You are on page 1of 129

National Frame Builders Association

Post-Frame Building Design Manual

Chapter 1: INTRODUCTION TO POST-FRAME BUILDINGS

1.1 General
1.1.1 Main Characteristics. Post-frame buildings are structurally efficient
buildings composed of main members such as posts and trusses and secondary
components such as purlins, girts, bracing and sheathing Snow and wind loads are
transferred from the sheathing to the secondary members. Loads are transferred to
the ground through the posts that typically are embedded in the ground or surface-
mounted to a concrete

or masonry foundation. Figure 1.1 illustrates the structural components of a post-


frame building.
1.1.2 Use. Post-frame construction is wellsuited for many commercial, industrial,
agricultural and residential applications. Post-frame offers unique advantages in
terms of design and construction flexibility and structural efficiency. For these
reasons, post-frame construction has experienced rapid growth, particularly in
nonagricultural applications.

Ridge cap Purlin

Roof cladding

Truss

Pressure preservative treated post

Concrete footing

Wall girt

Wall cladding Doorway Pressure preservative treated splash board

Figure 1.1. Simplified diagram of a post-frame building. Some components such as


permanent roof truss bracing and interior finishes are not shown.

1-1

National Frame Builders Association

Post-Frame Building Design Manual

1.2 Evolution
1.2.1 The concept of pole-type structures is not new. Archeological evidence exists
in abundance that pole buildings have been used for human housing for thousands of
years. In America, pole buildings began appearing on farms in the 19th century
(Norum, 1967).
1.2.2 Pole-type construction resurfaced in 1930 when Mr. H. Howard Doane introduced
the "modern pole barn" as an economical alternative to conventional barns (Knight,
1989). Mr. Doane was the founder of Doane's Agricultural Service, a firm
specializing in managing farms for absentee owners. These early pole barns were
constructed with red cedar poles that were naturally resistant to decay, trusses
spaced 2 ft oncenter, 1-inch nominal purlins and galvanized steel sheathing.
In the 1940s, pole barn construction was refined by using creosote preservative-
treated sawn posts, wider truss and purlin spacings, and improved steel sheathing.
Mr. Bernon G. Perkins, an employee of Doane's, is credited for many of the
refinements to Doane's original pole barn. In 1949, Mr. Perkins applied for the
first patent on the pole building concept through Doane's Agricultural Service, and
the patent was issued in 1953. Rather than protecting their patent, they publicized
the concept and encouraged its use throughout the world. In 1995, the post-frame
building concept was recognized as an Historic Agricultural Engineering Landmark by
the American Society of Agricultural Engineers.
1.2.3 In the past two decades, post-frame construction has been further enhanced by
the developments of metal-plate connected wood trusses, nail- and glue-laminated
posts, highstrength steel sheathing, fasteners and diaphragm design methods.
Composites such as laminated posts and structural composite lumber offer advantages
of superior strength and stiffness, dimensional stability, and they can be obtained
in a variety of sizes and pressure preservative treatments. Developments in
metalplate connected wood truss technology allow clear spans of over 80 feet.
Design procedures were introduced in the early 1980s to more accurately account for
the effect of diaphragm ac-

tion on post and foundation design (Knight, 1990). New roof panel constructions
using highstrength steel and customized screw fasteners have dramatically improved
diaphragm stiffness and strength.
1.3 Advantages
1.3.1 Reliability. Outstanding structural performance of post-frame buildings under
adverse conditions such as hurricanes is welldocumented. Professor Gurfinkel, in
his wood engineering textbook, cites superior performance of post-frame buildings
over conventional construction during hurricane Camille in 1969 (Gurfinkel, 1981).
Harmon et. al (1992) reported that post-frame buildings constructed according to
engineered plans generally withstood hurricane Hugo (wind gusts measured at 109
mph). Since post-frame buildings are relatively light weight, seismic forces do not
control the design unless significant additional dead loads are applied to the
structure (Faherty and Williamson, 1989; Taylor, 1996).
1.3.2 Economy. Significant savings can be obtained with post-frame construction in
terms of materials, labor, construction time, equipment and building maintenance.
For example, postframe buildings require less extensive foundations than other
building types because the wall sections between the posts are non-load bearing.
Embedded post foundations commonly used in post-frame require less concrete, heavy
equipment, labor, and construction time than conventional perimeter foundations.
Additionally, embedded post foundations are better-suited for wintertime
construction.
1.3.3 Versatility. Post-frame construction facilitates design flexibility. Posts
can be embedded into the ground or surface-mounted to a concrete foundation. Steel
sheathing can be replaced with wood siding, brick veneer, and conventional roofing
materials, to satisfy the appearance and service requirements of the customer. One-
hour fire-rated wall and roof/ceiling constructions have been developed for wood
framed assemblies. Exposed glued-laminated and solid-sawn timbers can be
substituted for trusses made from dimension lumber to achieve desired architectural
effects.

1-2

National Frame Builders Association

Post-Frame Building Design Manual

1.4 Industry Profile


1.4.1 Post-frame construction has experienced tremendous growth since World War II.
This growth was fueled by the abundant supplies of steel and pressure preservative-
treated wood, together with the need for low-cost structures. In the 1950s and
1960s, the pole barn industry was characterized by large numbers of independent
builders (Knight, 1989). During this time, pole builders were expanding from their
traditional agricultural base into other construction markets. This expansion into
code-enforced construction required rigorous documentation of engineering designs
and more involvement in the building code arena.
1.4.2 NFBA. Approximately 20 builders met in 1969 to discuss challenges facing the
postframe building industry. The group voted in favor of forming the National Frame
Builders Association (NFBA). The NFBA became incorporated in 1971 and the first
national headquarters was established in Chicago, Illinois. Today, the National
Frame Builders Association is headquartered in Lawrence, Kansas and includes over
300 contractors and suppliers, with regional branches throughout the U.S. In
addition, a Canadian Division of NFBA was created in 1984.
1.4.3 The post-frame industry has become one of the fastest growing segments of the
total construction industry. Based on light-gauge steel sales, post-frame industry
revenues are estimated to be from 2 to 2.5 billion dollars in 1990.
1.5 Terminology
AF&PA: American Forest & Paper Association (formerly National Forest Products
Association).
AITC: American Institute of Timber Construction.
ALSC: American Lumber Standard Committee.
ANSI: American National Standards Institute
APA: The Engineered Wood Association (formerly the American Plywood Association)

ASAE: The Society for engineering in agricultural, food, and biological systems
(formerly American Society of Agricultural Engineers).
Anchor Bolts: Bolts used to anchor structural members to a foundation. Commonly
used in post-frame construction to anchor posts to the concrete foundation.
ASCE: American Society of Civil Engineers.
AWC: American Wood Council. The wood products division of the American Forest &
Paper Association (AF&PA).
AWPB: American Wood Preservers Bureau.
Bay: The area between adjacent primary frames in a building. In a post-frame
building, a bay is the area between adjacent post-frames.
Bearing Height: Vertical distance between a pre-defined baseline (generally the
grade line) and the bearing point of a component.
Bearing Point: The point at which a component is supported.
Board: Wood member less than two (2) nominal inches in thickness and one (1) or
more nominal inches in width.
Board-Foot (BF): A measure of lumber volume based on nominal dimensions. To
calculate the number of board-feet in a piece of lumber, multiply nominal width in
inches by nominal thickness in inches times length in feet and divide by 12.
BOCA: Building Officials & Code Administrators International, Inc. The organization
responsible for maintaining and publishing the National Building Code.
Bottom Chord: An inclined or horizontal member that establishes the bottom of a
truss.
Bottom Plank: See Splashboard.
Butt Joint: The interface at which the ends of two members meet in a square cut
joint.

1-3

National Frame Builders Association

Post-Frame Building Design Manual

Camber: A predetermined curvature designed into a structural member to offset the


anticipated deflection when loads are applied.
Check: Separation of the wood that usually extends across the annual growth rings
(i.e., a split perpendicular-to-growth rings). Commonly results from stresses that
build up in wood during seasoning.
Cladding: The exterior and interior coverings fastened to the wood framing.
Clear Height: Vertical distance between the finished floor and the lowest part of a
truss, rafter, or girder.
Collars: Components that increase the bearing area of portions of the post
foundation, and thus increase lateral and vertical resistance.
Components and Cladding: Elements of the building envelope that do not qualify as
part of the main wind-force resisting system. In postframe buildings, this
generally includes individual purlins and girts, and cladding.
Diaphragm: A structural assembly comprised of structural sheathing (e.g., plywood,
metal cladding) that is fastened to wood or metal framing in such a manner the
entire assembly is capable of transferring in-plane shear forces.
Diaphragm Action: The transfer of load by a diaphragm.
Diaphragm Design: Design of roof and ceiling diaphragm(s), wall diaphragms
(shearwalls), primary and secondary framing members, component connections, and
foundation anchorages for the purpose of transferring lateral (e.g., wind) loads to
the foundation structure.
Dimension Lumber: Wood members from two (2) nominal inches to but not including
five (5) nominal inches in thickness, and 2 or more nominal inches in width.
Eave: The part of a roof that projects over the sidewalls. In the absence of an
overhang, the eave is the line along the sidewall formed by the intersection of the
wall and roof planes.

Fascia: Flat surface (or covering) located at the outer end of a roof overhang or
cantilever end.
Flashing: Sheet metal or plastic components used at major breaks and/or openings in
walls and roofs to insure weather-tightness in a structure.
Footing: Support base for a post or foundation wall that distributes load over a
greater soil area.
Frame Spacing: Horizontal distance between post-frames (see post-frame and post-
frame building). In the absence of posts, the frame spacing is generally equated to
the distance between adjacent trusses (or rafters). Frame spacing may vary within a
building.
Gable: Triangular portion of the endwall of a building directly under the sloping
roof and above the eave line.
Gable Roof: Roof with one slope on each side. Each slope is of equal pitch.
Gambrel Roof: Roof with two slopes on each side. The pitch of the lower slope is
greater than that of the upper slope.
Girder: A large, generally horizontal, beam. Commonly used in post-frame buildings
to support trusses whose bearing points do not coincide with a post.
Girt: A secondary framing member that is attached (generally at a right angle) to
posts. Girts laterally support posts and transfer load between wall cladding and
posts.
Glued-Laminated Timber: Any member comprising an assembly of laminations of lumber
in which the grain of all laminations is approximately parallel longitudinally, in
which the laminations are bonded with adhesives.
Grade Girt: See Splashboard.
Grade Line (grade level): The line of intersection between the building exterior
and the top of the soil, gravel, and/or pavement in contact with the building
exterior. For post-frame building

1-4

National Frame Builders Association

Post-Frame Building Design Manual

design, the grade line is generally assumed to be no lower than the lower edge of
the splashboard.
Header: A structural framing member that supports the ends of structural framing
members that have been cut short by a floor, wall, ceiling, or roof opening.
Hip Roof: Roof which rises by inclined planes from all four sides of a building.
IBC: International Building Code.
ICBO: International Conference of Building Officials. The organization responsible
for maintaining and publishing the Uniform Building Code.
Knee Brace: Inclined structural framing member connected on one end to a
post/column and on the other end to a truss/rafter.
Laminated Assembly: A structural member comprised of dimension lumber fastened
together with mechanical fasteners and/or adhesive. Horizontally- and vertically-
laminated assemblies are primarily designed to resist bending loads applied
perpendicular and parallel to the wide face of the lumber, respectively.
Laminated Veneer Lumber (LVL) A structural composite lumber assembly manufactured
by gluing together wood veneer sheets. Each veneer is orientated with its wood
fibers parallel to the length of the member. Individual veneer thickness does not
exceed 0.25 inches.
Loads: Forces or other actions that arise on structural systems from the weight of
all permanent construction, occupants and their possessions, environmental effects,
differential settlement, and restrained dimensional changes.
Dead Loads: Gravity loads due to the weight of permanent structural and
nonstructural components of the building, such as wood framing, cladding, and fixed
service equipment.
Live Loads: Loads superimposed by the construction, use and occupancy of the
building, not including wind, snow, seismic or dead loads.

Seismic Load: Lateral load acting in the horizontal direction on a structure due to
the action of earthquakes.
Snow Load: A load imposed on a structure due to accumulated snow.
Wind Loads: Loads caused by the wind blowing from any direction.
Lumber Grade: The classification of lumber in regard to strength and utility in
accordance with the grading rules of an approved (ALSC accredited) lumber grading
agency.
LVL: see Laminated Veneer Lumber.
Main Wind-Force Resisting System: An assemblage of structural elements assigned to
provide support and stability for the overall structure. Main wind-force resisting
systems in post-frame buildings include the individual postframes, diaphragms and
shearwall
Manufactured Component. A component that is assembled in a manufacturing facility.
The wood trusses and laminated columns used in post-frame buildings are generally
manufactured components.
MBMA: Metal Building Manufacturers Association.
NDS: National Design Specification for Wood Construction. Published by AF&PA.
Mechanically Laminated Assembly: A laminated assembly in which wood laminations
have been joined together with nails, bolts and/or other mechanical fasteners.
Metal Cladding: Metal exterior and interior coverings, usually cold-formed aluminum
or steel sheet, fastened to the structural framing.
NFBA: National Frame Builders Association.
NFPA: National Fire Protection Association
Nominal size: The named size of a member, usually different than actual size (as
with lumber).

1-5

National Frame Builders Association

Post-Frame Building Design Manual

Orientated Strand Board (OSB): Structural wood panels manufactured from


reconstituted, mechanically oriented wood strands bonded with resins under heat and
pressure.
Orientated Strand Lumber (OSL): Structural composite lumber (SCL) manufactured from
mechanically oriented wood strands bonded with resins under heat and pressure. Also
known as laminated strand lumber (LSL)
OSB: See Orientated Strand Board.
Parallel Strand Lumber (PSL): Structural composite lumber (SCL) manufactured by
cutting 1/8-1/10 inch thick wood veneers into narrow wood strands, and then gluing
and pressing the strands together. Individual strands are up to 8 feet in length.
Prior to pressing, strands are oriented so that they are parallel to the length of
the member.
Pennyweight: A measure of nail length, abbreviated by the letter d.
Plywood: A built-up panel of laminated wood veneers. The grain orientation of
adjacent veneers are typically 90 degrees to each other.
Pole: A round, unsawn, naturally tapered post.
Post: A rectangular member generally uniform in cross section along its length.
Post may be sawn or laminated dimension lumber. Commonly used in post-frame
construction to transfer loads from main roof beams, trusses or rafters to the
foundation.
Post Embedment Depth: Vertical distance between the bottom of a post and the lower
edge of the splashboard.
Post Foundation: The embedded portion of a structural post and any footing and/or
attached collar.
Post Foundation Depth: Vertical distance between the bottom of a post foundation
and the lower edge of the splashboard.
Post-Frame: A structural building frame consisting of a wood roof truss or rafters
connected to vertical timber columns or sidewall posts.

Post-Frame Building: A building system whose primary framing system is principally


comprised of post-frames.
Post Height: The length of the non-embedded portion of a post.
Pressure Preservative Treated (PPT) Wood: Wood pressure-impregnated with an
approved preservative chemical under approved treatment and quality control
procedures.
Primary Framing: The main structural framing members in a building. The primary
framing members in a post-frame building include the columns, trusses/rafters, and
any girders that transfer load between trusses/rafters and columns.
PSL: See Parallel Strand Lumber.
Purlin: A secondary framing member that is attached (generally at a right angle) to
rafters/ trusses. Purlins laterally support rafters and trusses and transfer load
between exterior cladding and rafters/trusses.
Rafter: A sloping roof framing member.
Rake: The part of a roof that projects over the endwalls. In the absence of an
overhang, the rake is the line along the endwall formed by the intersection of the
wall and roof planes.
Ridge: Highest point on the roof of a building which describes a horizontal line
running the length of the building.
Ring Shank Nail: See threaded nail.
Roof Overhang: Roof extension beyond the endwall/sidewall of a building.
Roof Slope: The angle that a roof surface makes with the horizontal. Usually
expressed in units of vertical rise to 12 units of horizontal run.
SBC: Standard Building Code (see SBCCI).
SBCCI: Southern Building Code Congress International, Inc. The organization
responsible for maintaining and publishing the Standard Building Code.

