ft (30 m) User spec
z Effective structure height
Exposure category B n
360 ft (110 m) 0.6h (26.11.4)
io
c Turbulence intensity factor at 10 m 0.3 Table 26.11-1
at
0.201 (26.11-7)
Iz Turbulence intensity at eff. height
m

ℓ Turbulence length scale at 33 ft (10 m) 320 ft (98 m) Table 26.11-1


ε Power law exponent of turbulent length 1/3 Table 26.11-1
or

scale profile
710 ft (216 m) (26.11-9)
Lz Turbulence length scale at eff. height
nf

0.616 (26.11-8)
Q2 Background response (squared)
rI

3.4 26.11.4
gQ Background load peak factor
Fo

3.4 26.11.4
gv Velocity peak factor
G Calculated gust-effect factor 0.860 (26.11-6)
Additional calculations for flexible

building (all n1 )

V Basic wind speed 115 mi/h (51 m/s)


0.2 Hz Analysis or rational approximation
n1 Fundamental natural frequency in
direction of wind
β Damping ratio 0.01 Rational assignment

57
 Power law exponent of mean wind speed 1/4.5 Table 26.11-1
profile
b Gust factor 1/F at 33 ft (10 m) 0.47 Table 26.11-1
135 ft/s (41.1 m/s) 26.11-16
Vz Mean wind speed at effective height
1.053 26.11-14
N1 Reduced natural frequency
0.128 26.11-13
Rn Resonance response factor for n
h Vertical decay parameter 4.094 26.11.5: 4.6n1h/𝑉

B Cross-wind decay parameter

LY
0.682 26.11.5: 4.6n1B/𝑉

L Along-wind decay parameter 2.285 26.11.5: 15.4n1L/𝑉

N
0.214 (26.11-15a)
Rh Resonant factor for h

O
0.666 (26.11-15a)
RB Resonant factor for B
RL Resonant factor for L 0.343
n (26.11-15a)
io
R2 Resonant response factor (squared) 1.261 (26.11-12)
at
3.787 (26.11-11)
gR Resonant peak factor
m

1.162 (26.11-10)
Gf Gust-effect factor
or

1
nf
rI
Fo

58
LY
N
O
n
io
1
2
at

3 The gust-effect factors account for loading effects in the along-wind direction caused by wind
m

4 turbulence–structure interaction. They do not include allowances for across-wind loading effects,
5 vortex shedding, instability caused by galloping or flutter, or amplification of aerodynamic
or

6 torsion caused by building vibration in a pure torsional mode. For structures susceptible to
nf

7 loading effects that are not accounted for in the gust-effect factor, information should be
8 obtained from recognized literature (Kareem 1992, 1985; Gurley and Kareem 1993; Solari
rI

9 1993a, b; Zhou et al. 2002; Chen and Kareem 2004; Bernardini et al. 2013a) or from wind tunnel
Fo

10 tests.
11 Along-Wind Response. The maximum along-wind displacement response can be approximated
12 by a static analysis of the structure under the action of loads multiplied by the appropriate gust-
13 effect factor, as defined in the standard. Such displacements are based on the static elastic curve
14 of the structure and are reasonably accurate when the resonant response is small compared to the
15 mean and background responses. For highly flexible structures, where the response is dominated
16 by resonance, more accurate values, including variation with height and dynamic responses such

59
1 as acceleration, can be calculated as described in the following sections. These response
2 components are needed for strength and serviceability limit states.
3 The maximum along-wind displacement as a function of height above the ground surface is
4 given by
 ( z )ρBhC fxVˆz2
5 X max ( z )  KG f (C26.11-1)
2 m1 (2 πn1 ) 2

6 where

7  (z) = Fundamental model shape, (z)  (z / h) ;

LY
8 ξ = Mode shape power-law exponent;
9 ρ = Air density;

N
10 C fx = Mean along-wind force coefficient;

O
h
11 m1 = Modal mass =  μ(z)2 (z)dz ;
0
n
io
12 n1 = Fundamental natural frequency;
at

13 μ(z) = Mass per unit height: K (1.65)


α̂
/(α̂ξ 1) ; and
m

14 Vˆz = 3-s gust speed at height z.


15
or
nf

 88 
16 Vˆz  bˆ( z / 33)α̂V  
 60 
rI

17 where V is the 3-s gust speed in Exposure C (mi/h) at the reference height (obtained from Figure
Fo

18 26.5-1); and b̂ and α̂ are given in Table 26.11-1.


19 The root-mean-square (RMS) along-wind acceleration σ ¨ ( z ) as a function of height above the
x

20 ground surface is given by


0.85φ( z )ρBhC fxVz2
21 σ ¨ ( z)  I z KR (C26.11-2)
x m1

22 where Vz is the mean hourly wind speed at height z (ft/s).

60
α
 z   88 
1 Vz  b   V   (C26.11-3)
 33   60 
2 where b and α are defined in Table 26.11-1.
3 The maximum along-wind acceleration as a function of height above the ground surface is given
4 by
¨
5 X max ( z )  g ¨ σ ¨ ( z ) (C26.11-4)
x x

6 where

LY
0.5772
7 g ¨  2 ln( n1T )  (C26.11-5)
x 2 ln( n1T )

N
8 where T = length of time over which the minimum acceleration is computed, usually taken to be
9 3,600 s to represent 1 h.

O
10 Example calculations of maximum along-wind displacement, RMS along-wind acceleration, and
11 maximum along-wind acceleration are given in Table C26.11-2.
n
io
12
13 Table C26.11-2. Example Calculation.
at

Example Calculation – Along-wind Response


m

See Table C26.11-1 for additional items not shown


Item Value Source
or

V Basic wind speed 115 mi/h Figure 26.5-1A


 Air density Site elevation near sea
0.0024slug/ft3
nf

level
0.8 Figure 27.4-1
Cpw External pressure coefficient, windward wall
rI

 0.8 Figure 27.4-1


Cpw External pressure coefficient, leeward wall
Fo

1.3
C fx Along-wind force coefficient

 Mode shape power law exponent 1 Analysis of structure


ˆ Peak velocity power law exponent 1/7.5 Table 26.11-1
b̂ Velocity profile parameter 0.84 Table 26.11-1

Vˆz
195 ft/s Equation (C26.11-1a)
3-s gust velocity at height z

K Modal load parameter 0.501

61
b Building density 12 lbm/ ft3 , Building design

0.3727slug/ft3
 Building mass per unit height 3,727 slug/ft
745,400 slug
m1 Modal mass
Maximum along-wind displacement at building
top
 ( h ) Mode shape at z  h 1.0
1.76 ft Equation (C26.11-1)
Xmax(h)

LY
RMS along-wind acceleration at building top
V Basic wind speed, 10-year MRI 76 mi/h Figure CC-1

N
89.1 ft/s Equation (C26.11-3)
Vz Mean velocity at height z

O
R Resonant response factor 1.123 Equation (26.11-12)

 (h)
2
0.176 ft/s Equation (C26.11-2)
¨
x
n
io
Maximum along-wind acceleration at building
top
at
T Time period for maximum 3600 s Traditional

g¨ Peak factor 3.79 Equation (C26.11-5)


m

¨ 0.668 ft/s2 Equation (C26.11-4)


X max (h)
or

1
nf

2 Approximate Fundamental Frequency. To estimate the dynamic response of structures,


3 knowledge of the fundamental frequency (lowest natural frequency) of the structure is essential.
rI

4 This value would also assist in determining whether the dynamic response estimates are
Fo

5 necessary. Most computer codes used in the analysis of structures would provide estimates of the
6 natural frequencies of the structure being analyzed. For the preliminary design stages, some

7 empirical relationships for building period Ta ( Ta 1/ n1) are available in the earthquake-related
8 chapters of this standard. However, these expressions are based on recommendations for
9 earthquake design with inherent bias toward higher estimates of fundamental frequencies (Goel
10 and Chopra 1997, 1998). For wind design applications, these values may be unconservative
11 because an estimated frequency higher than the actual frequency would yield lower values of the
12 gust-effect factor and thus a lower design wind pressure. However, Goel and Chopra also cite

62
1 lower-bound estimates of frequency that are more suited for use in wind applications and are
2 now given in Section 26.11.2; graphs of these expressions are shown in Figure C26.11-1.
3

LY
N
O
n
io
4
at

5 Figure C26.11-1. Equations for approximate lower-bound natural frequency na versus building
m

6 height.
or

7 Notes:
8 1. Equation (26.11-2):
nf

2 2 .2 / h 0 .8

9 2. Equation (26.11-3): 4 3 .5 / h 0 .9
rI

10 3. Equation (26.11-4): 75 / h
11
Fo

12 Because these expressions are based on regular buildings, limitations based on height and

13 slenderness are required. The effective length, Leff , uses a height-weighted average of the
14 along-wind length of the building for slenderness evaluation. The top portion of the building is
15 most important; hence, the height-weighted average is appropriate. This method is an appropriate
16 first-order equation for addressing buildings with setbacks. Explicit calculation of the gust-effect
17 factor by the other methods given in Section 26.11 can still be performed.