1-6

National Frame Builders Association

Post-Frame Building Design Manual

Secondary Framing: Structural framing members that are used to (1) transfer load
between exterior cladding and primary framing members, and/or (2) laterally brace
primary framing members. The secondary framing members in a post-frame building
include the girts, purlins and any structural wood bracing.
Self-Drilling Screw: A screw fastener that combines the functions of drilling and
tapping (thread forming). Generally used when one or more of the components to be
fastened is metal with a thickness greater than 0.03 inches
Self-Piercing Screw: A self-tapping (thread forming) screw fastener that does not
require a pre-drilled hole. Differs from a self-drilling screw in that no material
is removed during screw installation. Used to connect light-gage metal, wood,
gypsum wallboard and other "soft" materials.
SFPA: Southern Forest Products Association
Shake: Separation of annual growth rings in wood (splitting parallel-to-growth
rings). Usually considered to have occurred in the standing tree or during felling.
Shearwall: A vertical diaphragm in a structural framing system. A shearwall is any
endwall, sidewall, or intermediate wall capable of transferring in-plane shear
forces.
Siphon Break: A small groove to arrest the capillary action of two adjacent
surfaces.
Soffit: The underside covering of roof overhangs.
Soil Pressure: Load per unit area that the foundation of a structure exerts on the
soil.
Span: Horizontal distance between two points.
Clear Span: Clear distance between adjacent supports of a horizontal or inclined
member. Horizontal distance between the facing surfaces of adjacent supports.
Effective Span: Horizontal distance from center-of-required-bearing-width to
centerof-required-bearing-width, or the "clear

span" for rafters and joists in conventional construction.


Out-To-Out Span: Horizontal distance between the outer faces of supports. Commonly
used in specifying metal-plateconnected wood trusses.
Overall Span: Total horizontal length of an installed horizontal or inclined
member.
SPIB: Southern Pine Inspection Bureau.
Skirtboard: See Splashboard.
Splashboard: A preservative treated member located at grade that functions as the
bottom girt. Also referred to as a skirtboard, splash plank, bottom plank, and
grade girt.
Splash Plank: See Splashboard.
Stitch (or Seam) Fasteners: Fasteners used to connect two adjacent pieces of metal
cladding, and thereby adding shear continuity between sheets.
Structural Composite Lumber (SCL): Reconstituted wood products comprised of several
laminations or wood strands held together with an adhesive, with fibers primarily
oriented along the length of the member. Examples include LVL and PSL.
Threaded Nail: A type of nail with either annual or helical threads in the shank.
Threaded nails generally are made from hardened steel and have smaller diameters
than common nails of similar length.
Timber: Wood members five or more nominal inches in the least dimension.
Top Chord: An inclined or horizontal member that establishes the top of a truss.
TPI: Truss Plate Institute.
Truss: An engineered structural component, assembled from wood members, metal
connector plates and/or other mechanical fasteners, designed to carry its own
weight and superimposed design loads. The truss members form a

1-7

National Frame Builders Association

Post-Frame Building Design Manual

semi-rigid structural framework and are assembled such that the members form
triangles.
UBC: Uniform Building Code (see ICBO).
Wane: Bark, or lack of wood from any cause, on the edge or corner of a piece.
Warp: Any variation from a true plane surface. Warp includes bow, crook, cup, and
twist, or any combination thereof.
Bow: Deviation, in a direction perpendicular to the wide face, from a straight line
drawn between the ends of a piece of lumber.
Crook: Deviation, in a direction perpendicular to the narrow edge, from a straight
line drawn between the ends of a piece of lumber.
Cup: Deviation, in the wide face of a piece of lumber, from a straight line drawn
from edge to edge of the piece.
Twist: A curl or spiral of a piece of lumber along its length. Measured by laying
lumber on a flat surface such that three corners contact the surface. The amount of
twist is equal to the distance between the flat surface and the corner not
contacting the surface.
WCLIB: West Coast Lumber Inspection Bureau
Web: Structural member that joins the top and bottom chords of a truss. Web members
form the triangular patterns typical of most trusses.
WTCA: Wood Truss Council of America.
WWPA: Western Wood Products Association.
1.6 References
Faherty, K.F. and T.G. Williamson. 1989. Wood Engineering and Construction
Handbook. McGraw-Hill Publishing Company, New York, NY.
Gurfinkel, G. 1981. Wood Engineering (2nd Ed.). Kendall/Hunt Publishing Company,
Dubuque, Iowa.

Harmon, J.D., G.R. Grandle and C.L. Barth. 1992. Effects of hurricane Hugo on
agricultural structures. Applied Engineering in Agriculture 8(1):93-96.
Knight, J.T. 1989. A brief look back. Frame Building Professional 1(1):38-43.
Knight, J.T. 1990. Diaphragm design - technology driven by necessity. Frame
Building Professional 1(5):16,44-46.
Norum, W.A. 1967. Pole buildings go modern. Journal of the Structural Division,
ASCE, Vol. 93, No.ST2, Proc. Paper 5169, April, pp.47-56.
Taylor, S.E. 1996. Earthquake considerations in post-frame building design. Frame
Building News 8(3):42-49.

1-8

National Frame Builders Association

Post-Frame Building Design Manual

Chapter 2: BUILDING CODES, DESIGN SPECIFICATIONS AND ZONING REGULATIONS

2.1 Introduction
2.1.1 Definition. A building code is a legal document that helps ensure public
health and welfare by establishing minimum standards for design, construction,
quality of materials, use and occupancy, location and maintenance of all buildings
and structures.
2.1.2 Model Versus Active Codes. A model code is a code that is written for general
use (i.e., a code that is not written for use by a specific state, county, town,
village, company or individual). An active code is a model or specially written
code that has been adopted and is enforced by a regulatory agency such as a state
or local government. It follows that in a given jurisdiction, acceptance of a model
building code is voluntary until the model code becomes part of the active code in
the jurisdiction.
2.1.3 Active Code Variations. The content and administration of active building
codes varies not only between states, but frequently between municipalities within
a state. Some states have established a hierarchy structure of state, county and
township/village/city building codes. In this situation, more localized governing
areas can modify the state (or county) codes, provided the changes result in more
strict provisions.
Despite local differences in content and administration, most active building codes
share the common trait of regulating components of construction based on building
occupancy and use.
2.2 Major Model Building Codes
2.2.1 Current Codes. There are currently three primary model building codes in the
United States. These are the Uniform Building Code (UBC) published by the
International Congress of Building Officials, the National Building Code published
by the Building Officials and Code Administrators International (BOCA) and the
Standard Building Code published by the Southern Building Code Congress
International

(SBCCI). These model building codes are commonly referred to as the UBC, BOCA and
the Southern Building Code, respectively.
2.2.2 Adoption. Most states have adopted (and enforce) all or a major portion of
one of the three model building codes. As shown in figure 2.1, western states have
adopted the UBC, northeastern states the BOCA code, and states in the southwest the
Southern Building Code.
2.2.3 Development. Model building codes are consensus documents continually studied
and annually revised by building officials, industry representatives and other
interested parties.
2.2.4 International Building Code. On December 9, 1994, the three model building
code agencies (BOCA, ICBO and SBCCI) created the International Code Council (ICC).
The ICC was established in response to technical disparities among the three major
model codes. Since its founding, the ICC has worked to create a single model
building code for the U.S. This code, which is entitled the International Building
Code is now complete and will replace the three model codes over the next couple
years. With all states adopting the same model code, it will be less difficult for
building designers to work in different regions of the country.
2.3 Building Classification
2.3.1 General. Building codes give criteria for classifying buildings based on: (1)
use or occupancy, and (2) type of construction.
2.3.2 Occupancy Classifications. Occupancy classifications include assembly,
business, educational, factory and industrial, high-hazard, institutional,
mercantile, residential and storage. Occupancy classifications have requirements on
the number of occupants and building separation, height and area. Other limits
exist, for example on lighting, ventilation, sanitation, fire

2-1

National Frame Builders Association

Post-Frame Building Design Manual

Uniform Building Code (ICBO)

National Building Code (BOCA)

Standard Building Code (SBCCI)

Figure 2.1. Approximate areas of model building code influence. Wisconsin and New
York building codes are developed by their respective state code agencies and are
not necessarily influenced by current model codes.

protection and exiting, depending on the specific classification and building code.
2.3.2 Types of Construction. Classification by type of construction is primarily
based on the fire resistance ratings of the walls, partitions, structural elements,
floors, ceilings, roofs and exits. Specific requirements vary somewhat between
model building codes.
There are two primary source documents for determining the fire resistance of
assemblies: the Fire Resistance Design Manual, published by the Gypsum Association,
and the Fire Resistance Directory, published by Underwriters Laboratories, Inc.
The fire resistance of wood framed assemblies can generally be increased by using
fire retardant treated (FRT) wood or larger wood members. Codes allow FRT wood to
be used in cer-

tain areas of noncombustible construction. The superior fire resistance of large


timber members is recognized by the codes with the inclusion of a "heavy timber"
classification. To qualify for heavy timber construction, nominal dimensions of
timber columns must be at least 6- by 8inches and primary beams shall have nominal
width and depth of at least 6- by 10-inches.
2.3.2.1 NFBA Sponsored Fire Test. In January of 1990, the National Frame Builders
Association had Warnick Hersey International, Inc., conduct a one-hour fire
endurance test on the exterior wall shown in figure 2.2. The wall met all
requirements for a one-hour rating as prescribed in ASTM E119-88. The wall
sustained an applied load of 10,400 lbf per column throughout the test. Copies of
the fire test report can be obtained from NFBA.

2-2

National Frame Builders Association

Post-Frame Building Design Manual

Attach metal cladding 12 in. o.c. with 1.5 in. hex head screws with neoprene
washers

Metal cladding 29 gage

Unexposed nominal 2by 4-inch nailers 24 in. o.c.

Section A-A

Gold Bond 5/8 in. Fireshield G Type X, attached with 1-7/8 in. cement coated nails
(0.0195 in. shank, 1/4 head, 7 in. o.c.)

Fire side nailers, nominal 2- by 4-inches


24 in. o.c.

A
10 ft

Nominal 2- by 4-inch nailers, 24 in. o.c.

FIRE SIDE

4-1/16- by 5-1/4-inch glue-laminated column

3- by 24- by 48-inch
mineral wool, attach with 3 in. square cap nails (3
per 48 in. width)
Nominal 2- by 2-inch blocking between nailers (nailed to nominal 2- by
6-inch edge blocks)
Nail-laminated column fabricated from 3 nominal 2- by 6-inch No. 2 KD19 SP members
Nominal 2- by 4-inch blocking attached to column

Section B-B

1 ft

B
8 ft 1 ft

Figure 2.2. Construction details for exterior wall that obtained a one-hour fire
endurance rating during a January 1990 test conducted for the National Frame
Builders Association by Warnock Hersey International, Inc. Details of the test are
available from NFBA upon request.

2.4 Specifications and Standards


2.4.1 General. Design of buildings is covered in the model building codes either by
direct provisions or by reference to approved engineering specifications and
standards. Engineering specifications and standards provide criteria and data
needed for load calculation, design, testing and material selection. They are based
on the best available information and engineering judgment.
2.4.2 Wood Design Specifications. The tech

nical literature for wood design and construction is somewhat fragmented. New
design specifications and standards are continually under development, and existing
documents are periodically revised. Keeping abreast of this literature requires a
determined effort on the part of the design professional. To assist in this effort,
Table 2.1 gives a partial list of engineering design specifications, standards and
other technical references specifically related to post-frame construction. The
reader is encouraged to maintain communication with the organizations isted in
Table 2.1 concerning new and revised publications.

2-3

National Frame Builders Association

Post-Frame Building Design Manual

Of the documents listed in Table 2.1, the primary engineering design specification
cited by the model building codes for wood construction is the National Design
Specification for Wood Construction (NDS), published by the American Forest &
Paper Association (AF&PA). The NDS was first issued in 1944 and in 1992 it became a
consensus standard through the American National Standards Institute (ANSI).
2.5 Zoning Regulations
2.5.1 General. Zoning laws are established tocontrol construction activities and
regulate land use, in terms of types of occupancy, building

height, and density of population and activity. Zoning laws may also dictate
building appearance and location on property, parking signs, drainage, handicap
accessibility, flood control and landscaping. Typically land is zoned for
residential, commercial, industrial or agricultural uses.
2.5.2 Development and Enforcement. Zoning laws are developed by municipalities.
They (and building codes) are principally enforced by the granting of building
permits and inspection of construction work in progress. Certificates of occupancy
are issued when completed buildings satisfy all regulations.

Table 2.1. Partial list of technical references related to post-frame building


design and construction

Organization & Address

Publications

AF&PA American Forest & Paper Association 1111 19th Street, N.W., Suite 800
Washington, D.C. 20036 http://www.awc.org/

Allowable stress design (ASD) manual for engineered wood construction


National design specification (NDS) for wood construction NDS commentary Design
values for wood construction (NDS supplement) Load and resistance factor design
(LRFD) manual for engi-
neered wood construction Wood frame construction manual (WFCM) for one-and two-
family dwellings Span tables for joists and rafters

AITC American Inst. of Timber Construction 7012 S. Revere Parkway, Suite 140
Englewood, CO 80112

Timber construction manual

ANSI American National Standards Institute 11 West 42nd Street New York, NY 10036
http://www.ansi.org/

ANSI/AF&PA National design specification for wood construction (see AF&PA)


ANSI Standard A190 structural glued laminated

2-4

National Frame Builders Association

Post-Frame Building Design Manual

Table 2.1. Partial list of technical references related to post-frame building


design and construction

Organization & Address

Publications

APA The Engineered Wood Association P.O. Box 11700 7011 South 19th Street Tacoma,
WA 98411 http://www.apawood.org/

APA design/construction guide; residential and commercial Plywood design


specification (PDS) Diaphragms and shear walls Performance standard for APA EWS I-
joists Panel handbook & grade glossary

ASAE 2950 Niles Road St. Joseph, MI 49085-9659 http://asae.org/

ASAE EP288 Agricultural building snow and wind loads ASAE EP484.2 Diaphragm design
of metal-clad, wood-frame
rectangular buildings ASAE EP486 Post and pole foundation design ASAE EP558 Load
tests for metal-clad, wood-frame dia-
phragms ANSI/ASAE EP559 Design requirements and bending proper-
ties for mechanically laminated columns

ASCE American Society of Civil Engineers 1801 Alexander Bell Drive Reston, Virginia
20191-4400 http://www.asce.org/
ASCE Standard 7 Minimum Design Loads for Buildings and Other Structures
Standard for load and resistance factor design (LRFD) for engineered wood
construction
Guide to the use of the wind load provisions of ASCE 7-95

AWPA American Wood Preservers Assoc. P.O. Box 5690 Granbury, TX 76049

Standard C2 lumber, timbers, bridge ties and mine ties - preservative treatment by
pressure processes
Standard C15 wood for commercial-residential construction preservative treatment by
pressure processes
Standard C16 wood used on farms - preservative treatment by pressure processes
Standard C23 round poles and posts used in building construction - preservative
treatment by pressure processes
Standard M4 standard for the care of preservative-treated wood products

AWPI American Wood Preservers Institute 2750 Prosperity Avenue, Suite 550 Fairfax,
Virginia 22031-4312 http://www.awpi.org/

Answers to often-asked questions about treated wood Management of used treated wood
products booklet

Gypsum Association 810 First St., NE, #510 Washington DC, 20002
http://www.gypsum.org/

Fire resistance design manual GA-600 Design data - gypsum board GA-530

2-5

National Frame Builders Association

Post-Frame Building Design Manual

Table 2.1. Partial list of technical references related to post-frame building


design and construction

Organization & Address

Publications

ICC International Code Council http://www.intlcode.org/


BOCA International, Inc. 4051 West Flossmoor Road Country Club Hills, IL 50478-5794
http://www.bocai.org/
ICBO 5360 Workman Mill Road Whittier, CA 90601-2298 http://www.icbo.org/
SBCCI, Inc. 900 Montclair Road Birmingham, AL 35213-1206 http://www.sbcci.org/

International building code International energy conservation code International


zoning code International property maintenance code commentary International
property maintenance code International fuel gas code International mechanical code
commentary International mechanical code International mechanical code supplement
International private sewage disposal code International one and two family
dwelling code International plumbing code commentary International plumbing code

MBMA Metal Building Manufacturers Assoc. 1300 Sumner Ave Cleveland, OH 44115-2851
http://www.mbma.com/

Low rise building systems manual Metal building systems

NFBA National Frame Builders Association 4840 W. 15th St., Suite 1000 Lawrence, KS
66049-3876 http://www.postframe.org/

Post wall assembly fire test

NFPA National Fire Protection Association 1 Batterymarch Park Quincy, MA 02269-9101


http://www.nfpa.org/

NFPA 1: Fire prevention code NFPA 13: Installation of sprinkler NFPA 70: National
electrical code NFPA 72: National fire alarm code NFPA 101: Life safety code

SPIB Southern Pine Inspection Bureau 4709 Scenic Highway Pensacola, Fl. 32504-9094
http://www.SPIB.org/

Grading rules Standard for mechanically graded lumber Kiln drying southern pine

2-6

National Frame Builders Association

Post-Frame Building Design Manual

Table 2.1. Partial list of technical references related to post-frame building


design and construction

Organization & Address

Publications

SFPA & Southern Pine Council Southern Forest Products Association P. O. Box 641700
Kenner, LA 70064-1700 http://www.southernpine.com/ http://www.SFPA.org/

Southern pine use guide Southern pine joists & rafters: construction guide Southern
pine joists & rafters: maximum spans Post-frame construction guide Southern pine
headers and beams Pressure-treated southern pine Permanent wood foundations: design
& construction guide

TPI Truss Plate Institute 583 D'Onofrio Drive, Suite 200 Madison, WI 53719

ANSI/TPI 1-1995 National design standard for metal plate connected wood truss
construction
HIB-91 Summary sheet: handling, installing & bracing metal plate connected wood
trusses
HIB-98 Post frame summary sheet: recommendations for handling, installing &
temporary bracing metal plate connected wood trusses used in post-frame
construction
HET-80 Handling & erecting wood trusses: commentary and recommendations
DSB-89 Recommended design specifications for temporary bracing of metal plate
connected wood trusses

UL Underwriters Laboratories, Inc. 333 Pfingsten Road Northbrook, IL 60062-2096


http://www.ul.com/

Fire resistance directory

WTCA Wood Truss Council of America One WTCA Center 6425 Normandy Lane Madison, WI
53711 http://www.woodtruss.com/

Metal plate connected wood truss handbook Commentary for permanent bracing of metal
plate connected
wood trusses Standard responsibilities in the design process involving metal
plate connected wood trusses

WWPA Western Wood Products Association 522 SW Fifth Ave., Suite 500 Portland,
Oregon 97204-2122 http://www.wwpa.org/

Western woods use book Western lumber span tables Western lumber grading rules

2-7

National Frame Builders Association

Post-Frame Building Design Manual

2-8

National Frame Builders Association

Post-Frame Building Design Manual

Chapter 3: STRUCTURAL LOAD AND DEFLECTION CRITERIA

3.1 Introduction
3.1.1 Load Variations. Most structural loads exhibit some degree of random
behavior. For example, weather-related loads such as snow, wind and rain fluctuate
over time and locations. Extensive research has been conducted to characterize this
load variation, and to refine procedures for determining design loads within the
context of the intended building occupancy and use.
3.1.2 Codes. Calculation procedures for minimum design loads are given in the model
building codes. Buildings shall be designed to safely carry all loads specified by
the governing building code. In the absence of a code, minimum design loads shall
be calculated according to recommended engineering practice for the region and
application under consideration.
It is impractical to describe detailed load calculation procedures in this chapter
because of differences between building codes and frequent revisions of these
codes. Instead, general concepts and key references related to structural loads and
deflection criteria are presented, with an emphasis on issues that apply to post-
frame buildings.
3.2 Technical References on Structural Load Determination
3.2.1 ANSI/ASCE 7 Standard. The National Bureau of Standards published a report
titled Minimum Live Load Allowable for Use in Design of Buildings in 1924. The
report was expanded and published as ASA Standard A58.1-1945. This standard has
undergone several revisions to become the current ASCE Standard ANSI/ASCE 7 Minimum
Design Loads for Buildings and Other Structures. At the time this design manual was
written, the most recent revision of ASCE 7 was 1999 (ASCE, 1999); however, the
edition most commonly used is ASCE 7-93. The ASCE 7 standard is periodically
revised and balloted through the ANSI consensus

approval process, and then must be adopted by the model building codes. Design
professionals should check the governing building code for the latest adopted
edition. For clarity of presentation, this manual uses and will refer to ASCE 793.
ASCE 7-93 is the primary technical source used by the model codes concerning dead,
live, snow, wind, rain and seismic loads. Basically, the model codes attempt to
distill the rigorous ASCE 7-93 procedures into a simpler, easy-touse format. Many
specific load calculation procedures differ between the model codes; however, most
of the basic concepts mimic ASCE 793. Background information on the wind load
provisions in ASCE 7-88 (which are essentially the same as in ASCE 7-93) are given
by Mehta et al. (1991).
3.2.2 Low Rise Building Systems Manual. The Low Rise Building Systems Manual,
published by the Metal Building Manufacturers Association (1986), is recognized by
model building codes as an excellent technical resource document for calculating
structural loads on lowrise buildings (e.g. post-frame buildings). This document
will be referred to as MBMA-86 throughout this manual. Because wind and crane loads
frequently control the design of lowrise metal buildings, the coverage of these
loads within MBMA-86 is especially thorough. Another attractive feature of MBMA-86
is the extensive collection of example load calculations.
3.2.3 ASAE EP288.5 Standard. Agricultural buildings generally fall into a separate
class from other types of buildings due to the lower risks involved. The American
Society of Agricultural Engineers publishes a snow and wind load standard, EP288.5,
intended for agricultural buildings (ASAE, 1999). The major differences between
agricultural and other types of buildings are that lower values are used for
importance and roof snow conversion factors (due to relatively lower risk factors
for property and nonpublic use). If the local governing building code applies to
agricultural buildings, then the design load criteria in the code must be followed.

3-1

National Frame Builders Association

Post-Frame Building Design Manual

Table 3.1. Approximate Weights of Construction Materials (from Hoyle and Woeste,
1989)

Material

Weight (lb/ft2)

Material

Weight (lb/ft2)

Ceilings Acoustical fiber tile Gypsum board (see Walls) Mechanical duct allowance
Suspended steel channel system Wood purlins (see Wood, Seasoned) Light gauge steel
(see Roofs)
Floors Hardwood, 1-in. nominal Plywood (see Roofs) Linoleum, 1/4-in. Vinyl tile,
1/8-in.
Roofs Corrugated Aluminum 14 gauge 16 gauge 18 gauge 20 gauge Built-Up 3-ply 3-ply
with gravel 5-ply 5-ply with gravel Corrugated Galvanized steel 16 gauge 18 gauge
20 gauge 22 gauge 24 gauge 26 gauge 29 gauge Insulation, per inch thickness Rigid
fiberboard, wood base Rigid fiberboard, mineral base Expanded polystyrene
Fiberglass, rigid Fiberglass, batt Lumber (see Wood, Seasoned)

1.0
4.0 2.0
4.0
1.0 1.4
1.1 0.9 0.7 0.6
1.5 5.5 2.5 6.5
2.9 2.4 1.8 1.5 1.3 1.0 0.8
1.5 2.1 0.2 1.5 0.1

Roofs (continued) Plywood (per inch thickness) Roll roofing Shingles Asphalt Clay
tile Book tile, 2-in. Book tile, 3-in Ludowici Roman Slate, in. Wood

3.0 1.0
2.0 9.0-14.0
12.0 20.0 10.0 12.0 10.0 3.0
Walls Wood paneling, 1-in. Glass, plate, 1/4-in. Gypsum board (per 1/8-in.
thickMasonry, per 4-in. thickness Brick Concrete block Cinder concrete block Stone
Porcelain-enameled steel Stucco, 7/8-in. Windows, glass, frame, and sash

2.5 3.3 0.55


38.0 20.0 20.0 55.0 3.0 10.0 8.0

Wood, Seasoned
Cedar Douglas-fir Hemlock Maple, red Oak Poplar, yellow Pine, lodgepole Pine,
ponderosa Pine, Southern Pine, white Redwood Spruce

Density lb/ft3
32.0 34.0 31.0 37.0 45.0 29.0 29.0 28.0 35.0 27.0 28.0 29.0

3-2

National Frame Builders Association

Post-Frame Building Design Manual

3.3 Minimum Design Loads


Sections 3.4, 3.5, 3.6, 3.7, and 3.8 give general load requirements, sources of
load data and references for making detailed load calculations. Detailed
calculation procedures are not provided due to differences between the model codes
and the frequency of code revisions.
3.4 Dead Loads
3.4.1 Definition. Dead loads are the gravity loads due to the combined weights of
all permanent structural and nonstructural components of the building, such as
sheathing, trusses, purlins, girts and fixed service equipment. These loads are
constant in magnitude and location throughout the life of the building.
3.4.2 Code Application. Minimum design dead loads shall be determined according to
the governing building code. In the absence of a building code, dead load data can
be found in ASCE 7-93, or actual weights of materials and equipment can be used.
3.4.3 Special Considerations. Design dead loads that exceed the weights of
construction materials and permanent fixtures are permitted, except for when
checking building stability under wind loading. Using inflated design dead loads
may lead to conservative designs for gravity load conditions; however, it would not
be a conservative assumption for designing anchorage to counteract uplift,
overturning and sliding due to wind loads. In the cases of wind uplift and
overturning, the dead load used in design must not exceed the actual dead load of
the construction.
3.4.4 Weights of Construction Materials. Table 3.1 lists approximate weights of
materials. commonly used in post-frame construction.
3.5 Live Loads
3.5.1 Definition. Live loads are defined as the loads superimposed by the
construction, maintenance, use and occupancy of the building, and therefore do not
include wind, snow, seismic or dead loads.

Technical Note
Horizontal Uniform Dead Load Calculation
Many structural analysis programs (e.g. Purdue Plane Structures Analyzer) require
that the dead load associated with a sloping surface be represented as a uniform
load, wDL, acting on a horizontal plane as shown in figure 3.1. For a given
horizontal distance, bH, a sloping roof surface contains more material and is
heavier than a flat one. Thus, wDL increases as roof slope increases.
Load wDL is obtained by multiplying the unit weight of the roof assembly, wR, by
the slope length, bS, and dividing the resulting product by the horizontal length,
bH. Numerically, this is equivalent to dividing wR by the cosine of the roof slope.
Example: For a roof at a 4:12 slope, with materials weighing 4 lbm for each square
foot of roof surface area, the equivalent load, wDL, to apply to the horizontal
plane would be:
wDL = (4 lbm/ft2)/(cos 18.4) = 4.21 lbm/ft2
wDL

Roof assembly with weight wR per unit area


Rafter or truss top chord

bH

bS

Figure 3.1. Roof dead load represented by an equivalent uniform load acting on a
horizontal plane.
3.5.2 Code Application. Design live loads shall be determined so as to provide for
the service requirements of the building, but should never be lower than the
minimum live load specified in

3-3

National Frame Builders Association

Post-Frame Building Design Manual

the governing building code. In the absence of a governing building code, the
minimum live loads found in ASCE 7-93 are recommended. The minimum roof live load
recommended for agricultural buildings in ASAE Standard EP288.5 is 12 psf. Some
agricultural buildings do not necessarily pose a "low risk", and the ASAE higher
minimum live load reflects the possibility of highvalue agricultural constructions
now common in the United States
3.5.3 Reductions. In some cases, reductions are allowed for uniform loads to
account for the low likelihood of the loads simultaneously occurring over the
entire tributary area.

3.6 Snow Loads

3.6.1 Code Application. Minimum design snow loads shall be determined by the
provisions of the governing building code. The presentation of snow loads varies
among the model codes, but they all follow the basic concepts presented in ASCE 7-
93. In the absence of a building code, procedures given in ASCE 7-93 are
recommended. For low-risk agricultural buildings, snow load calculation procedures
given in ASAE EP288.5 are permitted.

3.6.2 Ground Snow Load Maps. ASCE 7-93 presents ground snow load maps that
correspond to a mean recurrence interval of 50 years. These maps do not give snow
load values for areas that are subject to extreme variations in snowfall, such as
western mountain regions. In some regions, the best and only reliable source for
ground snow loads is local climatic records.

3.6.3 Roof Snow Loads. Roof snow loads are influenced by a number of factors
besides ground snow load. These factors include roof slope, temperature and
coefficient of friction of the roof surface, and wind exposure. Snow loads are also
adjusted by an importance factor to account for risk to property and people. The
basic form of the snow load calculation found in ASCE 7-93 is:

pf = R Ce Ct I Cs Pg where:

(3-1)
pf = R=
Ce = Ct = I= Cs = Pg =

roof snow load in psf, roof snow factor that relates roof load to ground snowpack,
snow exposure factor, roof temperature factor, importance factor, roof slope
factor, and ground snow load in psf (50-yr mean recurrence).

The roof snow factor, R, varies from 0.6 for Alaska to 0.7 for the contiguous
United States. The snow exposure factor in the model codes accounts for the
combined effects of R and Ce given in Equation 3-1. The thermal factor defined in
ASCE 7-93 varies from 1.0 for heated structures to 1.2 for unheated structures. The
thermal factor is not included in the model building codes. The importance factors
range from 0.8 to 1.2 depending on the specific building code. Roof slope factors
vary linearly from 0 to 1 as roof slope increases from 15 to 70 degrees.

3.6.5 Special Considerations. Several factors, such as multiple gables, roof


discontinuities, and drifting can cause snow to accumulate unevenly on roofs. These
factors must be considered in the design. Specific recommendations and calculation
procedures are given in the model codes and ASCE 7-93.

3.7 Wind Loads

3.7.1 Controlling Factors. Wind loads are influenced by wind speed, building
orientation and geometry, building openings and exposure. Wind loading on
structures is a complex phenomenon and is being actively researched.
3.7.2 Code Application. Minimum design wind loads shall be determined by the
provisions of the governing building code. In the absence of a building code,
procedures given in ASCE 7-93 or MBMA-86 are recommended. For low-risk agricultural
buildings, wind load calculation procedures given in ASAE EP288.5 are permitted.
3.7.3 Design Wind Speed. ASCE 7-93 gives a map showing basic wind speeds throughout
the United States that correspond to a mean recurrence interval of 50 years. Local
weather rec-

3-4

National Frame Builders Association

Post-Frame Building Design Manual

ords should be used in regions that have unusual wind events. Detailed procedures
and illustrations for calculating wind loads on low-rise buildings are given in
MBMA-86.
Technical Note
Wind Speed
Wind speeds are derived from data which reflect both magnitude and duration. Wind
speeds can be reported as peak gusts, or can be averaged over some time interval.
The time interval may be fixed, as with mean hourly speeds, or variable, as with
fastest-mile wind speeds. Fastest-mile wind speeds are used in ANSI/ASCE 793 to
calculate design loads, and are defined on the basis of the period of time that one
mile of wind takes to pass an anemometer at a standard elevation of 10 meters. The
U.S. National Weather Service no longer collects fastest-mile wind speed data;
instead, they record 3-second gust speeds. The 1995 and later revisions of ASCE-7
base wind loads on 3-second gust wind speeds.

3.7.4 Effective Wind Velocity Pressure. The first step in determining wind loads is
to calculate the effective wind velocity pressure. The most severe exposure factors
that will apply during the service life of the structure should be used. Wind
velocity pressure is a function of the wind speed, exposure and importance. The
equation for calculating wind velocity pressure, qz , is given by:

qz = 0.00256 Kz (I V)2

(3-2)

where:

Kz =
I= V=

velocity pressure exposure coefficient, importance factor, and basic wind speed in
mph (50-year mean recurrence interval).

The velocity pressure exposure coefficient is a function of height above ground and
exposure category. Exposure categories account for the effects of ground surface
irregularities caused by natural topography, vegetation, location and building
construction features. ASCE 7-93 lists four wind exposure categories, whereas the

model codes publish fewer exposure categories. Importance factors vary from 0.95
for agricultural buildings (25-year recurrence interval) to 1.07 for buildings that
represent a high hazard to property and people in the event of failure (100year
recurrence interval). Wind pressure is related to the square of its speed,
therefore the terms V and I are squared in equation 3-2. The model building codes
simplify the calculation in equation 3-2 by publishing tables of effective wind
velocity pressures, Pb, for a base wind speed and various heights.