63
1 Observations from wind tunnel testing of buildings where frequency is calculated using analysis
2 software show that the following expression for frequency can be used for steel and concrete
3 buildings less than about 400 ft (122 m) in height:
4
5 For the average value of the frequency,
6 𝑛 100/ℎ (C26.11-6)
7 𝑛 30.48/ℎ (C26.11-6.SI)
8

LY
9 For the lower bound value of the frequency,
10 𝑛 75/ℎ (C26.11-7)

N
11 𝑛 22.86/ℎ (C26.11-7.SI)
12

O
13 Equations (C26.11-7) and (C26.11-7.SI) for the lower-bound value are provided in Section
14 26.11.3.
n
io
15 Based on full-scale measurements of buildings under the action of wind, the following
16 expression has been proposed for wind applications (Zhou and Kareem 2001a, Zhou et al. 2002):
at

17 𝑛 150/ℎ (C26.11-8)
m

18 𝑛 45.72/ℎ (C26.11-8.SI)
or

19
20 This frequency expression is based on older buildings and overestimates the frequency common
nf

21 in US construction for buildings less than 400 ft (122 m) in height, but it becomes more accurate
rI

22 for tall buildings greater than 400 ft (122 m) in height.


23 Studies in Japan involving a suite of buildings under low-amplitude excitations have led to the
Fo

24 following expressions for natural frequencies of buildings (Sataka et al. 2003):


25
26 For concrete buildings,
27 𝑛 220/ℎ (C26.11-9)
28 𝑛 67/ℎ (C26.11-9.SI)
29
30 For steel buildings,
31 𝑛 164/ℎ (C26.11-10)

64
1 𝑛 50/ℎ (C26.11-10.SI)
2
3 These expressions result in higher frequency estimates than those obtained from the general
4 expression given in Equations (C26.11-6) through (C26.11-8), particularly since the Japanese
5 data set has limited observations for the more flexible buildings sensitive to wind effects, and
6 Japanese construction tends to be stiffer.
7 For cantilevered masts or poles of uniform cross section (in which bending action dominates),

8 n1  (0.56/ h2)(EI / m)

LY
9 where EI is the bending stiffness of the section, and m is the mass per unit height. This formula
10 may be used for masts with a slight taper, using average value of EI and m (ECCS 1978).

N
11 An approximate formula for cantilevered, tapered, circular poles (ECCS 1978) is

O
12 n1  [λ / (2πh2 )](EI / m) (C26.11-12)

13 where h is the height of the pole, and E , I, and


nm are calculated for the cross section at the
io
14 base. λ depends on the wall thicknesses at the tip and base ( et and eb ) and external diameter at the
at

15 tip and base ( dt and db), according to the formula


m

  4d t    6.65 
16 λ  1.9 exp     0.9  ( e / e ) 0.666  (C26.11-13)
  db   
or

t b

17
nf

18 Equation (C26.11-12) reduces to Equation (C26.11-11) for uniform masts. For freestanding
rI

19 lattice towers (without added ancillaries, such as antennas or lighting frames),


20 𝑛 4,921𝑤 /ℎ (C26.11-14)
Fo

21 𝑛 1,500𝑤 /ℎ (C26.11-14.SI)

22 where wa is the average width of the structure and h is the tower height (Standards Australia
23 1994). An alternative formula for lattice towers (with added ancillaries) is
2/3 1/2
L   wb 
24 n1   N    (C26.11-15)
 h   h 

25 where wb is the tower base width and LN = 885.8 ft (270 m) for square base towers, or 7546 ft
26 (230 m) for triangular base towers (Wyatt 1984).

65
1 Structural Damping. Structural damping is a measure of energy dissipation in a vibrating
2 structure that results in bringing the structure to a quiescent state. It is defined as the ratio of the
3 energy dissipated in one oscillation cycle to the maximum amount of energy in the structure in
4 that cycle. There are as many structural damping mechanisms as there are modes of converting
5 mechanical energy into heat. The most important mechanisms are material damping and
6 interfacial damping.
7 In engineering practice, the damping mechanism is often approximated as viscous damping
8 because it leads to a linear equation of motion. This damping measure, in terms of the damping

LY
9 ratio, is usually assigned based on the construction material, for example, steel or concrete. The
10 calculation of dynamic load effects requires damping ratio as an input. In wind applications,

N
11 damping ratios of 1% and 2% are typically used in the United States for steel and concrete
12 buildings, respectively, at serviceability levels, while ISO (1997) suggests 1% and 1.5% for steel

O
13 and concrete, respectively. Damping ratios for buildings under ultimate strength design
14
n
conditions may be significantly higher, and 2.5% to 3% is commonly assumed. Damping values
io
15 for steel support structures for signs, chimneys, and towers may be much lower than for
16 buildings and may fall in the range of 0.15% to 0.5%. Damping values of special structures like
at

17 steel stacks can be as low as 0.2% to 0.6% and 0.3% to 1.0% for unlined and lined steel
m

18 chimneys, respectively (ASME 1992; CICIND 1999). These values may provide some guidance
19 for design. Damping levels used in wind load applications are smaller than the 5% damping
or

20 ratios common in seismic applications because buildings subjected to wind loads respond
nf

21 essentially elastically, whereas buildings subjected to design-level earthquakes respond


22 inelastically at higher damping levels.
rI

23 Because the level of structural response in the strength and serviceability limit states is different,
Fo

24 the damping values associated with these states may differ. Furthermore, because of the number
25 of mechanisms responsible for damping, the limited full-scale data manifest a dependence on
26 factors such as material, height, and type of structural system and foundation (Kijewski-Correa
27 et al. 2013). The Committee on Damping of the Architectural Institute of Japan suggests
28 different damping values for these states based on a large damping database described in Sataka
29 et al. (2003).
30 The NatHaz Group developed an interactive database of full-scale experimentally determined
31 modal damping ratios based on the database (Kareem et al. 2012). The database is publicly

66
1 available at https://vortex-winds.org and is equipped with a query-based Web interface for the
2 rapid identification of modal damping ratios that satisfy specific requirements, such as geometric
3 form, structural system, construction material, foundation type, and building use. A recent
4 publication offers a data-driven model of damping that has been validated with several full-scale
5 studies (Spence and Kareem 2014).
6 In addition to structural damping, aerodynamic damping may be experienced by a structure
7 oscillating in air. In general, the aerodynamic damping contribution is quite small compared to
8 the structural damping, and it is positive in low-to-moderate wind speeds. Depending on the

LY
9 structural shape, at some wind velocities the aerodynamic damping may become negative, which
10 can lead to unstable oscillations. In these cases, reference should be made to recognized literature

N
11 or a wind tunnel study.
12 Alternate Procedure to Calculate Wind Loads. The concept of the gust-effect factor implies

O
13 that the effect of gusts can be adequately accounted for by multiplying the mean wind load
14
n
distribution with height by a single factor. This is an approximation. If a more accurate
io
15 representation of gust effects is required, the alternative procedure in this section can be used. It
16 takes account of the fact that the inertial forces created by the building’s mass, as it moves under
at

17 wind action, have a different distribution with height than the mean wind loads or the loads
m

18 caused by the direct actions of gusts (Zhou and Kareem 2001a; Chen and Kareem 2004). The
19 alternate formulation of the equivalent static load distribution uses the peak base bending
or

20 moment and expresses it in terms of inertial forces at different building levels. A base bending
nf

21 moment, instead of the base shear as in earthquake engineering, is used for the wind loads
22 because it is less sensitive to deviations from a linear mode shape. For a more detailed discussion
rI

23 on this wind-loading procedure, see Zhou and Kareem (2001a, 2001b) and Chen and Kareem
Fo

24 (2004).
25 Alternate Procedure: Along-Wind Equivalent Static Wind Loading. The equivalent static
26 wind loading for the mean, background, and resonant components is obtained using the
27 procedure outlined in this section.

28 Mean wind load component, Pj , is given by

29 Pj  qj Cp  Aj G (C26.11-16)
30 where

67
1 j = Floor level,

2 qj = Velocity pressure at height zj ,

3 zj = Height of the j th floor above the ground level,

4 Cp = External pressure coefficient, and

5 G = 0.925(11.7gvIz )1 = gust velocity factor.

6 ˆ , at the j th floor level is given similarly by


Peak background wind load component, P
Bj

LY
7 PˆBj  Pj GB / G (C26.11-17)

N
8 where
 1.7 I z  g Q Q 

O
9 G B  0.925    is the background component of the gust-effect factor.
 1  1.7 g v I z 

10 Peak resonant wind load component, P


Rj n
ˆ , at the j th floor level is obtained by distributing the
io
11 resonant base bending moment response to each level:
at

12 PˆRj  CMj Mˆ R (C26.11-18)


m

w jφ j
13 CMj  (C26.11-19)
w φ z
or

j j j

14 ˆ  M •G / G
M (C26.11-20)
nf

R R

15 M P •z (C26.11-21)
rI

j j
j 1,n

16 where
Fo

17 CMj = Vertical load distribution factor,

18 Mˆ R = Peak resonant component of the base bending moment response,

19 wj = Portion of the total gravity load of the building located or assigned to level j,

20 φj = First structural mode shape value at level j,

21 M = Mean base bending produced by mean wind load,

68
1 GR = 0.925[(1.7Iz gRR)/(11.7gvIz )] = Resonant component of the gust-effect factor, and
2 n = Total stories in the building.
3
4 Alternate Procedure: Along-Wind Response. Through a simple static analysis, the peak
5 building response in the along-wind direction can be obtained by

6 rˆ  r  rˆB2  rˆR2 (C26.11-22)

7 where r, rˆB , and rˆR is the mean, peak background, and resonant response components of interest,

LY
8 for example shear forces, moment, or displacement. Once the equivalent static wind load
9 distribution is obtained, any response component, including acceleration, can be obtained using a

N
10 simple static analysis. Caution is suggested when combining the loads instead of response

O
11 according to the preceding expression, as for example in

12 Pˆ j  Pj  PˆBj2  PˆRj2 (C26.11-23)


n
io
13 because the background and the resonant load components have normally different distributions
14 along the building height. Additional background can be found in Zhou and Kareem (2001b);
at

15 Zhou et al. (2002), and Chen and Kareem (2004).


m

16 Aerodynamic Loads on Tall Buildings: An Interactive Database. Under the action of wind,
17 tall buildings oscillate simultaneously in the along-wind, across-wind, and torsional directions.
or

18 While the along-wind loads have been successfully treated in terms of gust loading factors based
nf

19 on quasi-steady and strip theories, the across-wind and torsional loads cannot be treated in this
20 manner because these loads cannot be related in a straightforward manner to fluctuations in the
rI

21 approach flow. As a result, most current codes and standards provide little guidance for the
Fo

22 across-wind and torsional response (Zhou et al. 2002, Chen and Kareem 2004, Kwon and
23 Kareem 2013, Bernardini et al. 2013a).
24 To provide some guidance at the preliminary design stages of buildings, an interactive
25 aerodynamic loads database for assessing dynamic wind-induced loads on a suite of generic
26 isolated buildings is introduced (Zhou et al. 2002; Kwon et al. 2008). Although the analysis
27 based on this experimental database is not intended to replace wind tunnel testing in the final
28 design stages, it provides users a methodology to approximate the previously untreated across-
29 wind and torsional responses in the early design stages. The database consists of high-frequency

69
1 base balance measurements involving seven rectangular building models, with side ratio (D/B,
2 where D is the depth of the building section along the oncoming wind direction, and B is the
3 width of the building) from 1/3 to 3, and three aspect ratios for each building model in two

4 approach flows, namely BL1 ( α  0.16 ) and BL2 ( α  0.35 ), corresponding to an open and an
5 urban environment, respectively. The data are accessible with a user-friendly, Java-based
6 Internet applet, the NatHaz Aerodynamic Loads Database, version 2.0 (NALD 2012). Through
7 the use of this interactive portal, users can select the geometry and dimensions of a model
8 building from the available choices and specify an urban or suburban condition. The

LY
9 aerodynamic load spectra for the along-wind, across-wind, or torsional directions are then
10 displayed with a Java interface permitting users to specify a reduced frequency (building

N
11 frequency × building dimension / wind velocity) of interest and automatically obtain the

O
12 corresponding spectral value. When coupled with the supporting Web documentation, examples,
13 and concise analysis procedure, the database provides a comprehensive tool for computation of
14 n
wind-induced response of tall buildings, suitable as a design guide in the preliminary stages.
io
15 Example: An example tall building is used to demonstrate the analysis using the database. The
at

16 building is a square steel tall building with size H W1 W2  656 × 131 × 131 ft (200 × 40 ×
m

17 40 m) and an average radius of gyration of 59 ft (18 m).


18
or

19 The three fundamental mode frequencies, f1 , are 0.2, 0.2, and 0.35 Hz in the X , Y , and Z
nf

20 directions, respectively; the mode shapes are all linear, or β = 1.0, and there is no modal
rI

21 coupling. The building density is 0.485 slugs/ft3 (250 kg/m3). This building is located in
Fo

22 Exposure A or close to the BL2 test condition of the Internet-based database (Zhou et al. 2002,
23 Kwon et al. 2008). In this location, the 3-s design gust speed at a 700-year recurrence interval is
24 115 mi/h (51 m/s) in ASCE 7-16. For serviceability requirements, 3-s design gust speed with 10-
25 year MRI is 76 mi/h (34 m/s) in ASCE 7-16. For the sake of illustration only, the first-mode

26 critical structural damping ratio, ζ1 , is to be 0.01 for both survivability and serviceability design.
27 Using these aerodynamic data and the procedures provided by Zhou et al. (2002) and Kwon et al.
28 (2008), the wind load effects are evaluated and the results are presented in Table C26.11-3. This
29 table includes base moments and acceleration response in the along-wind direction obtained by
70
1 the procedure in ASCE 7-16. It should be pointed out that the building experiences higher
2 across-wind load effects when compared to the along-wind response for this example, which
3 reiterates the significance of wind loads and their effects in the across-wind direction and the
4 need for such database-enabled design approaches.
5
6 Table C26.11-3. Along-Wind, Across-Wind, Torsional Moments, and Acceleration
7 Response.
Survivability Design Serviceability Design

LY
Aerodynamic Load Base Moments ( Aerodynamic Load Accelerations (milli-g
Coefficient Coefficient or rad/s2)
106 kips ft )
Corner

N
Load
σCM f1 CM( f1) M
f1 CM( f1) a X Y

O
components
ASCE 7-10 – – – 1.73 – – 1.95 2.77 3.24
Along-wind 0.109 0.193 0.046 1.72 0.292 0.022 2.03
Across-wind 0.133 0.193 0.093 1.82 n 0.292 0.024 2.64
io
Torsional 0.044 0.337 0.040 0.086 0.512 0.043 0.0001
8 Note: As this database is experimental in nature thus has limitation in scope, it can be
at

9 conveniently expanded using additional data as it becomes available.


m

10 C26.12 Enclosure Classification


or

11 The magnitude and sense of internal pressure depend on the magnitude and location of openings
nf

12 around the building envelope with respect to a given wind direction. Accordingly, the standard
rI

13 requires that a determination be made of the amount of openings in the envelope to assess
Fo

14 enclosure classification (enclosed, partially enclosed, partially open, or open). Openings are
15 defined as holes that allow air to flow through the building envelope during a design wind event.
16 Holes in the building envelope, including air intake and exhaust vents for ventilation systems,
17 should always be considered as openings. Whether doors, operable windows and skylights, and
18 flexible and operable louvers are considered as openings depends on their intended use during a
19 storm event. For example, a door of a fire station or hospital ambulance bay should be
20 considered an opening if the intended function for such essential facilities requires the door to
21 remain in an open position during a design wind event. However, if doors, windows, and
22 skylights are likely be in a closed position during a design wind event, they do not need to be
71
1 considered as openings. The porosity of an “enclosed building” is such that there are not
2 sufficient openings in the exterior building envelope to allow significant airflow into the
3 building. The porosity of a “partially enclosed” building is such that there are sufficient openings
4 in the building envelope windward wall to allow wind to enter the building; however, there are
5 not sufficient openings in the remaining portions of the building envelope to allow airflow out of
6 the building without a buildup of internal pressure. The porosity of a “partially open” building is
7 such that there exist sufficient openings in the building envelope windward wall to allow airflow
8 into the building, and sufficient openings exist in the remaining portions of the building envelope

LY
9 to allow some airflow out of the building, but with some buildup of internal pressure. The
10 porosity for the “open building” is such that air flow can enter and exit the building without a

N
11 significant buildup of internal pressure. The classification of a “partially open” building has been
12 added to the standard to help the user in the understanding that a building with openings and

O
13 significant porosity (such as an open parking garage, for example) that does not meet the
14
n
requirements of the “partially enclosed” classification does not automatically classify the
io
15 building as “open” or “enclosed.” Once the enclosure classification is known, the designer uses
16 Table 26.13-1 to select the appropriate internal pressure coefficient.
at

17 Wind-borne debris: This version of the standard has five terms applicable to enclosure: wind-
m

18 borne debris regions, glazing, impact-resistant glazing, impact-protective system, and


19 unprotected glazing. Wind-borne debris regions are specified to alert the designer to areas
or