3.7.5 Pressure Coefficients. Wind loads are calculated for each part of the
building by multiplying the effective wind pressure by a pressure coefficient. The
pressure coefficient, which may be different for each planar portion of the
building, accounts for building orientation, geometry and load sharing. It also
accounts for localized pressures at eaves, overhangs, corners, etc. Wind pressures,
qi, for the ith building surface are calculated by:

qi = Cpi qz

(3-3)

where:

Cpi = ith pressure coefficient, and qz = wind velocity pressure.

The wind velocity pressure is based on the wall height for the windward wall and on
the mean roof height for the leeward wall and roof. Wind pressures act normal to
the building surfaces. Inward pressures are denoted with positive signs, while
outward pressures (suction) are denoted with negative signs.

Technical Note
Components of Wind Load
Many structural analysis programs require uniform loads to be entered in terms of
their horizontal and vertical components. Wind loads act normal to building
surfaces, so an adjustment is needed for sloping members such as roof trusses. The
roof wind load, w, shown in figure 3.2a is equivalent to the horizontal and
vertical components shown in figure 3.2b. The relationship depicted in figure 3.2
can be proven as follows:

3-5

National Frame Builders Association


Post-Frame Building Design Manual

1. Convert the uniform wind load, w, to its resultant vector force.


R = w (span)/(cos )
2. Multiply resultant force, R, by cos to obtain its vertical component.
Fy = R (cos ) = w (span)
3. Divide the vertical component, Fy, by the span to obtain the horizontally
projected uplift pressure, whoriz.
whoriz = Fy /(span) = w (span)/(span) = w
The vertically projected uniform load can be proven similarly. A common mistake is
to multiply the normal pressure by sine and cosine of the roof slope to obtain the
two components.
w

(a)
w
w

(b)
Figure 3.2. Illustration of wind load acting normal to inclined surface and
equivalent horizontal and vertical load components. A common mistake is to multiply
the normal load by sin() and cos() for the vertical and horizontal components,
respectively.
3.7.6 Main Frames. Different pressure coefficients are used to calculate wind loads
on main frames as compared to components and cladding. Main frames include primary
structural systems such as rigid and braced frames, braced trusses, posts, poles
and girders. Since

these members have relatively large tributary areas, localized wind effects tend to
be averaged out over the tributary areas. Pressure coefficients for main members
reflect this averaging effect.
3.7.7 Components and Cladding. Wind pressures are higher on small areas due to
localized gust effects. This observation has been verified by wind tunnel studies
(MBMA, 1986), as well as site inspections of wind-induced building failures
(Harmon, et al., 1992). For this reason, components and cladding have higher
pressure coefficients than main frames. Components and cladding include members
such as purlins, girts, curtain walls, sheathing, roofing and siding.
3.7.8 Openings. Wind loads are significantly affected by openings in the structure.
ASCE 793 and the model building codes specify internal wind pressure coefficients
(or adjustments to external pressure coefficients) for structures with different
amounts and types of openings. Each model code has slightly different definitions
and wind load coefficients for open, closed and partially open buildings. In
general, "openings" refer to permanent or other openings that are likely to be
breached during high winds. For example, if window glazings are likely to be broken
during a windstorm, the windows are considered openings. However, if doors and
windows and their supports are designed to resist design wind loads, they need not
be considered openings. It should be noted that internal wind pressures act against
all interior surfaces and therefore do not contribute to sidesway loads on a
building.
3.8 Seismic Loads
3.8.1 Cause. Earthquakes produce lateral forces on buildings through the sudden
movement of the buildings foundation. Building response to seismic loading is a
complex phenomenon and there is considerable controversy as to how to translate
knowledge gained through research into practical design codes and standards.
3.8.2 Code Application. Seismic loads shall be determined by the provisions of the
governing building code. In the absence of a building code, procedures given in
ASCE 7-93 are recommended. Sweeping changes were made in the
3-6

National Frame Builders Association

Post-Frame Building Design Manual

1993 revision of ASCE 7 with respect to seismic loads. The seismic provisions in
ASCE 7-93 were based on work by the Building Seismic Safety Council under
sponsorship of the Federal Emergency Management Agency.

3.8.3 Lateral Force. Basic concept of seismic load determination for low-rise
buildings is to calculate an equivalent lateral force at the ground line as
follows:

V = Cs W where:

(3-4)

V= W=
Cs = =

total lateral force, or shear, at the building base total dead load, plus other
applicable loads specified in the code or ASCE 7-93. For most single-story post-
frame buildings, the only other minimum applicable load is a portion (20% minimum)
of the flat roof snow load. If the flat roof snow load is less than 30 psf, the
applicable load to be included in W is permitted to be taken as zero. seismic
design coefficient 1.2 Av S/(T2/3 R)

Av =
S=
R= T=

coefficient representing effective peak velocity-related acceleration coefficient


for the soil profile characteristics response modification factor fundamental
period of the building

3.8.4 Seismic loads rarely control post-frame building design because of the
relatively low building dead weight as compared with other types of construction
(Taylor, 1996; Faherty and Williamson, 1989). For post-frame buildings, lateral
loads from wind usually are much greater than those from seismic forces.

3.9 Load Combinations for Allowable Stress Design

3.9.1 Code Application. Every building element shall be designed to resist the most
critical load combinations specified in the governing building code.

3.9.2 Load Combinations. Except when applicable codes provide otherwise, the
following load combinations shall be considered (as a minimum) and the combination
which results in the most conservative design for each building element shall be
used. Note that different load combinations may control the design of different
components of the structure.
Case 1: Dead + Floor Live + Roof Live (or Snow) Case 2: Dead + Floor Live + Wind
(or Seismic) Case 3: Dead + Floor Live + Wind + Snow Case 4: Dead + Floor Live +
Wind + Snow Case 5: Dead + Floor Live + Snow + Seismic
3.9.3 Floor Live Loads. Most post-frame buildings are single story and therefore
would not have floor live loads acting on the post-frames. When a concrete floor is
used in a single story building, consideration must be given to anticipated live
and equipment loading.
3.9.4 Reductions. Reductions in some of the load terms in Cases 1 through 5 are
permitted, depending on governing building code or reference document. With some
exceptions, the model building codes permit allowable stresses used in allowable
stress design to be increased one-third when considering wind or seismic forces
either acting alone or when combined with vertical loads. The allowable stress
increase for wind loading can be traced back to the New York City Building Code of
1904 (Ellifritt, 1977), and appears to be based on judgment rather than engineering
theory. It should be noted that ASCE 7-93 does not include the one-third increase
factor, but instead specifies load combination factors that are intended to account
for the low probability of maximum live, seismic, snow and wind loads occurring
simultaneously. The commentary of ASCE 7-93 implies the stress increase for wind
and seismic found in codes is not appropriate if the combined load effects are also
reduced by the load combination factors published in ASCE 7-93. Finally, the
National Design Specification (NDS) for Wood Construction (NF&PA, 199) addresses
the issue of load combination versus load duration factors by stating, The load
duration factors, CD, in Table 2.3.2 and Appendix B are independent of load
combination factors, and both shall be permitted to be used in design
calculations.

3-7

National Frame Builders Association

Post-Frame Building Design Manual

3.10 Load Duration Factors for Wood


It is well documented that wood has the property of being able to carry
substantially greater loads for short durations than for long durations of loading.
This property is accounted for in design through the application of load duration
factors to all allowable design values except modulus of elasticity and compression
perpendicular to grain. Additional restrictions and details on load duration
adjustments can be found in Chapter 2 and Appendix B of the NDS (AF&PA, 1997).
3.10.1 Snow Load. The cumulative duration of maximum snow load over the life of a
structure is generally assumed to be two months. It should be emphasized that the
two-month period does not necessarily mean that the design snow load from any one
event would last two months. Rather, it means that the total time that the roof
supports the full design snow load over the life of the structure is two months. If
the cumulative full design load is two months, an allowable stress increase of 15
percent is allowed (AF&PA, 1997). However, in some situations, such as unheated or
heavily insulated buildings in cold climates, longer snow load durations may occur
and the stress increase may not be justified.
3.10.2 Wind Load. The cumulative duration of maximum wind (and seismic) loads over
the life of a structure is generally assumed to be 10 minutes (AF&PA, 1997), if
design wind loads are based on ASCE 7-93, and the corresponding load duration
factor is 1.6. Other load duration adjustments may be appropriate when design wind
loads are based on earlier versions of ASCE 7-93 or other standards (with different
wind gust duration assumptions).
3.11 Deflection
3.11.1 Code Application. Post-frame building components must meet deflection limits
specified in the governing building code.
3.11.2 Exception to Code Requirements. Girts supporting corrugated metal siding are
typically not subjected to deflection limitations unless their deflection
compromises the integrity of an interior wall finish. Because of the inherent

flexibility of corrugated metal siding, girt deflections present no serviceability


problems, and consequently, girt size is generally only stress dependent.
3.11.3 Time Dependent Deflection. In certain situations, it may be necessary to
limit deflection under long term loading. Published modulus of elasticity, E,
values for wood are intended for the calculation of immediate deflection under
load. Under sustained loading, wood members exhibit additional time-dependent
deformation (i.e. creep). It is customary practice to increase calculated
deflection from long-term loading by a factor of 1.5 for glued-laminated timber and
seasoned lumber, or 2 for unseasoned lumber (see Appendix F, AF&PA, 1997). Thus,
total deflection is equal to the immediate deflection due to long-term loading
times the creep deflection factor, plus the deflection due to the short-term or
normal component of load. For applications where deflection is critical, the
published value of E (which represents the average) may be reduced as deemed
appropriate by the designer. The size of the reduction depends on the coefficient
of variation of E. Typical values of E variability are available for different wood
products (see Appendix F, AF&PA, 1997).
3.11.4 Shear Deflection. Shear deflection is usually negligible in the design of
steel beams; however, shear deflection can be significant in wood beams.
Approximately 3.4 percent of the total beam deflection is due to shear for wood
beams of usual span-to-depth proportions (i.e. 15:1 to 25:1). For this reason, the
published value of E in the Supplement to the National Design Specification is 3.4
percent less than the true flexural value (AF&PA, 1993). This correction
compensates for the omission of the shear term in handbook beam deflection
equations. For span-to-depth ratios over 25, the predicted deflection using the
published E value will exceed the actual deflection. Similarly, for span-todepth
ratios less than 15, predicted deflections will be significantly less than actual.
This could lead to unconservative designs (with respect to serviceability) for
post-frame members such as door headers. Practical information on the effects of
shear deformation on beam design is given in Appendix D of Hoyle and Woeste (1989)
for rectangular wood beams and Triche (1990) for wood I-beams.

3-8

National Frame Builders Association


3.12 References
American Forest & Paper Association (AF&PA). 1997. ANSI/AF&PA NDS-1997 - National
Design Specification for Wood Construction. AF&PA, Washington, D.C.
American Forest & Paper Association (AF&PA). 1993. Commentary to the National
Design Specification for Wood Construction. AF&PA, Washington, D.C.
ASAE. 1999. ASAE EP288.5: Agricultural building snow and wind loads. ASAE Standards
1999, 46th edition, ASAE, St. Joseph, MI.
American Society of Civil Engineers (ASCE). 1993. Minimum design loads for
buildings and other structures. ANSI/ASCE 7-93, ASCE, New York, NY.
American Society of Civil Engineers (ASCE). 1999. Minimum design loads for
buildings and other structures. ANSI/ASCE 7-99, ASCE, New York, NY.
Ellifritt, D.S. 1977. The mysterious 1/3 stress increase. American Institute of
Steel Construction Engineering Journal (4):138-140.
Faherty, K.F. and T.G. Williamson. 1989. Wood Engineering and Construction
Handbook. McGraw-Hill, New York, NY.
Hoyle, R.J. and F.E. Woeste. 1989. Wood Technology in the Design of Structures.
Ames, IA: Iowa State University Press.
Mehta, K.C., R.D. Marshall and D.C. Perry. 1991. Guide to the Use of the Wind Load
Provisions of ASCE 7-88 (formerly ANSI A58.1). American Society of Civil Engineers,
New York, NY.
Metal Building Manufacturers Association (MBMA). 1986. Low rise building systems
manual. MBMA, Cleveland, OH.
Taylor, S.E. 1996. Earthquake considerations in post-frame building design. Frame
Building News 8(3):42-49.
3-9

Post-Frame Building Design Manual

National Frame Builders Association

Post-Frame Building Design Manual


3-10

National Frame Builders Association

Post-Frame Building Design Manual

Chapter 4: STRUCTURAL DESIGN OVERVIEW

4.1 Introduction
4.1.1 General. The aim of this chapter is to give a broad overview of post-frame
building design, and then highlight unique aspects of post-frame that require
special design considerations. Postframe is a special case of light-frame wood
construction. Light-frame construction is accepted by all model building codes, and
the design procedures are well documented. The design rules that apply to light-
frame wood construction also apply to post-frame. However, there are some aspects
of post-frame that are not as familiar to building designers, such as diaphragm
design, interaction between post-frames and diaphragms, and post foundation design.
Hence, Chapters 5, 6, 7 and 8 focus on these topics in more detail.
4.1.2 Primary Framing. Primary framing is the main structural framing in a
building. In a postframe building, this includes the columns, trusses (or rafters),
and any girders that transfer load between trusses and columns. Each truss and the
post(s) to which it is attached form an individual "post-frame". Post-frames
collect and transfer load from roof purlins and wall girts to the foundation. In
the context of wind loading in standards and building codes, post-frames are an
integral part of the main wind-force resisting system. Specific sections dedicated
to primary framing include: Section 4.2 Posts, Section 4.3 Trusses, Section 4.4
Girders, and Section 4.5 Knee braces.
4.1.3 Secondary Framing. Secondary framing includes any framing member used to (1)
transfer load between cladding and primary framing members, and/or (2) laterally
brace primary framing members. The secondary framing members in a post-frame
building include the girts, purlins and any structural wood bracing such as
permanent truss bracing. Specific sections dedicated to secondary framing include:
Section 4.6 Roof Purlins, Section 4.7 Wall Girts, and Section 4.8 Large Doors.
4.1.4 Diaphragms and Shearwalls. When cladding is fastened to the wood frame of a
post-frame building, large shearwalls and roof

and ceiling diaphragms are formed that can add considerable rigidity to the
building. In many post-frame buildings, diaphragms and shearwalls are carefully
designed and become an integral part of the main wind-force resisting system. Roof
and Ceiling Diaphragms are covered in Section 4.9 and Shearwalls in Section 4.10.
4.1.5 Limitations. The structural design of buildings involves making many
judgments, such as determining design loads, structural analogs and analyses, and
selecting materials that can safely resist the calculated forces. New research or
testing could justify a change of design procedure for the industry or for an
individual designer. The considerations presented here are not exhaustive, since
many issues in a specific building design will require unique treatment.
4.2 Posts
4.2.1 General. The function of the wood post is to carry axial and bending loads to
the foundation. Posts are embedded in the ground or attached to either a
conventional masonry or concrete wall or a concrete slab on grade. Posts can be
solid sawn, mechanically laminated, glued-laminated or wood composite. Any portion
of a post that is embedded or exposed to weather must be pressure-treated with
preservative chemicals to resist decay and insect damage.
4.2.2 Controlling Load Combinations. The load combination that usually controls
post design is dead plus wind plus one-half snow; however, local codes may
stipulate different load combinations. It is possible for any one of the
combinations to be critical; therefore, they all should be considered for a
specific building design. For example, maximum gravity load will govern truss-to-
post bearing and post foundation bearing; whereas wind minus dead load will govern
the truss-to-post connection (for uplift).
4.2.3 Force Calculations. The diaphragm analysis method presented in Chapter 5 is
the most accurate method to determine design

4-1

National Frame Builders Association

Post-Frame Building Design Manual

moments, and axial and shear forces in posts. Historically, some designers
calculated the maximum post moment for embedded posts by using the simple
structural analog of a propped cantilever (i.e. fixed reaction at the post bottom
and pin reaction at the top). The implicit assumption of this analog is that the
roof diaphragm and shearwalls are infinitely stiff. This model may be adequate for
buildings with extremely stiff roof diaphragms and for conservatively estimating
shear forces in the roof diaphragm; however, it may underestimate the maximum post
moment for many post-frame buildings. The analysis procedures described in Chapter
5 are more reliable since they account for the flexible behavior of the roof
diaphragm.
If posts are embedded, generally two bending moments must be calculated - one at
the groundline and the other above ground. Groundline bending moment and shear
values are used in embedded post foundation design calculations. For surface-
attached posts, the bottom reaction can be modeled as a pin, and generally only one
bending moment is calculated.
4.2.4 Combined Stress Analysis. Forces involved in post design subject the posts to
combined stress (bending and axial) and must be checked for adequacy using the
appropriate interaction equation from the NDS (AF&PA, 1997). In theory, every post
length increment must satisfy the interaction equation, but in practice, a minimum
of two locations are checked: the point of maximum interaction near the ground
level (column stability factor, Cp, equal to 1.0) and the upper section of the
posts where the maximum moment occurs in conjunction with column action (Cp<1.0).
4.2.5 Shear Stress. The shear stress due to lateral loading (wind or seismic)
rarely controls post design, but should always be checked as a matter of good
practice. Other loads such as bulk loads from stored materials may influence final
post design.
4.2.6 Deflection. A post deflection limit is not normally specified for post-frame
buildings, but interior finishes may require it. Refer to the deflection criteria
in Chapter 3.