20 requiring consideration of test-missile-impact design and potential openings in the building


nf

21 envelope. Glazing is defined as any glass or transparent or translucent plastic sheet used in
22 windows, doors, skylights, or curtain walls. Impact-resistant glazing is specifically defined as
rI

23 glazing that has been shown by testing to withstand the impact of test missiles. Impact-protective
Fo

24 systems over glazing can be shutters or screens designed to withstand wind-borne debris impact.
25 Unprotected glazing is glazing that is not impact-resistant and does not have an impact-
26 protective system. Impact resistance of glazing and protective systems can be tested using the
27 test method specified in ASTM E1886, with test missiles, impact speeds, and pass/fail criteria
28 specified in ASTM E1996. Glazing in sectional doors, rolling doors, and flexible doors can be
29 tested for impact resistance with test missiles, impact speeds, and pass-fail criteria specified in
30 ANSI/DASMA 115. Sectional doors, rolling doors, and flexible doors are commonly used for
31 vehicle access and material transportation in warehouses, retail buildings, and other structures

72
1 requiring garage-type entry. A sectional door, sometimes known as a garage door, is a door made
2 of two or more horizontal sections, spanning between supports at the sides of the opening,
3 hinged together so as to provide a door capable of closing the entire opening, and which is raised
4 and lowered by means of the track rollers, in tracks which are in turn supported by the structure.
5 A rolling door is a vertically operating, coiling door typically used in commercial or industrial
6 applications. A flexible door is a door in which a flexible fabric or other flexible sheet material
7 forms the panel portion, even though it may have a rigid frame, rigid reinforcements, rigid
8 support means for one or more edges thereof, or combinations of these features. Other methods

LY
9 involving opening protection of building envelope systems are acceptable when approved by the
10 Authority Having Jurisdiction. The origins of the test-missile-impact provisions contained in

N
11 these standards are summarized by Minor (1994) and Twisdale et al. (1996).
12 Section 26.12.3 requires glazing in Risk Category II, III, and IV buildings in wind-borne debris

O
13 regions to be protected with an impact-protective system or to be made of impact-resistant
14
n
glazing to reduce the amount of interior wind and water damage to buildings during design
io
15 windstorm events (Surry et al. 1977; Reinhold 1982; Stubbs and Perry 1993). An exception in
16 Section 26.12.3.1 permits glazing to be unprotected when it is located more than 60 ft (18.3 m)
at

17 above grade and more than 30 ft (9.1 m) above aggregate-surfaced roofs located within 1,500 ft
m

18 (457 m) of the building. Unprotected glazing permitted by the Section 26.12.3.1 exception is not
19 considered an opening. Unprotected glazing higher than 60 ft (18.3 m) above grade may be
or

20 broken by wind-borne debris when a debris source is present. At these higher elevations, loose
nf

21 roof aggregate has been the predominant source of debris damage in previous wind events. This
22 includes gravel or stone used as ballast that is not protected by a sufficiently high parapet.
rI

23 Accordingly, if an existing aggregate-surfaced roof is within 1,500 ft (457 m) of a new building,


Fo

24 glazing in the new building, from 30 ft (9.1 m) above the source building to grade, needs to be
25 protected with an impact-protective system or be made of impact-resistant glazing. If loose roof
26 aggregate is proposed for the new building, it too should be considered as a debris source,
27 because aggregate can be blown off the roof and be propelled into glazing on the leeward side of
28 the building. The requirement for protection 30 ft (9.1 m) above the roof aggregate debris source
29 is to account for debris that can be lifted during flight. The following references provide further
30 information regarding debris damage to glazing: Beason et al. (1984), Minor (1985, 1994),
31 Kareem (1986), and Minor and Behr (1993, 1994). See C26.13 for discussion regarding breakage

73
1 of unprotected glazing higher than 60 ft (18.3 m) above grade by wind-borne debris other than
2 roof aggregate.
3 Although wind-borne debris can occur in just about any condition, the level of risk in
4 comparison to the postulated debris regions and test-missile-impact criteria may also be lower
5 than that determined for the purpose of standardization. For example, individual buildings may
6 be sited away from likely debris sources that would generate significant risk of impacts similar in
7 magnitude to pea gravel (i.e., small roof aggregate as simulated by 2 g steel balls in impact tests)
8 or butt-on 2  4 impacts, as required in impact-testing criteria. This situation describes a

LY
9 condition of low vulnerability only as a result of limited debris sources in the vicinity of the
10 building. In other cases, potential sources of debris may be present, but extenuating conditions

N
11 can lower the risk. These extenuating conditions include the type of materials and surrounding
12 construction, the level of protection offered by surrounding exposure conditions, and the basic

O
13 wind speed. Therefore, the risk of impact may differ from those postulated as a result of the
14
n
conditions specifically enumerated in the standard and the referenced impact standards. There
io
15 are vastly differing opinions regarding the significance of these parameters that are not fully
16 considered in developing standardized debris regions or referenced test-missile-impact criteria.
at

17 The definition of the wind-borne debris regions for Risk Category II buildings and structures was
m

18 chosen such that the coastal areas included in the wind-borne debris regions are approximately
19 consistent with those given in ASCE 7-05 and prior editions. Thus, the wind speed contours that
or

20 define the wind-borne debris regions in Section 26.12.3.1 are not direct conversions of the wind
nf

21 speed contours that are defined in ASCE 7-05, as shown in Table C26.5-7.
22 While the coastal areas included in the wind-borne debris regions for Risk Category II are
rI

23 approximately consistent with those given in ASCE 7-05, significant reductions in the area of
Fo

24 wind-borne debris regions for this Risk Category occur around Jacksonville, FL, in the eastern
25 part of the Florida Panhandle, and inland from the coast of North Carolina.
26 The introduction in ASCE 7-10 of separate maps for different Risk Categories provided a means
27 for achieving a more risk-consistent approach for defining wind-borne debris regions. The
28 approach selected was to link the geographical definition of the wind-borne debris regions to the
29 wind speed contours in the maps that correspond to the particular Risk Category, resulting in
30 expansion of the wind-borne debris region for some Risk Category III and all Risk Category IV
31 buildings and structures. A review of the types of buildings and structures currently included in

74
1 Risk Category III suggests that in the expanded wind-borne debris region, life-safety issues
2 would be most important for health-care facilities. Consequently, the committee chose to apply
3 the expanded wind-borne debris protection requirement to these types of Risk Category III
4 facilities and not to all Risk Category III buildings and structures.

5 C26.13 Internal Pressure Coefficients

6 The internal pressure coefficient values in Table 26.13-1 were obtained from wind tunnel tests
7 (Stathopoulos et al. 1979) and full-scale data (Yeatts and Mehta 1993). Even though the wind

LY
8 tunnel tests were conducted primarily for low-rise buildings, the internal pressure coefficient
9 values are assumed to be valid for buildings of any height. The values (GCpi) = +0.18 and –0.18

N
10 are for enclosed buildings. It is assumed that the building has no dominant opening or openings

O
11 and that the small leakage paths that do exist are essentially uniformly distributed over the
12 building’s envelope. The internal pressure coefficient values for partially enclosed buildings
13 n
assume that the building has a dominant opening or openings. For such a building, the internal
io
14 pressure is dictated by the exterior pressure at the opening and is typically increased substantially
at
15 as a result. Net loads (i.e., the combination of the internal and exterior pressures) are therefore
16 also significantly increased on the building surfaces that do not contain the opening. Therefore,
m

17 higher ( GCpi ) values of  0.55 and  0.55 are applicable to this case. These values include a
or

18 reduction factor to account for the lack of perfect correlation between the internal pressure and
nf

19 the external pressures on the building surfaces not containing the opening (Irwin 1987; Beste and
20 Cermak 1996). Taken in isolation, the internal pressure coefficients can reach values of  0.8 (or
rI

21 possibly even higher on the negative side).