4.2.7 Connections. Truss-to-post connection must be designed for bearing as well as


uplift. Connection design procedures are given in the NDS (AF&PA, 1997). This
connection should be modeled as a pin unless moment-carrying capacity can be
justified. Direct end grain bearing is desirable and is often achieved by notching
the post to receive the truss. When designing the truss-to-post connection for
uplift, it is important to accurately estimate the weights of construction
materials if any counteracting credit is to be taken.
For surface-attached posts, the bottom connection needs to be checked for maximum
shear and uplift forces. For embedded posts attached to collars or footings, the
connections must be properly designed to withstand gravity and uplift loads, and
corrosion-resistant fasteners must be used.
4.2.8 Construction Alternatives. The posts in post-frame buildings can be solid
sawn, mechanically-laminated, glued-laminated or wood composite. Allowable design
stresses are published in the NDS or are available from the manufacturers. Treated
wood is used for the embedded part of the post, but no treatment is required on the
parts that are not in contact with the ground and are protected by the building
envelope.
4.2.9 Foundation. Post-frame building foundations include posts embedded in the
ground or surface-attached on a concrete foundation. Embedded posts shall be
designed to resist sidesway and overturning forces due to wind or seismic loads, as
well as wind uplift, and gravity loads. Post foundation design is an important
aspect of post-frame building design that is not well known in the structural
engineering design community, and therefore Chapter 8 is dedicated to this subject.
If a concrete slab is used, it only needs to be designed for interior loads since
exterior building loads are transferred directly to the ground through the posts.
Another option is to attach the posts to a concrete foundation. In this case, the
concrete must be designed to carry the exterior building loads as well as interior.
Connections must be designed to attach the posts to the concrete.

4-2

National Frame Builders Association

Post-Frame Building Design Manual

4.2.10 Pressure Preservative Treatment (PPT). Treated foundation systems have been
accepted by the model codes and have a history of successful performance in
residential wood construction. The most common pressure preservative treatment used
in post-frame construction is chromated copper arsenate (CCA). CCA can increase the
potential for metal-fastener corrosion, and may require hot-dipped galvanized or
stainless steel fasteners. The minimum waterborne treatment retention for
structural posts used in post-frame buildings is 0.6 lb/ft3 (pcf) as defined in
AWPA Standard C15 (AWPA, 1995a).
Technical Note
Wood Preservative Treatments
When the moisture content of wood exceeds 20% on a dry weight basis in the presence
of oxygen, it is vulnerable to attack by insects and decaying fungi. Although some
species of wood (and the heartwood of other species) are naturally resistant to
these types of attack, most structural woods used in North America are not. These
structural wood species must be chemically treated to protect them from decay and
maintain their strength throughout the structural design life.
The chemicals used for preservative treatment of the wood are typically injected
into the wood using pressure processes. Wood that has been chemically treated in
this manner is accepted by all major building codes. The type of preservative
treatment and the required amount of retention by the wood depends on the end use
of the wood component. It is assumed that the designer is already familiar with the
use of preservative treated wood for above-ground applications (such as wood
decks); this section will concentrate specifically on preservative treatments and
retention levels appropriate for use in post foundations.
Preservative chemicals abate wood decay by altering the wood as a potential food
source for insects and fungi. The preservatives typically used in North America are
waterborne arsenicbased, pentachlorophenol (penta) and creosote. Waterborne
arsenic-based preservatives include chromated copper arsenate (CCA), ammoniacal

copper arsenate (ACA), and ammoniacal copper zinc arsenate. CCA is available in
three formulations: CCA-A, CCA-B, and CCA-C. CCA-C is the most popular of the three
formulations due to its increased resistance to leaching.
Penta is an oil-borne preservative, and creosote is a coal-tar based preservative
that is its own carrier. While penta and creosote offer superior resistance in high
salt environments, waterborne preservatives are typically more popular since the
final product has a clean surface, is paintable, and is relatively odorless.
Waterborne preservatives do provide a strong potential for corrosion of metal
connectors and fasteners; follow the manufacturers recommendations for the use of
stainless or hot-dipped galvanized fasteners.
While the major building codes endorse the use of preservative-treated wood for
foundation applications, it is imperative that the preservative retention
guidelines be followed. The American Wood Preservers Association has published
standards for the preservative treatment of wood for various applications (AWPA,
1991). Care must be taken that the appropriate standard is considered when
specifying treatment for post foundation systems. For example, most waterborne
preservative-treated lumber sold has a preservative retention level of 0.4 pcf
(pounds of preservative per cubic foot of wood), which is the retention level
specified by AWPA Standard C2 for lumber in contact with the ground. This differs,
however, from the AWPA Standard C15 governing the treatment of structural posts
used in foundations; the required preservative retention for waterborne
preservatives under this standard is 0.6 pcf. The AWPA C15 required retention level
for post foundations using penta as a preservative is 0.6 pcf, while the required
retention level for creosote is 12.0 pcf.
The rate at which treatments are absorbed into wood, and the depth of penetration
of the treatment, varies from wood species to wood species. Whereas Southern Pine
species take treatment quite well, most other species must be incised to comply
with AWPA retention requirements. Incising can adversely affect lumber strength
properties. Consult AF&PA for specifications regarding the use of incised wood in
structural applications.

4-3

National Frame Builders Association

Post-Frame Building Design Manual

Quality assurance is critical to the performance of treated wood. The treating


industry has developed a quality control and treatment quality marking program
accredited by the American Lumber Standards Committee. Any treated members
specified for use in a post foundation should be stamped by an approved agency
(e.g., AWPA, Southern Pine Inspection Bureau (SPIB), etc.) to assure that the
members have been treated in accordance with AWPA Standard C15 and to the
appropriate retention level.
Treated wood suppliers provide Material Safety Data Sheets (MSDS) or Consumer
Information Sheets with the product. These sheets contain special instructions
about the care, handling and disposal of treated wood. Federal law dictates that
these sheets must be provided to all employees exposed to the materials.
Saw cuts or drilled holes made after treatment may expose untreated wood. This
problem is especially critical if the newly exposed wood is in the splash zone or
in contact with the ground. When using nail-laminated posts, the cut end of the
treated lumber should be placed upward, above the ground level; otherwise,
brushapplied, soaked, or dipped field treatments are recommended. AWPA Standard M4
outlines procedures for field treatment; some chemicals require a certified
pesticide applicators license to apply. The chemical suppliers should be consulted
for application restrictions.
4.3 Trusses
4.3.1 General. Together with posts, wood trusses are primary structural elements of
postframe buildings. Two excellent sources of technical information on trusses are
the Truss Plate Institute (TPI) and the Wood Truss Council of America (WTCA).
Trusses must be properly designed, handled and installed. These responsibilities
are shared by the building owner, contractor and designer, and the truss designer
and manufacturer. The importance of a clear understanding of responsibilities among
these parties cannot be overstated, and is covered in WTCA 1-1995 Standard
Responsibilities in the Design Process Involving Metal Plate Connected Wood Trusses
and ANSI/TPI-1-1995 National Design

Standard for Metal Plate Connected Wood Truss Construction.


4.3.2 Design Loads. The controlling load combination for truss design often is snow
plus dead load. The unbalanced snow load case should be checked per the applicable
building code, or for agricultural buildings, engineering practice ASAE EP288.5
(ASAE, 1999a) should be consulted. However, all other applicable load combinations
must be checked. For example, a wind load combination may cause stress reversal in
some truss elements as discussed later in this chapter.
Truss loads are normally represented by listing the top-chord live and dead, and
bottom-chord live and dead loads, respectively. Truss design loads are typically
expressed in units of pounds per square foot (psf). An example of truss loading
would be 20-4-0-1 (psf is implied). Both live and dead loads apply to the
vertically projected tributary areas of the top and bottom chords. Often, a bottom-
chord live load is not required, so the preceding nomenclature would be shortened
to 20-4-1 psf.
4.3.3 Design. This design manual does not present specifics of roof truss design.
Metalplate-connected wood trusses in the United States are designed according to
the provisions of ANSI/TPI 1-1995. Other designs are based on proprietary test
information, along with design criteria from the NDS (AF&PA, 1997). Model building
codes recognize either of these approaches.
ANSI-TPI 1-1995 mentions two types of structural analyses. The simplified method
is a type of pin joint analysis that has been calibrated to account for partial
joint fixity. This method uses tables of factors to determine chord moments and
member buckling lengths. The simplified method has been the predominate method for
a number of years; however, it will eventually be phased out by TPI. The other type
of analysis which is sometimes referenced as the exact method, is a stiffness
matrix method of analysis. Plane frame structures analyzers are becoming more
commonly used and provide for more sophisticated and accurate analyses. Regardless
of analysis methods, structural modeling assumptions are important and can
dramatically influence the design (Brakeman, 1994).

4-4

National Frame Builders Association

Post-Frame Building Design Manual

For example, partial fixity at truss plate joints as well as eccentricity at heel
joints, can be modeled a variety of ways. The heel joint usually gets the most
attention since heel joint modeling decisions can greatly influence truss design.
The size, and in some cases the orientation, of truss plates is dependent on
proprietary design values. These values are available from the manufacturers or
from research reports prepared by the model code agencies, Such as ICBO, SBCCI and
BOCA.
Trusses can be obtained pre-engineered from the manufacturer. It is important to
consider wind loading on trusses as stress reversals can occur and overstress some
members. This design is complicated by the fact that wind loads are influenced by
building geometry, so this information must be communicated to the truss designer.
Any structural bracing (e.g. knee braces) or redundant supports must be included in
the truss design.
4.3.4. Truss-to-Post Connection. The connection between the truss and post is
critical. Designers must consider both gravity forces and uplift forces. With some
truss-to-post connection designs, it might be necessary to examine the impact of
the connection on the forces induced in the truss chords, heel joints, and post.
Observations from several building investigations revealed that the individual
trusses and posts were designed properly, but the connection between the two units
was not. Many different methods and hardware have been used to design the
connection, such as bolts, nails, truss anchors, and combinations of the same.
Unless otherwise governed by a specific code, the design of this connection should
meet NDS (AF&PA, 1997) requirements.
4.3.5 Stress Reversal. The trusses used in post-frame buildings are typically long
span and, consequently, have long webs. When the truss becomes part of a post-frame
building, it is possible, under certain loading conditions, for a tension web in
the truss design to become a compression web.
Stress reversal can also occur in truss chords due to a wind uplift loading
combined with dead

load. This load case may not frequently control the size of the truss chord lumber,
but it makes compression in the bottom chord possible. This situation is one reason
that lateral bracing of the bottom chord is required (TPI, 1989; 1991a; 1991b).
4.3.6 Temporary Bracing. Temporary bracing is required to ensure stability of
trusses during their installation and until permanent bracing for trusses and the
building are in place. This area is the most difficult to manage in the field.
According to WTCA 1-1995 and ANSI/TPI 11995, determination and installation of
temporary bracing is the responsibility of the building contractor. Truss Plate
Institute (TPI) publication HIB-98 is a summary sheet that contains
recommendations for handling, installing and temporary bracing metal plate
connected wood trusses used in post-frame construction. Another TPI summary sheet
(i.e., HIB-91) contains similar recommendation for trusses with on center spacings
two feet or less and spans less than 60 feet. Both HIB-98 and HIB-91 are formatted
as accident-prevention brochures for use by builders, building contractors,
licensed contractors, erectors, and erection contractors.
4.3.7 Permanent Bracing. Permanent truss bracing is critical to the performance of
the roof system. Roof trusses are designed with the assumption that their elements
are held sufficiently in-plane (ANSI/TPI, 1995). The primary function of permanent
roof-truss bracing is to hold all trusses of the roof in the intended vertical
plane. HIB-98, provides guidance for the placement of temporary truss bracing,
which, if left in place, may function as part of the permanent bracing system.
Building designers are responsible for designing permanent bracing. For trusses
spaced 4 ft or less, DSB-89 (TPI, 1989) provides a calculation method for temporary
and permanent bracing designs. For trusses spaced greater than 4 ft (1.22 m) on-
center, similar principles can be used, but designers must consider that the longer
lengths involved may cause the bracing members to buckle. A commentary covering
permanent bracing of metal plate connected wood trusses is available from WTCA
(1999).

4-5

National Frame Builders Association

Post-Frame Building Design Manual

4.4 Girders
4.4.1 General. Girders are heavy beams used to span large openings (e.g., doors)
and to support trusses located between posts. For example, when roof truss spacing
is less than the post spacing, girders (sometimes called headers) are needed to
carry the intermediate trusses. This is a common occurrence over large door
openings. These beams are considered main wind-force resisting members. Vertically
nail-laminated lumber, structural composite lumber, glued-laminated beams and steel
Ibeams are all commonly used as girders. There is an abundant supply of structural-
composite lumber products from manufacturers who publish their own allowable
stresses. Often, the critical load combination is dead plus snow load, although all
applicable load combinations must be checked.
4.4.2 Design Criteria. Girders are designed as bending members. Any one of the four
criteria used for the design of bending members can control design (i.e. bending,
shear, compression perpendicular to grain, and deflection). Shear can often control
girder design. Also note that formulae found in most handbooks account for bending
but not shear deflection. Designers should consider the impact of shear deflection
on the total deflection of a girder. Hoyle and Woeste (1989) provide formulae for
calculating shear deflection of wood beams.
4.4.3 Vertically Laminated Lumber. The design of girders for a post-frame building
is routine structural design except when a girder is fabricated by vertically
laminating three or more pieces of dimension lumber. In this case, the allowable
bending stress can be increased using the repetitive member factors published in
ANSI/ASAE EP559 (ASAE, 1999b). These values are given in table 7.3.
4.4.4 Connections. Girder attachment to posts and individual roof trusses is a
fundamental part of girder design. When designing girder-to-post connections, both
uplift and gravity must be considered. When designing truss-to-girder connections,
special consideration must be given to situations in which trusses are hung off the
side of the girder. In such cases truss-togirder connections should be designed to
prevent rotation between the trusses and girder, or

the girder must be sized to handle additional stresses due to torsion.