Fo

22 For partially enclosed buildings containing a large unpartitioned space, the response time of the
23 internal pressure is increased, and this increase reduces the ability of the internal pressure to
24 respond to rapid changes in pressure at an opening. The gust factor applicable to the internal
25 pressure is therefore reduced. Equation (26.13-1), which is based on Vickery and Bloxham
26 (1992) and Irwin and Dunn (1994), is provided as a means of adjusting the gust factor for this
27 effect on buildings or other structures with large internal spaces, such as stadiums and arenas.
28 Breached glazing. Unprotected glazing may be breached by wind-borne debris during any type
29 of windstorm. However, the vast majority of damage has occurred during hurricanes and

75
1 tornadoes. See C26.12 for discussion of the requirements in Section 26.12.3 for protection of
2 glazing openings. The following discussion provides guidance on determining the internal
3 pressure coefficient with respect to glazing breaches caused by wind-borne debris.
4 Outside, and within, wind-borne debris regions, it is common to design new buildings for
5 enclosed conditions. Where unprotected glazing is permitted in wind-borne debris regions [i.e.,
6 greater than 60 ft (18.3 m)] above grade and not susceptible to roof aggregate per Section
7 26.12.3.1), the glazing can be considered as not being an opening. Numerous wind-damage
8 investigations have shown that extensive breakage of unprotected glazing higher than 60 ft (18.3

LY
9 m) above grade by wind-borne debris other than roof aggregate is very unlikely. It is possible for
10 a few, or several, panes of permitted unprotected glazing to be breached by debris. The breaching

N
11 may result in development of increased internal pressure in the area (compartment) behind the
12 breached glazing. The increased pressure may overload nonstructural building envelope

O
13 elements, as illustrated in Figures C26.13-1 and C26.13-2.
14
n
io
at
m
or
nf
rI
Fo

76
LY
N
O
n
io
at
m

1
or

2 Figure C26.13-1. Glazing breaches and non-load-bearing wall failures (Hurricane Ivan).
nf

3 Source: FEMA (2004).


4
rI

5 At the building shown in Figure C26.13-1, some glazing was protected by shutters (blue dashed
Fo

6 line), but most of the glazing was unprotected. Some of the glazing was breached by debris
7 (which was not roof aggregate), and some glazing assemblies collapsed (red dashed line) because
8 they were inadequately attached to the wall framing. See Figure C26.13-2 for a view of the
9 sidewall in the area shown by the red and blue lines (the exterior of the sidewall experienced
10 suction pressures).

77
LY
N
O
1
2 Figure C26.13-2. View of sidewall damage. n
io
3 Source: FEMA.
at

4
m

5 The wall shown by the red line is at the floor with the collapsed glazing assembly. The wall was
6 composed of stucco/gypsum board, metal framing, and gypsum board. In some areas, the entire
or

7 wall blew away. At the area shown by the yellow line, the stud track detached from the floor slab
nf

8 and deformed outward. The wall shown by the blue line is at the floor with the protected glazing.
9 At this wall area, the metal framing and most of the interior and exterior gypsum board was still
rI

10 in place. The difference between the amount of damage between the two wall areas was judged
Fo

11 to be due to development of increased internal pressure caused by collapse of the glazing


12 assembly. The wall was also judged to have inadequate resistance to wind loads derived from
13 ASCE 7 for an enclosed building.
14 Although the probability is low, buildings that are designed in accordance with ASCE 7, using
15 the internal pressure coefficients for an enclosed building, could experience breaching of
16 permitted unprotected glazing that results in development of increased internal pressure.
17 However, if properly constructed and maintained, damage to other portions of the building

78
1 envelope or the MWFRS would not be expected during a design event because of the additional
2 strength provided by safety and load factors.
3
4 Regarding distribution of breached glazing, the dominant opening(s) will initially be on a
5 windward façade. However, during hurricanes, the angle of attack may change and result in
6 breached glazing in more than one façade.
7 The ASCE/SEI Prestandard for Performance-Based Wind Design (2019) provides glazing
8 guidance for designers retained by building owners that desire to achieve higher wind

LY
9 performance.
10 Compartmentalization. The influence of compartmentalization on the distribution of increased

N
11 internal pressure has not been researched. If the space behind large breaches of the building
12 envelope (such as broken glazing or collapsed doors or walls) is separated from the remainder of

O
13 the building by a sufficiently strong and reasonably airtight compartment, the increased internal
14
n
pressure would likely be confined to that compartment. However, if the compartment is breached
io
15 (e.g., by an open corridor door or by collapse of the compartment wall), the increased internal
16 pressure will spread beyond the initial compartment quite rapidly. The next compartment may
at

17 contain the higher pressure, or it too could be breached, allowing the high internal pressure to
m

18 continue to propagate. For buildings with more than one floor, increased internal pressure may
19 spread to other floors.
or

20 Aircraft hangers. Because of the great amount of air leakage that often occurs at large hangar
nf

21 doors, designers of hangars should consider using the internal pressure coefficients in
22 Table 26.13-1 for partially enclosed buildings.
rI

23 Lobby entry vestibules. Vestibules can be exposed to increased internal pressure due to
Fo

24 operation or breach of exterior entry doors under high winds, subjecting interior vestibule doors,
25 walls, and ceilings to external pressures. For those Risk Category III and IV buildings where
26 operation of exterior entry doors either during or after a high wind event is critical to building
27 function, the interior vestibule envelope could be designed for the same pressures as the exterior
28 wall and doors. For other buildings, the interior vestibule envelope could be reasonably designed
29 to withstand external pressures at wind speeds expected during normal operation of exterior
30 entry doors, which would otherwise be closed in high wind events.

79
1 Remedial work on existing buildings located in a wind-borne debris region. If existing
2 glazing does not comply with the requirements of Section 26.12.3, then the existing glazing
3 should be replaced with new impact-resistant glazing or retrofitted with impact-protective
4 systems in accordance with Section 26.12.3. If the existing noncompliant glazing is to remain,
5 then it should be assumed that it can be breached by wind-borne debris, and load calculations for
6 new building-envelope elements should include a confirmation of building enclosure (i.e.
7 enclosed versus partially enclosed) to determine the correct internal pressure coefficient. If
8 replacing a roof system, uplift resistance of the roof deck and deck support structure should be

LY
9 considered.

18 REFERENCES

N
O
19 Almeida, G. P., D. F. G. Durao, and M. V. Heitor 1993. “Wake flows behind two-dimensional
20 model hills.” Experimental Thermal and Fluid Science 7: 87-101.
21 n
ANSI/DASMA 115-17 Standard Method for Testing Sectional Garage Doors: Determination of
io
22 Structural Performance Under Missile Impact and Cyclic Wind Pressure. Cleveland, OH:
at
23 DASMA.
24 ASCE/SEI. 2019. Prestandard for performance-based wind design. Reston, VA: ASCE.
m

25 ASME (American Society of Mechanical Engineers). 1992. Steel stacks. STS-1. New York:
or

26 ASME.
27 ARA (Applied Research Associates). 2001. Hazard mitigation study for the Hawaii Hurricane
nf

28 Relief Fund. Report 0476. Raleigh, NC: ARA.


rI

29 ARA (Applied Research Associates). 2019. Development of Wind Speed-ups and Hurricane
30 Hazard Maps for Puerto Rico, Final Draft Report, Raliegh, NC.
Fo

31 ASTM E330/E330M – 14 Standard Test Method for Structural Performance of Exterior


32 Windows, Doors, Skylights and Curtain Walls by Uniform Static Air Pressure Difference
33 ASTM E1886-19 Standard Test Method for Performance of Exterior Windows, Curtain Walls,
34 Doors, and Impact Protective Systems Impacted by Missile(s) and Exposed to Cyclic
35 Pressure Differentials West. Conshohocken, PA; ASTM
36 ASTM E1996-17 Standard Specification for Performance of Exterior Windows, Curtain Walls,
37 Doors, and Impact Protective Systems Impacted by Windborne Debris in Hurricanes.
38 West Conshohocken, PA; ASTM

80
1 Banks, D., N. Ellison, and W. Esterday. 2019. Kern County Special Wind Region Study, Final
2 Report. CPP Project 9413 Rev02. https://bit.ly/CPPKernCountySiteStudy.
3 Beason, W. L., G. E. Meyers, and R. W. James. 1984. “Hurricane related window glass damage
4 in Houston.” J. Struct. Eng. 110 (12): 2843–2857.
5 Behr, R. A., and J. E. Minor. 1994. “A survey of glazing system behavior in multi-story
6 buildings during Hurricane Andrew.” Structural Design of Tall Buildings 3: 143–161.
7 Bernardini, E., S. M. J. Spence, and A. Kareem. 2013a. “A probabilistic approach for the full
8 response estimation of tall buildings with 3D modes using the HFFB.” Struct. Saf. 44:

LY
9 91–101.
10 Bernardini, E., S. M. J. Spence, and A. Kareem. 2013b. “An efficient performance-based design

N
11 approach to the high frequency force balance.” In Proc., 11th Int. Conf. on Structural
12 Safety and Reliability, ICOSSAR 2013, G. Deodatis, B. Ellingwood, and D. Frangopol,

O
13 eds. London: Taylor & Francis, 1777–1784.
14
n
Beste, F., and J. E. Cermak. 1996. “Correlation of internal and area-averaged wind pressures on
io
15 low-rise buildings.” In Proc., 3rd Int. Colloquium on Bluff Body Aerodynamics and
16 Applications, Virginia Polytechnic Institute, Blacksburg.
at

17 Chen, X., and A. Kareem. 2004. “Equivalent static wind loads on buildings: New model.” J.
m

18 Struct. Eng. 130 (10): 1425–1435.


19 Chock, G., L. Peterka, and G. Yu. 2005. “Topographic wind speed-up and directionality factors
or

20 for use in the city and county of Honolulu building code.” In Proc., 10th Americas Conf.
nf

21 on Wind Eng., Louisiana State University, Baton Rouge.


22 Cleveland, W. S., and S. J. Devlin. 1988) “Locally weighted regression: An approach to
rI

23 regression analysis by local fitting.” J. Am. Stat. Assn. 83 (403): 596–610.


Fo

24 doi:10.2307/2289282.
25 CICIND (Comité International des Cheminées Industrielles). 1999. Model code for steel
26 chimneys, Revision 1-1999. Zurich.
27 CSA A123.21-14: Standard Test Method for the Dynamic Wind Uplift Resistance of Membrane-
28 Roofing Systems-Davenport, A.G. 1960. “Rationale for determining design wind
29 velocities.” Journal of the Structural Division 86(5): 39-68.

81
1 Davenport, A. G., C. S. B. Grimmond, T. R. Oke, and J. Wieringa. 2000. “Estimating the
2 roughness of cities and sheltered country.” Preprint, 12th AMS Conf. Applied Climatol.,
3 American Meteorological Society, Boston, 96–99.
4 Donelan, M. A., B. K. Haus, N. Reul, W. J. Plant, M. Stiassnie, H. C. Graber, et al. 2004. “On
5 the limiting aerodynamic roughness of the ocean in very strong winds.” Geophys. Res.
6 Lett. 31: 1–5.
7 Durst, C. S. 1960. “Wind speeds over short periods of time.” Meteorol. 89: 181–187.
8 Ellingwood, B. 1981. “Wind and snow load statistics for probabilistic design.” J. Struct. Div. 107

LY
9 (7): 1345–1350.
10 Ellingwood, B., J. G. MacGregor, T. V. Galambos, and C. A. Cornell. 1982. “Probability based

N
11 load criteria: Load factors and load combinations.” In J. Struct. Div. 108 (5): 978–997.
12 ECCS (European Convention for Structural Steelwork). (1978). Recommendations for the

O
13 calculation of wind effects on buildings and structures. Technical Committee T12.
14 Brussels: ECCS.
n
io
15 ESDU (Engineering Sciences Data Unit). 1982. Strong winds in the atmospheric boundary layer,
16 Part 1: Mean hourly wind speed. ESDU 82026. London: ESDU.
at

17 ESDU. 1993. Strong winds in the atmospheric boundary layer, Part 2: Discrete gust speeds.
m

18 ESDU 83045. London: ESDU.


19
or

20 FEMA (Federal Emergency Management Agency). 2021a. Safe rooms for tornadoes and
nf

21 hurricanes: Guidance for community and residential safe rooms. P-361, 4th ed..
22 Washington, DC: FEMA https://www.fema.gov/emergency-managers/risk-
rI

23 management/safe-rooms.
Fo

24 FEMA (Federal Emergency Management Agency). 2021b. Taking shelter from the storm:
25 Building a safe room for your home. P-320. 5th ed.,Washington, DC: FEMA.
26 https://www.fema.gov/emergency-managers/risk-management/safe-rooms.
27
28 Goel, R. K., and A. K. Chopra. 1997. “Period formulas for moment-resisting frame buildings.” J.
29 Struct. Eng. 123 (11): 1454–1461.
30 Goel, R. K., and A. K. Chopra. 1998. “Period formulas for concrete shear wall buildings.” J.
31 Struct. Eng. 124 (4): 426–433.

82
1 Gurley, K., and A. Kareem. 1993. “Gust loading factors for tension leg platforms.” Appl. Ocean
2 Res. 15 (3): 137–154.
3 Harris, R. I., and D. M. Deaves. 1981. “The structure of strong winds.” In Proc., CIRIA Conf.
4 Wind Eng. in the Eighties. London: Construction Industry Research and Information
5 Association.
6 Ho, E. 1992. Variability of low building wind lands. Ph.D. thesis, University of Western Ontario,
7 Dept. of Civil Engineering, London, ON.
8 ICC (International Code Council). 2020. ICC/NSSA standard for the design and construction of

LY
9 storm shelters. ICC 500-2020. International Code Council and National Storm Shelter
10 Association.

N
11 ISO (International Organization for Standardization). 1997. “Wind actions on structures.” ISO
12 4354. Geneva: ISO.

O
13 Irwin, P. A. 1987. “Pressure model techniques for cladding loads.” J. Wind Eng. Indust.
14 Aerodyn. 29: 69–78.
n
io
15 Irwin, P. A. 2006. “Exposure categories and transitions for design wind loads.” J. Struct. Eng.
16 132 (11): 1755–1763.
at

17 Irwin, P. A., and G. E. Dunn. 1994. “Review of internal pressures on low-rise buildings.” RWDI
m

18 Report 93-270 for Canadian Sheet Building Institute, Feb. 23.


19 Isyumov, N., E. Ho, and P. Case. 2013. “Influence of wind directionality on wind loads and
or

20 responses.” In Proc., 12th Americas Conf. Wind Eng., D. Reed and A. Jain, eds.
nf

21 Jackson, P. S., and J. C. R. Hunt. 1975. “Turbulent wind flow over a low hill.” Quart. J. Royal
22 Meteorol. Soc. 101: 929–955.
rI

23 Jung, S., and F. J. Masters. 2013. “Characterization of open and suburban boundary layer wind
Fo

24 turbulence in 2008 Hurricane Ike.” Wind Struct. 17: 135–162.


25 Kareem, A. 1985. “Lateral-torsional motion of tall buildings.” J. Struct. Eng. 111 (11): 2479–
26 2496.
27 Kareem, A. 1986. “Performance of cladding in Hurricane Alicia.” J. Struct. Eng. 112 (12):
28 2679–2693.
29 Kareem, A. 1992. “Dynamic response of high-rise buildings to stochastic wind loads.” J. Wind
30 Eng. Indust. Aerodyn. (41–44): 1101-1112

83
1 Kareem, A., D. K. Kwon, and Y. Tamura. 2012. “Cyberbased analysis, modeling and simulation
2 of wind load effects in VORTEX-winds.” Proc., 3rd American Assn. Wind Eng.
3 Workshop, Hyannis, MA, August 12–14.
4 Kelly, M., R. A. Cersosimo, and J. Berg. 2019. “A universal wind profile for the inversion-
5 capped neutral atmospheric boundary layer.” Quart. J. Royal Meteorol. Soc. 1–11.
6 Kern County. 2017. Title 17 Buildings and Construction, Chapter 17.08 Building
7 Code.17.08.420, Section 1609.3 amended, Ultimate design wind speed.
8 https://library.municode.com/ca/kern_county/codes/code_of_ordinances?nodeId=TIT17B

LY
9 UCO_CH17.08BUCO.
10 Kijewski-Correa, T., A. Kareem, Y. L. Guo, R. Bashor, and T. Weigand. 2013. “Performance of

N
11 tall buildings in urban zones: Lessons learned from a decade of full-scale monitoring.”
12 Int. J. High-Rise Bldgs. 2 (3): 179–192.

O
13 Kwon, D.-K., and A. Kareem. 2013. “Comparative study of major international wind codes and
14
n
standards for wind effects on tall buildings.” Eng. Struct. 51: 23–25.
io
15 Kwon, D.-K., T. Kijewski-Correa, and A. Kareem. 2008. “e-Analysis of high-rise buildings
16 subjected to wind loads.” J. Struct. Eng. 133 (7): 1139–1153.
at

17 Lemelin, D. R., D. Surry, and A. G. Davenport. 1988. “Simple approximations for wind speed-
m

18 up over hills.” J. Wind Eng. Indust. Aerodyn. 28: 117–127.


19 Lettau, H. 1969. “Note on aerodynamic roughness element description.” J. Appl. Meteorol. 8:
or

20 828–832.
nf

21 Lombardo, F., A. Pintar, P.J. Vickery, E. Simiu, and M. Levitan. 2016. Development of new
22 wind speed maps for ASCE 7–16. NIST Special Publication.
rI

23 Means, B., T. A. Reinhold, and D. C. Perry. 1996. “Wind loads for low-rise buildings on
Fo

24 escarpments.” In Building an international community of structural engineers, S. K.


25 Ghosh and J. Mohammadi, eds. Reston, VA: ASCE, 1045–1052.
26 Mehta, K. C., and R. D. Marshall. 1998. Guide to the use of the wind load provisions of ASCE 7-
27 95. Reston, VA: ASCE.
28 Minor, J. E. 1985. “Window glass performance and hurricane effects.” In Hurricane Alicia: One
29 year later, A. Kareem, ed. Reston, VA: ASCE, 151–167.
30 Minor, J. E. 1994. “Windborne debris and the building envelope.” J. Wind Eng. Indust. Aerodyn.
31 53: 207–227.