4.5 Knee Braces
4.5.1 General. Knee braces are intended to supplement the resistance of post frames
under lateral loads, and can influence the unsupported length of columns. They have
been used less and less in recent years.
4.5.2 Effectiveness. Knee brace effectiveness is highly dependent on the stiffness
of its connections to the post and truss. If the connections at the ends of the
brace are flexible or not very stiff due to the use of a few nails, the roof
diaphragm carries the bulk of the load, and the brace is ineffective (Gebremedhin
and Woeste, 1986). If the brace connections are made very stiff (by installing many
nails or bolts) the brace could effectively resist the wind loading but could
overload the truss.
4.5.3 Analysis. Knee braces induce primary bending moments in truss chords if
attached between panel points. Knee braces induce secondary bending moments when
attached directly to panel points. If knee braces are to be used in a post-frame
design, load sharing among the truss, post, knee brace, connections, and diaphragm
(when applicable) must be included in the structural analysis.
4.6 Roof Purlins
4.6.1 General. Roof purlins are typically 2- by 4inch or 2- by 6-inch lumber, and
are key structural elements of the roof assembly. They resist gravity loads, wind
loads, roof diaphragm chord forces, and provide lateral bracing to truss top chords
(or rafters). To fulfill the chord-bracing role, the purlins must be supported
against lateral movement by attachment to sheathing or metal cladding that provides
the needed roof diaphragm strength. Not all roof cladding materials provide
diaphragm strength and/or purlin lateral support; one example is standing seam
roofing, which is fastened with clips that allow adjacent sheets to slide.
4.6.2 Classification. Purlins in post-frame buildings fall into the category of
component

4-6

National Frame Builders Association

Post-Frame Building Design Manual

and cladding, which is recognized by all three model building codes. Components
and cladding collect the loads and distribute them to the primary structural
elements, identified as the main wind-force resisting system. Wind loads are much
greater at eaves, ridges, edges, corners and other discontinuities. Purlin spacing
and fasteners are critical in these areas. If these areas fail under extreme wind
loading, the building envelope will be breached, and internal wind pressures will
change dramatically.
4.6.3 Orientation. Purlins are installed on-edge or flat. When they are used on-
edge, they may be either placed on top of the truss or recessed between the
trusses. Purlins placed on-edge are frequently overlapped and fastened together at
the overlap. When used flat, purlins are installed on top of the trusses.
4.6.4 Truss Chord Bracing. Purlin spacing is a factor in truss design since purlins
provide lateral support to the truss top chord. In some cases, the slenderness
ratio for weak-axis truss chord buckling between purlins can be greater than that
for strong-axis buckling. Therefore, when specifying trusses, the building designer
should inform the truss-design engineer of the planned purlin spacing.
4.6.5 Design Loads. Purlin design often is controlled by the dead plus snow load
combination, or dead plus wind load (especially in the edge zones of the roof).
Dead loads used for design may exceed actual weights for gravity load calculations;
however, inflated dead loads cannot be used to offset wind uplift or wind overturn
moments. In these cases, offsetting loads cannot exceed actual weights of
materials.
4.6.6 Design Criteria. Purlins members should be checked for bending strength,
shear capacity, and deflection. If the roof assembly is functioning as a structural
diaphragm, purlins will also be subjected to axial forces. Purlins shall be
designed to carry bending about both axes. Weak axis bending may be omitted if it
can be demonstrated by test or analysis that the roof sheathing provides support.
The connections between the purlins and rafters should be designed for both gravity
loads and wind uplift forces. Purlin hangers are often used when pur-

lins are recessed, and their capacity should be verified for the various loading
cases. In general, the provisions of the NDS (AF&PA, 1997) apply for the
connections and stress analysis.
4.7 Wall Girts
4.7.1 General. Girts are used to collect windinduced wall loads and distribute them
to the post frames. For end walls, the wind loads are distributed to structural
end-wall posts.
4.7.2 Classification. Girts belong to the component and cladding category for
determining the design wind load.
4.7.3 Orientation. Girts are either installed flat on post faces or recessed
between the posts. Girts recessed between posts are almost always orientated with
the narrow edge facing the cladding, and in this position, are frequently used to
support both interior and exterior cladding/sheathing.
4.7.4 Post Bracing. Girts provide lateral support to side-wall columns. With girts
securely installed, the slenderness ratio of the post weak axis is greatly reduced.
Therefore, posts can usually be designed to carry the axial loads using the
slenderness ratio of the strong axis.
4.7.5 Design Loads. Girts are normally designed to resist only wind load. Wind
loads are much greater at corners and other discontinuities. Girt spacing and
fasteners are critical in these areas. If these areas fail, the building envelope
will be breached, and internal wind pressures will change dramatically.
The dead load of the girt and attached steel is normally negligible for girt
design. Cladding is attached to the girts by nails or screws, and the stiffness of
these connections does not allow the girts to undergo significant bending stress or
deflection from the action of the small dead loads present. However, the wall dead
load should be included in total dead load calculations for the post foundation.
Girts must be design to resist forces induced by stored materials, especially
granular materials such as fertilizer or seeds/grain. Care should be

4-7

National Frame Builders Association

Post-Frame Building Design Manual

taken to assure that the capacity of wall panels, fasteners and girts are not
exceeded by these forces.
4.7.6 Design Criteria. Girts are designed as bending members for which the usual
bendingmember design criteria apply. The critical connections between the girts and
the post should be checked for both wind pressure and suction. The top wall girt
may be constructed to carry chord forces from the roof diaphragm and, if so, must
be checked for the appropriate axial loads. The NDS (AF&PA, 1997) provisions apply
for the connections and stress analysis.
4.8 Large Doors
4.8.1 General. Large doors are common in post-frame buildings. Door components must
be designed to withstand design wind loads, and are treated as components and
cladding for such calculations.
4.8.2 Open Doors. It is not uncommon for building owners to leave large doors open,
even during periods of high wind. If an owner anticipates that this will occur, the
building must be designed accordingly. Note that a large opening on one side of the
structure is generally associated with increased internal wind pressure
coefficients, and thus can significantly increase roof uplift forces.
4.9 Roof and Ceiling Diaphragms
4.9.1 General. Roof and ceiling diaphragms are used to resist lateral (sidesway)
forces applied to the building by wind, earthquake and stored material. Under
lateral load, roof and ceiling diaphragms act as large stiff plates. These plates
support and distribute loads to wall posts. Conceptually, diaphragm design is easy
to understand, but the application of the procedure requires analysis tools and
data.
Diaphragms made from plywood are well documented, as well as those made entirely
from steel. Less information is available about woodframed, metal-clad diaphragms
which are prevalent in the post-frame building industry. This is a major factor in
post-frame building design and is covered in more depth in Chapter 5.

4.9.2 Design Properties. Diaphragm performance depends on factors such as the


steel, steel sheet-to-sheet fasteners, steel-to-wood fasteners, and the wood frame.
There is no standard steel panel construction, so diaphragm strength and stiffness
depend on the specific construction used. Strength and stiffness data on laboratory
test panels are generally required to derive design values. Most post-frame
buildings have much greater spans than laboratory test panels; therefore, test data
must be extrapolated to prac-tical building sizes as explained in Chapter 6.
4.10 Shearwalls
4.10.1 General. A large portion of the shear forces induced in roof and ceiling
diaphragms is transferred to the building foundation by shearwalls. In many post-
frame buildings, the only walls available to transfer this shear are exterior walls
(i.e., endwalls and sidewalls). Where present, interior partition walls can be
designed to transfer additional shear.
4.10.2 Endwalls. Endwalls in post-frame buildings resist wind loads perpendicular
to the building end wall and simultaneously help transmit roof shears (due to
parallel-to-end wall wind components) to the ground. In the diaphragm design
procedure described in Chapter 5, maximum roof shears occur at the endwalls. The
roof shear is transferred into the top truss chord or rafter of the endwall,
through the endwall to the ground level, and finally to the ground by posts or to
posts connected to a concrete slab. In addition to shear forces, the end wall is
subject to overturning forces. Wirt et al. (1992) have published procedures for
analyzing and designing end-wall foundations.
4.10.3 Wall Openings. Allowances must be made for openings in shearwalls. One
common practice in post-frame construction is to place large doorways in the
building endwalls. Procedures for accounting for the opening and ways to reinforce
the remaining wall are given in Chapter 5.
4.10.4 Partitioning. Partitioning of the building into structural segments is one
method to reduce maximum roof shears and endwall shears. For example, if it is not
practical to reinforce an

4-8

National Frame Builders Association

Post-Frame Building Design Manual

endwall that has a large door installed, the alternative is to install a structural
partition in the center of the building. The structural partition must meet the
shear requirements delivered by the roof diaphragm. Buttresses, inside or outside
the walls, can be used to reduce the effective length of the building with respect
to maximum roof and end-wall shears.
4.11 References
ASAE. 1999a. ASAE EP 288.5: Agricultural building snow and wind loads. ASAE
Standards, 46th edition. ASAE, St. Joseph, MI.
ASAE. 1999b. ANSI/ASAE EP 559: Design requirements and bending properties for
mechanically-laminated posts. ASAE Standards, 46th edition. ASAE, St. Joseph, MI.
American Forest & Paper Association (AF&PA). 1997. National design specification
for wood construction. AF&PA, Washington, D.C.
American Wood-Preservers' Assc. (AWPA). 1995a. Wood for commercial-residential
construction. Preservative treatment by pressure process, C-15. In Book of
Standards. AWPA, Stevensville, MD.
American Wood-Preservers Assc. (AWPA). 1995b. Lumber, timbers, bridge ties, and
mine ties, pressure treatment, C2-90. In Book of Standards. AWPA, Stevensville, MD.
American Wood-Preservers' Assc. (AWPA). 1995c. Care of pressure treated wood
products, M4-90. In Book of Standards. AWPA, Stevensville, MD.
Brakeman, D.B. 1994. Which truss design method is the correct one? Peaks 16(1):1-3.
Gebremedhin, K.G., and F.E. Woeste. 1986. Diaphragm design with knee brace slip for
postframe buildings. Transactions of the American Society of Agricultural Engineers
23(2):538-542.
Hoyle, R.J. and F.E. Woeste. 1989. Wood technology in the design of structures.
Fifth edition. Iowa State University Press, Ames, IA.
Truss Plate Institute (TPI). 1989. Recommended design specifications for temporary
bracing of

metal plate connected wood trusses. DSB-89. TPI, Madison, WI.


Truss Plate Institute (TPI). 1998. HIB-98 summary sheet. TPI, Madison, WI.
Truss Plate Institute (TPI). 1995. ANSI/TPI 11995 National design standard for
metal plate connected wood truss construction. TPI, Madison, WI.
Wirt, D.L., F.E. Woeste, D.E. Kline and T.E. McLain. 1992. Design procedures for
post-frame end walls. Applied Engineering in Agriculture 8(1):97-105.
Wood Truss Council of America (WTCA). 1995. Standard responsibilities in the design
process involving metal plate connected wood trusses. WTCA 1-1995. WTCA, Madison,
WI.
Wood Truss Council of America (WTCA). 1999. Commentary for permanent bracing of
metal plate connected wood trusses. WTCA, Madison, WI.

4-9

National Frame Builders Association

Post-Frame Building Design Manual

4-10

National Frame Builders Association

Post-Frame Building Design Manual

Chapter 5: DIAPHRAGM DESIGN

5.1 Introduction
5.1.1 2-D Frame Analysis. Prior to the 1980s, the common method of analysis for
post-frame structures in agricultural, commercial and light industrial applications
was to consider the structure as a system of independently-acting, twodimensional
(2-D) post-frames. Although a 2-D frame analysis method works well for designing
frames under vertical loadings; it is often too conservative for designing
buildings against sidesway. In addition, many 2-D frames offer little or no
resistance to loads acting normal to the frames (e.g., wind acting normal to the
endwalls).
5.1.2 Diaphragm Action. A considerable portion of the horizontal load applied to
many postframe structures is actually resisted by roof and ceiling diaphragms and
shearwalls. As previously stated (section 4.9), roof and ceiling diaphragms are
large plates that are formed when cladding is attached to roof and ceiling framing,
respectively. These large plates help redistribute load throughout the structure.
This redistribution of load by the diaphragms is called diaphragm action. A
shearwall is any wall interior or exterior with a measurable amount of racking
resistance. Most of the load to which a diaphragm is subjected, is transferred to
the foundation by shearwalls orientated parallel to the direction of applied load.
Figure 5.1 illustrates a situation in which wind load directed at a sidewall, is
transferred via the roof diaphragm to the endwalls and one interior wall. Under
this loading, the two endwalls and the one interior wall function as shearwalls.
When the same wind load is directed toward the endwall, the sidewalls function as
shearwalls in transferring the load from the roof diaphragm to the foundation
system.
5.1.3 Post-Frame Contributions. Whenever load is applied normal to the sidewall of
a structure, any post-frame with measurable racking resistance functions like the
interior shearwall in figure 5.1. The amount of load that an individual post-frame
will transfer to the foundation is dependent on (1) the in-plane shear stiffness of
the diaphragm, and (2) the racking stiffness of the

post-frame relative to that of other post-frames and shearwalls. If a diaphragm is


constructed in such a way that it is quite stiff in shear, diaphragm action will be
enhanced and the diaphragm will transfer load from post-frames with low racking
stiffness to shearwalls and postframes with high racking stiffness. However, if the
shear stiffness of the diaphragm is relatively low, load transfer will be minimal
and the behavior of the structure will be much more in accordance with the
assumption of independently acting post-frames.
Eave displacement
Wind load
Intermediate shearwall
Roof diaphragm End shearwall
Deformed structure Undeformed structure
Figure 5.1. Example of diaphragm action in which the roof diaphragm transfers load
to three shearwalls one interior and two exterior walls.
5.1.4 Endwall Loadings. Virtually all postframe buildings are longer than they are
wide. It follows, that diaphragms in such buildings, when viewed from the endwall,
appear as narrow, deep plates. For endwall loadings, these narrow, deep diaphragms
are generally assumed to have an infinite shear stiffness, which means that every
structural element attached to the diaphragm, shifts the same amount when the
diaphragm shifts without rotating. For example, under an endwall loading, the roof
diaphragm would ensure equal displacement of the top of endwall posts and the top
of each sidewall.

5-1

National Frame Builders Association

Post-Frame Building Design Manual

5.1.5 Diaphragm Design. When diaphragm action is accounted for in overall building
design, the design process is referred to as diaphragm design. Diaphragm design is
a relatively straight forward process when a diaphragm is (1) assumed to have
infinite shear stiffness, and/or (2) only attached to two shearwalls/post-frames
(as is generally the case with endwall loadings). When neither of these conditions
applies (generally true with loads normal to the sidewall) diaphragm design is more
complex.
5.1.6 ASAE EP484.2. The current diaphragm design procedure is outlined in ANSI/ASAE
EP484.2: Diaphragm Design of Metal-Clad, Wood-Frame Rectangular Buildings (ASAE,
1999a). This procedure, which is outlined in the following sections, can be broken
into five steps: Step 1. Construct a finite element model of
the building by breaking the structure into frame, shearwall, and diaphragm
elements (Section 5.2) Step 2. Assign stiffness values to frames and shearwall
elements (Section 5.3) and diaphragm elements (Section 5.4). Step 3. Calculate
structural loads (i.e., eave loads) for the model (Section 5.5). Step 4. Determine
the distribution of load to individual elements (Section 5.6). Step 5. Check to
make sure that loads do not exceed allowable values (Section 5.7).
5.2 Structural Model
5.2.1 General. The model developed in this section is only applicable for
determining the distribution of loads that are applied parallel to individual post-
frames (a.k.a., primary frames). As previously stated, an individual post-frame
consists of an individual truss and any attached posts.
5.2.2 Diaphragm Sectioning. The process of modeling a post-frame building for
diaphragm design begins with the dividing of individual roof and ceiling diaphragms
into sections, herein referred to as diaphragm sections. Diaphragm sectioning is a
straight-forward process with interior post-frames, interior shearwalls, ridge
lines and any other abrupt changes in roof and ceiling slopes servings as lines of
demarcation between diaphragm sections.

(a) Diaphragm "a"


Diaphragm "b"
Diaphragm "c"
(b) 1 2 3 45
1a 2a 3a 4a 1b 2b 3b 4b
(c) 1 2 3 45
1c 2c 3c 4c
(d) Figure 5.2. (a) Perspective view of a four-bay post-frame building with (b)
roof and ceiling diaphragms. Sectioning of (c) roof diaphragms, and (d) ceiling
diaphragm.

5-2

National Frame Builders Association

Post-Frame Building Design Manual

Figures 5.2a shows a post-frame building with three interior post-frames. Drawing a
line along each interior frame and the ridge results in the eight (8) roof
diaphragm sections shown in figure 5.2c, and the four ceiling diaphragm sections
shown in figure 5.2d.
To avoid confusion when assigning properties to diaphragm sections, it is helpful
to identify each diaphragm section with a two-digit identifier. The first digit
identifies the bay associated with the section. Bays are generally numbered from
leftto-right, as shown in figures 5.2c and 5.2d. The second digit identifies the
specific roof or ceiling slope. In figure 5.2, letters have been used to identify
these slopes, with letters a and b representing different roof slopes, and
letter c used to identify ceiling sections.
5.2.3 Discretization. The process of breaking a structure into elements for
analysis is referred to as discretization. For diaphragm design, a structure is
broken into frame elements and diaphragm elements. Each post-frame is considered a
separate frame element, as is each shearwall orientated in the same direction as
the post-frames. The example building shown in figure 5.2 would be modeled with
five (5) frame elements. These frame elements have been identified in figures 5.2c
and 5.2d with the encircled numbers (as with individual bays, numbering is
generally from left-to-right). Each diaphragm element models the diaphragm sections
within a single bay. For example, diaphragm sections 1a, 1b, and 1c in figure 5.2
would be represented with a single diaphragm element. It follows that the number of
diaphragm elements is equal to the number of building bays, which in turn, is one
less than the number of frame elements. Discretization of a four-bay building is
shown in figure 5.3a.
5.2.4 Spring Model. To determine the distribution of horizontally applied loads to
individual diaphragm and frame elements requires only a single stiffness property
for each element. For this reason, diaphragm and frame elements are generally
represented with simple springs. As shown in figure 5.3b, frame elements are
represented with springs of stiffness, k, and diaphragm elements are represented as
springs with stiffness Ch. The element (or spring) connection points (a.k.a. nodes)
are taken to repre-

sent locations at the eave of each frame/shearwall. Horizontal components of


applied building loads are typically uniformly distributed along the length of the
building as shown in figure 5.3a. For modeling purposes, this uniform load is
converted into a set of equivalent concentrated loads that are applied at the nodes
as shown in figure 5.3b. Because of the location of their application, these forces
are referred to as eave loads.
12345
1 234
(a)
k1 k2 k3 k4 k5
C C C Ch1 h2 h3 h4 r1 r2 r3 r4 r5
(b)
Figure 5.3. (a) Top view of a four-bay building showing individual elements and
applied horizontal loads. Encircled numbers identify frame elements, other numbers
identify diaphragm elements. (b) Corresponding spring model.

5-3

National Frame Builders Association

Post-Frame Building Design Manual

5.3 Frame Stiffness, k

5.3.1 Definition. To be compatible with a model in which nodes represent points


along the eave line (figure 5.3b), frame element stiffness, k, must equal the force
required for a unit displacement of the frame at the eave (figure 5.4). In equation
form:

k = P/

(5-1)

where:

k = frame stiffness, lbf/in (N/mm) P = load applied at eave, lbf (N) = lateral
displacement at eave result-
ing from applied load P, in (mm)

k=P/

Figure 5.4. Definition of frame stiffness, k.

5.3.2 Calculation. Frame stiffness is generally obtained with a plane-frame


structural analysis program, e.g., PPSA (Purdue Research Foundation, 1993), METCLAD
(Gebremedhin, 1987b), and SOLVER (Gebremedhin, 1987a). For post-frames in which (1)
all posts are assumed to be pin-connected to the truss (or rafters), and (2) there
are no special members (e.g., knee braces) connecting posts to the truss, frame
stiffness can be calculated as:

n
k = kp,i
i=1

(5-2)

where:

kp,i = stiffness of post i, lbf/in (N/mm) n = number of posts in the post-frame

Post stiffness, kp, is graphically defined in figure 5.5. For a post with a
constant flexural rigidity (E x I) that is assumed to be fixed at the base, post
stiffness is given as:

kp = 3 E I / Hp3

(5-3)

where:

kp =
E= I= Hp =

stiffness of post that is fixed at the base and pinned at the top, lbf/in (N/mm)
modulus of elasticity of post, lbf/in2 (N/mm2) moment of inertia of post, in4 (mm4)
post height from fixed base to truss connection post (see figure 5.5), in (mm)

P
Post -to-truss connection Hp point
kp = P /

Figure 5.5. Definition of post stiffness, kp.


5.3.3 Shearwalls. End shearwalls and intermediate shearwalls, like post-frames, are
modeled as frame elements (see Section 5.2.3). Consequently, their stiffness, like
that for post-frames, is defined as the ratio of a horizontal force, P, applied at
the eave of the wall, to the resulting horizontal eave displacement, .
The stiffness of shearwalls can be obtained using validated structural models, or
from tests of functionally equivalent assemblies. ASAE EP558 (1999b) gives
laboratory test procedures that can be used to determine the stiffness of
functionally equivalent walls. This topic is also discussed in Section 6.5.

5-4

National Frame Builders Association

Post-Frame Building Design Manual

Technical Note Embedded Post Analogs


When a post is embedded in the soil, calculated post stiffness (and consequently
calculated frame stiffness) is highly dependent on how the embedded portion is
modeled. Traditionally, engineers have ignored soil properties and have modeled
embedded posts using the analogs shown in figures 5.6a and 5.6b. An inherent
deficiency of these analogs is that the pin supports used to fix the post below
grade do not allow the post to naturally displace. For this reason, post stiffness
values predicted using the analogs should be applied with caution. It should also
be noted that the analogs in figures 5.6a and 5.6b produce a reduced post stiffness
when depth of embedment, d, is increased. In reality, anytime a post is embedded
deeper into the ground, the stiffness associated with the post increases.
To accurately model post movement below grade requires accounting for soil
stiffness. Bohnhoff (1992) developed equations for predicting post stiffness
assuming soil stiffness increased linearly with depth below grade and inversely
with post width. Bohnhoff also assumed that the post had infinite flexural
stiffness

below grade. Meador (1997) developed similar equations, but unlike Bohnhoff, Meador
assumed that soil stiffness was not a function of post width. Meador also
investigated the assumption of infinite post stiffness below grade, and established
limits for applicability of the equations he developed. McGuire (1998) used the
work of both Bohnhoff and Meador to propose an analog where soil is modeled as a
series of linear springs whose stiffness increases linearly below grade (figure
5.6c). McGuire verified Bohnhoffs results and also showed that for the case of
non-constrained posts, analogs like those shown in figure 5.6a may incorrectly
predict the sense of base moment (see Chapter 8).
Current impediments to the wide spread adoption of analogs that account for soil
stiffness include: (1) complexity of equations, and (2) unrealistically low post
stiffness values obtained using published soil stiffness data.
It is important for the post-frame designer to realize that fixing the post at
grade (figure 5.5) generally produces conservative values for post base moments,
especially for the nonconstrained post case. Conversely, forces calculated in the
diaphragm using this model might be non-conservative.

Hp
0.34 d d 0.1 d

Ground surface

Hp

0.7 d d

Hp

Floor slab

Springs used to model soil stiffness

(a) (b)

(c)

Figure 5.6. Structural analog traditionally used for (a) non-constrained and (b)
constrained posts. (c) A more realistic non-constrained post analog that accounts
for soil stiffness.

5-5

National Frame Builders Association


5.4 Diaphragm Stiffness, Ch
5.4.1 Definition. As shown in figure 5.7, the stiffness of a diaphragm element is
the horizontal load required to cause a unit shift (in a direction parallel to the
trusses/rafters) of the roof/ceiling assembly over a frame spacing (a.k.a. bay
width), s. This stiffness is commonly referred to as the total horizontal shear
stiffness, Ch, of the diaphragm.
12 3 4

s1 s2 s3 (a)
PP

s4

Post-Frame Building Design Manual

Ch = ch,i =
n=

total horizontal shear stiffness of diaphragm element, lbf/in (N/mm) horizontal


shear stiffness of diaphragm section i (from Section 6.4.4), lbf/in (N/mm) number
of diaphragm sections comprising the diaphragm element

5.5 Eave Loads, R

5.5.1 Definition. For diaphragm design, building loads are replaced by an


equivalent set of horizontally acting, concentrated (i.e., point) loads. These
loads are located at the eave of each frame element (i.e., post-frame and end
shearwall, and intermediate shearwall) and therefore are referred to as eave loads.
Eave loads and applied building loads are equivalent when they horizontally
displace the eave an equal amount.
Roof Gravity Loads

s x qwr

s x qlr

i
C h,i = P /

Ceiling Gravity Loads

s xqww s xqlw

si

(b)

Figure 5.7. (a) Top view of a four-bay building. (b) Definition of diaphragm
stiffness, Ch, for a single diaphragm element.
5.4.2 Calculation. The total horizontal shear stiffness of a diaphragm element is
simply equal to the sum of the horizontal shear stiffness values of the diaphragm
sections that comprise the element. In equation form:

Ch =

n
ch,i
i=1

(5-4)

where:
Figure 5.8. Typical structural analog for obtaining eave load, R.
5.5.2 Calculation by Plane-Frame Structural Analysis. A horizontal restraint
(vertical roller) is placed at the eave line as shown in figure 5.8 and the
structural analog is analyzed with all external loads in place. The horizontal
reaction at the vertical roller support is numerically equal to the eave load, R.
The vertical roller should always be placed at the same location that horizontal
load P was placed when determine frame stiffness (see figure 5.4). The value of R
is very dependent on the magnitude of forces with

5-6

National Frame Builders Association

horizontal components (i.e., wind and stored materials).

5.5.3 Calculation Using Frame-Base Fixity Factors. When: (1) posts are assumed to
be pin-connected to trusses/rafters, (2) the only applied loads with horizontal
components are due to wind, and (3) wind pressure is uniformly distributed on each
wall and roof surface, then eave load, R, can be estimated as:

R = s (hwr qwr hlr qlr + hww fw qww hlw fl qlw)


where:

(5-5)

R= s=
=
hwr = hlr = hww = hlw = qwr =
qlr =
qww =
qlw =
fw =
fl =

eave load, lbf (N) frame spacing for interior postframes and shearwalls, ft (m)
one-half the frame spacing for endwalls, ft (m) windward roof height, ft (m)
leeward roof height, ft (m) windward wall height, ft (m) leeward wall height, ft
(m) design windward roof pressure, lbf/ft2 (N/m2) design leeward roof pressure,
lbf/ft2 (N/m2) design windward wall pressure, lbf/ft2 (N/m2) design leeward wall
pressure lbf/ft2 (N/m2) frame-base fixity factor, windward post frame-base fixity
factor, leeward post

Inward acting wind pressures have positive signs, outward acting pressures are
negative (figure 5.8). In buildings with variable frame spacings, set s equal to
the average of the frame spacings on each side of the eave load.

Frame-base fixity factors, fw and fl, determine how much of the total wall load is
transferred to the eave, and how much is transferred directly to the ground. The
greater the resistance to rotation at the base of a wall, the more load will be
attracted directly to the base of the wall. For substantial fixity against rotation
at the groundline, set the frame-base fixity factor(s) equal to 3/8. For all other
cases, set the frame-base fixity factor(s) equal to 1/2.

Post-Frame Building Design Manual

For symmetrical base restraint and frame geometry, equation 5-5 reduces to:

R = s [hr (qwr qlr) + hw f (qww qlw)] (5-6)


where:

hr = hw = f=

roof height, ft (m) wall height, ft (m) frame-base fixity factor for both leeward
and windward posts

5.6 Load Distribution

5.6.1 General. The distribution of horizontal loads to frames, shearwalls, and


various diaphragm sections can be determined after stiffness values have been
assigned to each frame and diaphragm element, and eave loads have been established.
5.6.2 Analysis Tools. Any finite element or plane-frame structural analysis program
can be used to analyze the structural model shown in figure 5.3b. However, to
expedite this process, computer program DAFI was developed (Bohnhoff, 1992). Once
eave loads and frame and diaphragm element stiffness values are input, DAFI
calculates eave displacements, frame element loads and diaphragm element shear
forces. DAFI can be downloaded at no cost from the NFBA web site
(http://www.postframe.org/).
An iterative method for hand-calculating load distribution was developed by
Anderson and others (1989). This method, which is referred to as the force
distribution method, is procedurally identical to the classical method of moment
distribution.
5.6.3 mS and mD Tables. Forces in the most highly loaded diaphragm and frame
elements, can be calculated using tables 5.1 and 5.2 when all five of the following
conditions exist: (1) all diaphragm elements have the same stiffness Ch, (2) all
interior frame elements have the same stiffness, k, (3) both exterior frame
elements (i.e., the two elements representing the endwalls) have the same
stiffness, ke, (4) eave load, R, is the same at each interior frame, and (5) the
eave load for each exterior frame is equal to one-half that for an interior frame.
These five requirements are generally met in buildings with a fixed bay spacing,
endwalls that are virtually identical in construction, and interior frames that

5-7

National Frame Builders Association

dont vary in overall design. When tables 5.1 and 5.2 are applicable, the analysis
tools discussed in Section 5.6.2 are generally not needed.

Input parameters required for tables 5.1 and 5.2 include: number of frame elements
(i.e., the number of interior frames + 2); ratio of diaphragm element to interior
frame element stiffness, Ch / k; and ratio of exterior to interior frame element
stiffness, ke / k.

The most highly loaded diaphragm element (in any building that meets the preceding
five conditions) is the element located adjacent to the endwalls. The maximum shear
force in this diaphragm element, Vh, is equal to the appropriate shear modifier
value, mS, from table 5.1, multiplied by the eave load, R, for an interior frame.
In equation form:

Vh = R mS

(5-7)

where:

Vh =
mS = R=

maximum diaphragm element shear force, lbf (N) shear force modifier from Table 5.1
eave load at interior frame, lbf (N)

The value obtained from equation 5-7 is simply equal to one-half of the total
horizontal eave load that is not carried by the interior frames.

The most highly loaded interior frame element (in any building that meets the
preceding five conditions) is the element located closest to the building
midlength. Because of diaphragm action, the total horizontal load that this
critical frame must resist is reduced from that which it would have to resist
without diaphragm action. The magnitude of this reduction is referred to the
horizontal restraining force because in reality, it is a restraining force applied
to the frame by the roof (and/or ceiling) diaphragms. Numerically, the horizontal
restraining force, Q, is equal to the product of the eave load R, and the
appropriate sidesway restraining force factor, mD from table 5.2. In equation form:

Q = R mD

(5-8)

where:

Post-Frame Building Design Manual

Q= mD =
R=

sidesway restraining force, lbf (N) sidesway restraining force factor from Table
5.2 eave load at interior frame, lbf (N)

5.6.4 In-Plane Shear Force in a Diaphragm Section, Vp. The analysis tools discussed
in Section 5.6.2 (and equation 5-7) output diaphragm element forces. In most cases,
each element is comprised of two or more diaphragm sections. The in-plane shear
force in each of these diaphragm sections is calculated as:

Vp,i = (ch,i / Ch) Vh / (cos i) where:

(5-9)

Vp,i = Vh = ch,i = i =

in-plane shear force in diaphragm section i, lbf (N) horizontal shear force in the
diaphragm element, lbf (N) horizontal shear stiffness of diaphragm section i,
lbf/in. (N/mm)
slope of diaphragm section i

5.6.5 Forces Applied to Frames by Individual Diaphragms. The horizontal movement of


most building frames is resisted by roof/ceiling diaphragms. The total horizontal
resisting force applied to an individual frame by the roof/ceiling diaphragms was
previously defined as the sidesway restraining force, Q. To accurately model a
frame with the resisting forces applied by the roof and ceiling diaphragms,
requires that the sidesway restraining force, Q, first be divided up between the
individual diaphragms (e.g., diaphragms a, b, and c in figure 5.2b). This is
accomplished using the following equation:

Q i = Q (ch,i / Ch) where:


(5-10)

Qi = Q= Ch =
ch,i =

sidesway resisting force due to diaphragm i, lbf (N) total sidesway resisting force
acting on the frame, lbf (N) horizontal shear stiffness for a width s of the
roof/ceiling assembly, lbf/in. (N/mm) horizontal shear stiffness of diaphragm i
with width s, lbf/in. (N/mm)

5-8

National Frame Builders Association

Post-Frame Building Design Manual

Table 5.1. Shear Force Modifier (mS)

ke / k

Ch / k

Number of frames (endwalls are counted as frames)