84
1 Minor, J. E., and R. A. Behr. 1993. “Improving the performance of architectural glazing in
2 hurricanes.” In Hurricanes of 1992: Lessons learned and implications for the future.
3 Reston, VA: ASCE, 476–485.
4 NALD (NatHaz Aerodynamic Loads Database. 2012. NatHaz aerodynamic loads database,
5 version 2. http://aerodata.ce.nd.edu/.
6 NHC (National Hurricane Center). 2015. Saffir–Simpson Hurricane Wind Scale.
7 http://www.nhc.noaa.gov/aboutsshws.php.
8 Palutikof, J. P., B. B. Brabson, D. H. Lister, and S. T. Adcock. 1999. “A review of methods to

LY
9 calculate extreme wind speeds.” Meteorological Applications 6: 119–132.
10 Panofsky, H. A. and J.A. Dutton. 1984. Atmospheric turbulence: models and methods for

N
11 engineering applications. 1st ed. Wiley-Interscience.
12 Peterka, J. A. 2006. Colorado Front Range Gust Map: CPP report for use by Colorado Structural

O
13 Engineers Association and Colorado Front Range communities.
14
n
https://bit.ly/2006ColoradoFrontRangeGustMap.
io
15 Peterka, J. A. 2010. “Colorado Front Range Gust Map – ASCE 7-10 Compatible”. Structural
16 Engineers Association of Colorado, Aurora. https://seacolorado.org/docs/FINAL-
at

17 COLORADO-FRONT-RANGE-GUST-MAP-2013.pdf.
m

18 Peterka, J. A., and S. Shahid. 1998. “Design gust wind speeds in the United States.” J. Struct.
19 Eng. 124 (2): 207–214.
or

20 Pickands, James, III. 1971. “The two-dimensional poisson process and extremal processes.” J.
nf

21 Appl. Prob. 8 (4): 745–56. doi:10.2307/3212238.


22 Pintar, A., E. Simiu, F. Lombardo, and M. Levitan. 2015. “Maps of non-hurricane non-tornadic
rI

23 wind speeds with specified mean recurrence intervals for the contiguous united states
Fo

24 using a two-dimensional poisson process extreme value model and local regression.”
25 Special Publication 500-301. National Institute of Standards and Technology.
26 http://dx.doi.org/10.6028/NIST.SP.500-301.
27 Powell, M. D. 1980. “Evaluations of diagnostic marine boundary-layer models applied to
28 hurricanes.” Monthly Weather Rev. 108 (6): 757–766.
29 Powell, M. D., P. J. Vickery, and T. A. Reinhold. 2003. “Reduced drag coefficients for high
30 wind speeds in tropical cyclones.” Nature 422: 279–283.

85
1 Reinhold, T. A., ed. 1982. “Wind tunnel modeling for civil engineering applications.” Proc., Int.
2 Workshop on Wind Tunnel Modeling Criteria and Techniques in Civil Engineering
3 Applications. Gaithersburg, MD: Cambridge University Press.
4 Sataka, N., K. Suda, T. Arakawa, A. Sasaki, and Y. Tamura. 2003. “Damping evaluation using
5 full-scale data of buildings in Japan.” J. Struct. Eng. 129 (4): 470–477.
6 SEAOC (Structural Engineers Association of California). 2017. “Recommendations for action to
7 address design wind speeds in California”. W1-2017. Wind Committee, Structural
8 Engineers Association of California, Sacramento.

LY
9 https://www.seaoc.org/store/ListProducts.aspx?catid=597243.
10 Searer, et. al. 2010. SEAOSC Summary Report: Study of Historical and Design Wind Speeds in

N
11 the Los Angeles Area. Los Angeles: Structural Engineers Association of Southern
12 California.

O
13 Simiu, E. 2011. Design of buildings for wind: A guide for ASCE 7-10 Standard users and
14
n
designers of special structures. 2nd ed. New York: Wiley.
io
15 Simiu, E., and R. H. Scanlan. 1996. Wind effects on structures. 3rd Ed. New York: Wiley.
16 Simiu, E., L. Shi, and D. Yeo. 2016. “Planetary boundary-layer modelling and tall building
at

17 design.” Boundary-Layer Meteorology 159 (1): 173-181.


m

18 Simiu, E., P. Vickery, and A. Kareem. 2007. “Relation between Saffir-Simpson hurricane scale
19 wind speeds and peak 3-s gust speeds over open terrain.” J. Struct. Eng. 133 (7): 1043–
or

20 1045.
nf

21 Simiu, E., and D. Yeo. 2019. Wind effects on structures: Modern structural design for wind. 4th
22 ed. Oxford: Wiley-Blackwell.
rI

23 Smith, R. L. 1989. “Extreme Value Analysis of Environmental Time Series: An Application to


Fo

24 Trend Detection in Ground-Level Ozone.” Stat. Science 4 (4): 367–377.


25 Solari, G. 1993a. “Gust buffeting. I: Peak wind velocity and equivalent pressure.” J. Struct. Eng.
26 119 (2): 365–382.
27 Solari, G. 1993b. “Gust buffeting. II: Dynamic alongwind response.” J. Struct. Eng. 119 (2):
28 383–398.
29 Solari, G., and A. Kareem. 1998. “On the formulation of ASCE 7-95 gust effect factor.” J. Wind
30 Eng. Indust. Aerodyn. 77-78: 673–684.

86
1 Spence, S., and A. Kareem. 2014. “Tall buildings and damping: A concept-based data driven
2 model.” J. Struct. Eng. 140 (5).
3 Standards Australia. (1994). “Design of steel lattice towers and masts.” AS3995-1994, North
4 Sydney: Standards Australia.
5 Stathopoulos, T., D. Surry, and A. G. Davenport. 1979. “Wind-induced internal pressures in low
6 buildings.” In Proc., 5th Int. Conf. on Wind Engineering, J. E. Cermak, ed. Colorado
7 State University, Fort Collins.
8 Stubbs, N., and D. C. Perry. 1993. “Engineering of the building envelope: To do or not to do.” In

LY
9 Hurricanes of 1992: Lessons learned and implications for the future, R. A. Cook and M.
10 Sotani, eds. Reston, VA: ASCE Press, 10–30.

N
11 Surry, D., R. B. Kitchen, and A. G. Davenport. (1977). “Design effectiveness of wind tunnel
12 studies for buildings of intermediate height.” Can. J. Civ. Eng. 4 (1): 96–116.

O
13 Tse, K. T., S. W. Li, P. W. Chan, H. Y. Mok, and A. U. Weerasuriya. 2013. “Wind profile
14
n
observations in tropical cyclone events using wind-profilers and doppler SODARs.”
io
15 Journal of Wind Engineering and Industrial Aerodynamics 115: 93-103.
16 Twisdale, L. A., P. J. Vickery, and A. C. Steckley. 1996. Analysis of hurricane windborne debris
at

17 impact risk for residential structures. Bloomington, IL: State Farm Mutual Automobile
m

18 Insurance Companies.
19 USNRC (US Nuclear Regulatory Commission). 2011) Technical basis for regulatory guidance
or

20 on design-basis hurricane wind speeds for nuclear power plants. NUREG/CR-7005.


nf

21 Washington, DC: US Nuclear Regulatory Commission.


22 Vickery, B. J., and C. Bloxham. 1992. “Internal pressure dynamics with a dominant opening.” J.
rI

23 Wind Eng. Indust. Aerodyn. 41–44: 193–204.


Fo

24 Vickery, P. J., and P. F. Skerlj. 2005. “Hurricane gust factors revisited.” J. Struct. Eng. 131 (5):
25 825–832.
26 Vickery, P. J., P. F. Skerlj, A. C. Steckley, and L. A. Twisdale. (2000). “Hurricane wind field
27 model for use in hurricane simulations.” J. Struct. Eng. 126 (10): 1203–1221.
28 Vickery, P. J., and D. Wadhera. 2008a. “Development of design wind speed maps for the
29 Caribbean for application with the wind load provisions of ASCE 7.” ARA Report
30 18108-1, prepared for Pan American Health Organization, Regional Office for the

87
1 Americas, World Health Organization, Disaster Management Programme, Washington,
2 DC.
3 Vickery, P. J., and D. Wadhera. 2008b. “Statistical models of the Holland pressure profile
4 parameter and radius to maximum winds of hurricanes from flight level pressure and H*
5 wind data.” J. Appl. Meteor. 47: 2497–2517.
6 Vickery, P. J., D. Wadhera, J. Galsworthy, J. A. Peterka, P. A. Irwin, and L. A. Griffis. 2010.
7 “Ultimate wind load design gust wind speeds in the United States for use in ASCE-7.” J.
8 Struct. Eng. 136: 613–625.

LY
9 Vickery, P. J., D. Wadhera, M. D. Powell, and Y. Chen. 2009a. “A hurricane boundary layer and
10 wind field model for use in engineering applications.” J. Appl. Meteorology 48: 381–405.

N
11 Vickery, P. J., D. Wadhera, L. A. Twisdale, Jr., and F. M. Lavelle. 2009b. “U.S. hurricane wind
12 speed risk and uncertainty.” J. Struct. Eng. 135 (3): 301–320.

O
13 Walmsley, J. L., P. A. Taylor, and T. Keith. 1986. “A simple model of neutrally stratified
14
n
boundary-layer flow over complex terrain with surface roughness modulations.”
io
15 Boundary-Layer Meteorol. 36: 157–186.
16 Wieringa, J., A. G. Davenport, C. S. B. Grimmond, and T. R. Oke. 2001. “New revision of
at

17 Davenport roughness classification.”.


m

18 Wyatt, T. A. 1984. “Sensitivity of lattice towers to fatigue induced by wind gusts.” Eng. Struct.
19 6: 262–267.
or

20 Yeatts, B. B., and K. C. Mehta. 1993. “Field study of internal pressures.” In Proc., 7th U.S. Nat.
nf

21 Conf. on Wind Engineering, vol. 2, 889–897.


22 Zhou, Y., and A. Kareem. 2001a. “Gust loading factor: New model.” J. Struct. Eng. 127 (2):
rI

23 168–175.
Fo

24 Zhou, Y., and A. Kareem. 2001b. “Equivalent static lateral forces on buildings under seismic and
25 wind effects.” J. Wind Eng. 89: 605–608.
26 Zhou, Y., T. Kijewski, and A. Kareem. 2002. “Along-wind load effects on tall buildings:
27 Comparative study of major international codes and standards.” J. Struct. Eng. 128 (6):
28 788–796.
29 Zilitinkevich, S. S., and I. N. Esau. 2002. “On integral measures of the neutral barotropic
30 planetary boundary layer.” Boundary-Layer Meteorology 104 (3): 371-379.
31

88
1 OTHER REFERENCES (NOT CITED)
2 ASCE. 1987. “Wind tunnel model studies of buildings and structures.” Manuals and Reports of
3 Engineering Practice No. 67. Reston, VA: ASCE.
4 ASTM International. 2006. “Standard specification for rigid poly(vinyl chloride) (PVC) siding.”
5 ASTM D3679-06a. West Conshohocken, PA: ASTM.
6 ASTM. 2007. “Standard test method for wind resistance of sealed asphalt shingles (uplift
7 force/uplift resistance method).” ASTM D7158-07. West Conshohocken, PA: ASTM.
8 Cook, N. (1985). The designer’s guide to wind loading of building structures, Part I:

LY
9 Background, damage survey, wind data and structural classification. London: Building
10 Research Establishment and Butterworths.

N
11 ESDU. 1990. Characteristics of atmospheric turbulence near the ground. Part II: Single point
12 data for strong winds (neutral atmosphere). ESDU 85020. London: ESDU.

O
13 Georgiou, P. N. 1985. “Design wind speeds in tropical cyclone regions.” PhD dissertation,
14
n
University of Western Ontario, Dept. of Civil Engineering.
io
15 Georgiou, P. N., A. G. Davenport, and B. J. Vickery. 1983. “Design wind speeds in regions
16 dominated by tropical cyclones.” J. Wind Eng. Indust. Aerodyn. 13: 139–152.
at

17 Holmes, J. D. 2001. Wind loads on structures. New York: SPON Press / Taylor & Francis.
m

18 Isyumov, N. 1982. “The aeroelastic modeling of tall buildings.” In Proc., Int. Workshop on Wind
19 Tunnel Modeling Criteria and Techniques in Civil Engineering Applications, T.
or

20 Reinhold, ed. New York: Cambridge University Press, 373–407.


nf

21 Jeary, A. P., and B. R. Ellis. 1983. “On predicting the response of tall buildings to wind
22 excitation.” J. Wind Eng. Indust. Aerodyn. 13: 173–182.
rI

23 Kala, S., T. Stathopoulos, and K. Kumar. 2008. “Wind loads on rainscreen walls: Boundary-
Fo

24 layer wind tunnel experiments.” J. Wind Eng. Indust. Aerodyn. 96 (6-7): 1058–1073.
25 Kareem, A., and C. W. Smith. 1994. “Performance of off-shore platforms in Hurricane Andrew.”
26 In Hurricanes of 1992: Lessons learned and implications for the future, R. A. Cook and
27 M. Soltani, eds. New York: ASCE, 577–586.
28 Kijewski, T., and Kareem, A. 1998. “Dynamic wind effects: A comparative study of provisions
29 in codes and standards with wind tunnel data.” J. Wind and Struct. 1 (1): 77–109.
30 Krayer, W. R., and R. D. Marshall. 1992. “Gust factors applied to hurricane winds.” Bull. Am.
31 Meteorol. Soc. 73: 613–617.

89
1 Kwon, D-K., and A. Kareem. (2013). “A multiple database-enabled design module with
2 embedded features of international codes and standards.” International Journal of High-
3 Rise Buildings (Council on Tall Buildings and Urban Habitat): 2 (3): 257–269.
4 Kwon, K. K., S. M. J. Spence, and A. Kareem. 2014, “A cyberbased data-enabled design
5 framework for high-rise buildings driven by synchronously measured surface pressures.”
6 Adv. Eng. Softw. 77: 13–27.
7 Liu, H. 1999. Wind engineering: A handbook for structural engineers. New York: Prentice Hall.
8 MacDonald, P. A., K. C. S. Kwok, and J. H. Holmes. 1986. “Wind loads on isolated circular

LY
9 storage bins, silos and tanks: Point pressure measurements.” Research Report R529.
10 School of Civil and Mining Engineering, University of Sydney, Australia.

N
11 Myer, M. F., and G. F. White. 2003. “Communicating damage potentials and minimizing
12 hurricane damage.” In Hurricane! Coping with disaster: Progress and challenges since

O
13 Galveston, 1900, R. Simpson, ed. Washington, DC: American Geophysical Union.
14
n
Perry, D. C., N. Stubbs, and C. W. Graham. 1993. “Responsibility of architectural and
io
15 engineering communities in reducing risks to life, property and economic loss from
16 hurricanes.” In Hurricanes of 1992: Lessons learned and implications for the future.
at

17 Reston, VE: ASCE Press.


m

18 Solari, G., and A. Kareem. 1998. “On the formulation of ASCE 7-95 gust effect factor.” J. Wind
19 Eng. Indust. Aerodyn. 77-78: 673–684.
or

20 Standards Australia. (1989). Australian standard SAA loading code, Part 2: Wind loads. North
nf

21 Sydney: Standards House.


22 Stathopoulos, T., D. Surry, and A. G. Davenport. 1980. “A simplified model of wind pressure
rI

23 coefficients for low-rise buildings.” In Proc., 4th Colloquium on Industrial


Fo

24 Aerodynamics, Part 1, Gesellschaft der Freunde der Fachhochschule Aachen, 17–31.


25 Stubbs, N., and A. C. Boissonnade. 1993) “A damage simulation model for building contents in
26 a hurricane environment.” In Proc., 7th US Nat. Conf. on Wind Engineering, vol. 2, 759–
27 771.
28 Vickery, B. J., A. G. Davenport, and D. Surry. 1984. “Internal pressures on low-rise buildings.”
29 In Proc., 4th Canadian Workshop on Wind Engineering, Canadian Wind Engineering
30 Association, Ottawa.

90
1 Vickery, P. J., and P. F. Skerlj. 1998. “On the elimination of exposure D along the hurricane
2 coastline in ASCE-7.” Report for Andersen Corporation by Applied Research Associates,
3 Project 4667. Raleigh, NC: Applied Research Associates.
4 Vickery, P. J., P. F. Skerlj, and L. A. Twisdale. (2000). “Simulation of hurricane risk in the U.S.
5 using empirical track model.” J. Struct. Eng. 126 (10): 1222–1237.
6 Vickery, P. J., and L. A. Twisdale. 1995a. “Prediction of hurricane wind speeds in the United
7 States.” J. Struct. Eng. 121 (11): 1691–1699.
8 Vickery, P. J., and L. A. Twisdale. 1995b. “Wind-field and filling models for hurricane wind-

LY
9 speed predictions.” J. Struct. Eng. 121 (11): 1700–1709.
10 Womble, J. A., B. B. Yeatts, and K. C. Mehta. 1995. “Internal wind pressures in a full and small

N
11 scale building.” In Proc., 9th Int. Conf. on Wind Engineering. New Delhi: Wiley Eastern.
12 Zhou, Y., and A. Kareem, A. 2003. “Aeroelastic Balance.” Journal of Engineering Mechanics

O
13 129 (3): 283–292.
14
n
Zhou, Y., A. Kareem, and M. Gu. 2000. “Equivalent static buffeting loads on structures.” J.
io
15 Struct. Eng. 126 (8): 989–992.
16 Zhou, Y., T. Kijewski, and A. Kareem. 2003. “Aerodynamic loads on tall buildings: Interactive
at

17 database.” J. Struct. Eng. 129 (3): 394–404.


m

18
or
nf
rI
Fo

91

You might also like