3 4 5 6 7 8 9 10 11 12 13 14 15 16

5 5 0.88 1.14 1.33 1.45 1.53 1.59 1.62 1.65 1.66 1.67 1.68 1.68 1.68 1.68 5 10 0.89
1.19 1.42 1.59 1.72 1.82 1.89 1.94 1.98 2.00 2.02 2.04 2.05 2.06 5 20 0.90 1.22
1.48 1.68 1.85 1.98 2.08 2.16 2.23 2.29 2.33 2.36 2.39 2.41 5 50 0.91 1.24 1.51
1.74 1.93 2.10 2.23 2.35 2.45 2.53 2.60 2.67 2.72 2.77 5 100 0.91 1.24 1.53 1.76
1.97 2.14 2.29 2.42 2.53 2.63 2.72 2.80 2.87 2.93 5 200 0.91 1.25 1.53 1.77 1.98
2.16 2.32 2.46 2.58 2.69 2.79 2.87 2.95 3.02 5 500 0.91 1.25 1.54 1.78 1.99 2.18
2.34 2.48 2.61 2.73 2.83 2.92 3.01 3.08 5 1000 0.91 1.25 1.54 1.78 2.00 2.18 2.35
2.49 2.62 2.74 2.84 2.94 3.02 3.10 5 10000 0.91 1.25 1.54 1.79 2.00 2.19 2.35 2.50
2.63 2.75 2.86 2.95 3.04 3.12

10 5 0.91 1.23 1.46 1.62 1.73 1.81 1.86 1.89 1.91 1.92 1.93 1.93 1.94 1.94 10 10
0.93 1.29 1.58 1.81 1.99 2.13 2.23 2.31 2.36 2.40 2.44 2.46 2.48 2.49 10 20 0.94
1.33 1.66 1.94 2.17 2.36 2.52 2.66 2.76 2.85 2.92 2.98 3.03 3.06 10 50 0.95 1.35
1.70 2.02 2.30 2.55 2.76 2.96 3.12 3.27 3.40 3.51 3.61 3.70 10 100 0.95 1.36 1.72
2.05 2.35 2.62 2.86 3.08 3.27 3.45 3.61 3.76 3.89 4.01 10 200 0.95 1.36 1.73 2.07
2.37 2.65 2.91 3.14 3.36 3.56 3.74 3.90 4.06 4.20 10 500 0.95 1.36 1.74 2.08 2.39
2.68 2.94 3.19 3.41 3.62 3.82 4.00 4.17 4.32 10 1000 0.95 1.36 1.74 2.08 2.40 2.68
2.95 3.20 3.43 3.64 3.84 4.03 4.20 4.37 10 10000 0.95 1.36 1.74 2.08 2.40 2.69 2.96
3.21 3.45 3.66 3.87 4.06 4.24 4.41

20 5 0.93 1.28 1.54 1.73 1.85 1.94 2.00 2.03 2.06 2.07 2.09 2.09 2.10 2.10 20 10
0.95 1.35 1.68 1.95 2.16 2.33 2.45 2.55 2.62 2.67 2.71 2.74 2.76 2.78 20 20 0.96
1.39 1.76 2.09 2.38 2.62 2.83 3.00 3.14 3.25 3.35 3.43 3.49 3.54 20 50 0.97 1.41
1.82 2.20 2.54 2.85 3.14 3.39 3.62 3.83 4.01 4.17 4.32 4.44 20 100 0.97 1.42 1.84
2.23 2.60 2.95 3.26 3.56 3.83 4.09 4.32 4.54 4.74 4.92 20 200 0.97 1.42 1.85 2.25
2.63 2.99 3.33 3.65 3.95 4.24 4.50 4.75 4.99 5.21 20 500 0.98 1.43 1.86 2.27 2.65
3.02 3.38 3.71 4.03 4.33 4.62 4.90 5.16 5.41 20 1000 0.98 1.43 1.86 2.27 2.66 3.03
3.39 3.73 4.06 4.37 4.66 4.95 5.22 5.48 20 10000 0.98 1.43 1.86 2.27 2.67 3.04 3.40
3.75 4.08 4.40 4.70 5.00 5.28 5.55

50 5 0.95 1.31 1.59 1.79 1.93 2.03 2.09 2.14 2.16 2.18 2.19 2.20 2.20 2.21 50 10
0.97 1.38 1.74 2.04 2.28 2.46 2.61 2.72 2.80 2.86 2.91 2.94 2.97 2.99 50 20 0.98
1.43 1.83 2.20 2.52 2.80 3.04 3.25 3.41 3.55 3.67 3.77 3.84 3.91 50 50 0.99 1.45
1.90 2.32 2.71 3.08 3.42 3.73 4.01 4.26 4.50 4.70 4.89 5.06 50 100 0.99 1.46 1.92
2.36 2.78 3.18 3.57 3.93 4.27 4.60 4.90 5.18 5.45 5.69 50 200 0.99 1.47 1.93 2.38
2.82 3.24 3.65 4.04 4.42 4.79 5.14 5.47 5.79 6.09 50 500 0.99 1.47 1.94 2.40 2.84
3.28 3.70 4.12 4.52 4.91 5.29 5.66 6.02 6.37 50 1000 0.99 1.47 1.94 2.40 2.85 3.29
3.72 4.14 4.55 4.96 5.35 5.73 6.11 6.47 50 10000 0.99 1.47 1.94 2.40 2.86 3.30 3.74
4.16 4.58 5.00 5.40 5.80 6.19 6.57

100 5 0.95 1.32 1.61 1.82 1.96 2.06 2.13 2.17 2.20 2.22 2.23 2.24 2.24 2.25 100 10
0.97 1.40 1.76 2.07 2.32 2.51 2.67 2.78 2.87 2.93 2.98 3.02 3.05 3.06 100 20 0.98
1.44 1.86 2.24 2.58 2.87 3.12 3.34 3.52 3.67 3.79 3.89 3.98 4.05 100 50 0.99 1.47
1.92 2.36 2.77 3.16 3.52 3.85 4.16 4.43 4.69 4.91 5.12 5.30 100 100 0.99 1.48 1.95
2.40 2.85 3.27 3.68 4.07 4.44 4.79 5.13 5.44 5.73 6.01 100 200 0.99 1.48 1.96 2.43
2.89 3.33 3.77 4.19 4.61 5.00 5.39 5.76 6.12 6.46 100 500 1.00 1.48 1.97 2.44 2.91
3.37 3.83 4.27 4.71 5.14 5.56 5.98 6.38 6.78 100 1000 1.00 1.48 1.97 2.45 2.92 3.39
3.85 4.30 4.75 5.19 5.62 6.05 6.48 6.89 100 10000 1.00 1.49 1.97 2.45 2.93 3.40
3.86 4.32 4.78 5.23 5.68 6.12 6.56 7.00

1000 1000 1000 1000 1000 1000 1000 1000 1000

5 10 20 50 100 200 500 1000 10000

0.95 0.98 0.99 1.00 1.00 1.00 1.00 1.00 1.00

1.33 1.41 1.45 1.48 1.49 1.49 1.50 1.50 1.50

1.63 1.78 1.88 1.95 1.97 1.99 1.99 2.00 2.00

1.84 2.10 2.28 2.40 2.45 2.47 2.49 2.49 2.50

1.99 2.36 2.63 2.83 2.91 2.95 2.98 2.98 2.99

2.09 2.56 2.93 3.24 3.36 3.42 3.46 3.48 3.49

2.16 2.72 3.20 3.62 3.79 3.89 3.95 3.97 3.98

2.20 2.84 3.43 3.97 4.21 4.34 4.42 4.45 4.48

2.23 2.93 3.62 4.30 4.61 4.78 4.90 4.94 4.97

2.25 3.00 3.78 4.60 4.99 5.22 5.37 5.42 5.47

2.27 3.05 3.91 4.87 5.35 5.64 5.83 5.90 5.96

2.27 3.09 4.02 5.12 5.69 6.05 6.29 6.37 6.45

2.28 3.12 4.11 5.34 6.02 6.44 6.74 6.85 6.94

2.28 3.14 4.18 5.54 6.32 6.83 7.18 7.31 7.43

10000 10000 10000 10000 10000 10000 10000 10000 10000

5 10 20 50 100 200 500 1000 10000

0.96 1.33 1.63 1.84 1.99 2.09 2.16 2.21 2.24 2.26 2.27 2.28 2.28 0.98 1.41 1.79
2.10 2.36 2.57 2.72 2.85 2.94 3.01 3.06 3.10 3.12 0.99 1.45 1.89 2.28 2.63 2.94
3.21 3.43 3.63 3.79 3.92 4.03 4.12 1.00 1.48 1.95 2.40 2.84 3.25 3.63 3.98 4.31
4.61 4.89 5.14 5.36 1.00 1.49 1.98 2.45 2.92 3.37 3.80 4.22 4.62 5.01 5.37 5.72
6.05 1.00 1.50 1.99 2.48 2.96 3.43 3.90 4.35 4.80 5.24 5.66 6.08 6.48 1.00 1.50
2.00 2.49 2.98 3.47 3.96 4.44 4.92 5.39 5.86 6.32 6.78 1.00 1.50 2.00 2.50 2.99
3.49 3.98 4.47 4.96 5.44 5.93 6.41 6.88 1.00 1.50 2.00 2.50 3.00 3.50 4.00 4.50
4.99 5.49 5.99 6.49 6.98

2.29 3.14 4.19 5.57 6.35 6.87 7.23 7.36 7.48

5-9

National Frame Builders Association

Post-Frame Building Design Manual

Table 5.1. Shear Force Modifier (mS), cont.

ke / k

Ch / k

Number of frames (endwalls are counted as frames)

17 18 19 20 21 22 23 24 25 26 27 28 29 30

5 5 1.69 1.69 1.69 1.69 1.69 1.69 1.69 1.69 1.69 1.69 1.69 1.69 1.69 1.69 5 10 2.06
2.07 2.07 2.07 2.07 2.08 2.08 2.08 2.08 2.08 2.08 2.08 2.08 2.08 5 20 2.43 2.44
2.46 2.46 2.47 2.48 2.48 2.49 2.49 2.49 2.49 2.49 2.50 2.50 5 50 2.81 2.84 2.87
2.89 2.92 2.94 2.95 2.97 2.98 2.99 3.00 3.01 3.01 3.02 5 100 2.98 3.03 3.07 3.11
3.14 3.18 3.20 3.23 3.25 3.27 3.29 3.30 3.32 3.33 5 200 3.09 3.14 3.19 3.24 3.28
3.32 3.36 3.39 3.42 3.45 3.48 3.50 3.52 3.54 5 500 3.15 3.22 3.28 3.33 3.38 3.43
3.47 3.51 3.55 3.58 3.61 3.64 3.67 3.70 5 1000 3.18 3.24 3.30 3.36 3.41 3.46 3.51
3.55 3.59 3.63 3.66 3.70 3.73 3.75 5 10000 3.20 3.27 3.33 3.39 3.45 3.50 3.54 3.59
3.63 3.67 3.71 3.74 3.78 3.81

10 5 1.94 1.94 1.94 1.94 1.94 1.94 1.94 1.94 1.94 1.94 1.94 1.94 1.94 1.94 10 10
2.50 2.50 2.51 2.51 2.51 2.52 2.52 2.52 2.52 2.52 2.52 2.52 2.52 2.52 10 20 3.09
3.12 3.14 3.15 3.16 3.17 3.18 3.19 3.19 3.20 3.20 3.20 3.21 3.21 10 50 3.77 3.84
3.89 3.94 3.99 4.02 4.06 4.09 4.11 4.13 4.15 4.17 4.18 4.19 10 100 4.12 4.21 4.30
4.38 4.45 4.52 4.58 4.63 4.68 4.72 4.76 4.80 4.83 4.86 10 200 4.33 4.45 4.56 4.66
4.76 4.84 4.92 5.00 5.07 5.13 5.19 5.25 5.30 5.35 10 500 4.47 4.61 4.74 4.86 4.97
5.08 5.18 5.27 5.36 5.44 5.52 5.60 5.67 5.73 10 1000 4.52 4.66 4.80 4.93 5.05 5.16
5.27 5.37 5.47 5.56 5.65 5.73 5.81 5.88 10 10000 4.57 4.72 4.86 4.99 5.12 5.24 5.36
5.47 5.57 5.67 5.76 5.86 5.94 6.03

20 5 2.10 2.10 2.10 2.10 2.10 2.10 2.10 2.10 2.10 2.10 2.10 2.10 2.10 2.10 20 10
2.79 2.80 2.80 2.81 2.81 2.81 2.82 2.82 2.82 2.82 2.82 2.82 2.82 2.82 20 20 3.58
3.62 3.64 3.66 3.68 3.69 3.71 3.71 3.72 3.73 3.73 3.74 3.74 3.74 20 50 4.56 4.65
4.74 4.82 4.88 4.94 4.99 5.03 5.07 5.11 5.14 5.16 5.18 5.20 20 100 5.08 5.24 5.38
5.51 5.62 5.73 5.83 5.91 5.99 6.07 6.13 6.20 6.25 6.30 20 200 5.42 5.61 5.80 5.97
6.13 6.28 6.42 6.55 6.67 6.79 6.90 7.00 7.09 7.18 20 500 5.65 5.88 6.09 6.30 6.50
6.69 6.87 7.04 7.20 7.36 7.51 7.65 7.78 7.91 20 1000 5.73 5.97 6.20 6.42 6.64 6.84
7.03 7.22 7.40 7.58 7.74 7.90 8.06 8.21 20 10000 5.81 6.06 6.30 6.54 6.77 6.98 7.20
7.40 7.60 7.79 7.97 8.15 8.33 8.50

50 5 2.21 2.21 2.21 2.21 2.21 2.21 2.21 2.21 2.21 2.21 2.21 2.21 2.21 2.21 50 10
3.00 3.01 3.02 3.02 3.03 3.03 3.03 3.03 3.03 3.04 3.04 3.04 3.04 3.04 50 20 3.96
4.00 4.03 4.06 4.08 4.10 4.11 4.12 4.13 4.14 4.14 4.15 4.15 4.16 50 50 5.20 5.33
5.45 5.55 5.64 5.72 5.79 5.85 5.90 5.95 5.99 6.03 6.06 6.08 50 100 5.92 6.13 6.33
6.51 6.67 6.83 6.97 7.10 7.21 7.32 7.42 7.51 7.59 7.67 50 200 6.39 6.66 6.93 7.18
7.41 7.64 7.85 8.05 8.24 8.42 8.59 8.75 8.90 9.04 50 500 6.71 7.04 7.36 7.67 7.97
8.26 8.54 8.81 9.07 9.32 9.57 9.80 10.03 10.25 50 1000 6.83 7.18 7.52 7.85 8.18
8.50 8.80 9.10 9.40 9.68 9.96 10.23 10.50 10.75 50 10000 6.94 7.31 7.68 8.03 8.38
8.72 9.06 9.39 9.72 10.04 10.35 10.66 10.97 11.27

100 5 2.25 2.25 2.25 2.25 2.25 2.25 2.25 2.25 2.25 2.25 2.25 2.25 2.25 2.25 100 10
3.08 3.09 3.10 3.10 3.11 3.11 3.11 3.11 3.11 3.12 3.12 3.12 3.12 3.12 100 20 4.10
4.14 4.18 4.21 4.23 4.25 4.27 4.28 4.29 4.30 4.30 4.31 4.31 4.31 100 50 5.46 5.61
5.74 5.85 5.95 6.04 6.12 6.19 6.24 6.30 6.34 6.38 6.42 6.45 100 100 6.26 6.50 6.72
6.93 7.12 7.29 7.45 7.60 7.74 7.86 7.98 8.08 8.18 8.27 100 200 6.79 7.10 7.41 7.69
7.97 8.23 8.48 8.72 8.94 9.15 9.35 9.54 9.72 9.89 100 500 7.16 7.54 7.91 8.27 8.62
8.96 9.29 9.62 9.93 10.24 10.53 10.82 11.10 11.37 100 1000 7.30 7.70 8.10 8.49 8.87
9.24 9.61 9.97 10.33 10.67 11.01 11.35 11.68 12.00 100 10000 7.43 7.85 8.28 8.69
9.11 9.51 9.92 10.32 10.72 11.11 11.50 11.88 12.27 12.64

1000 1000 1000 1000 1000 1000 1000 1000 1000

5 10 20 50 100 200 500 1000 10000

2.28 3.15 4.24 5.72 6.61 7.20 7.62 7.78 7.92

2.29 3.16 4.29 5.88 6.87 7.56 8.05 8.24 8.41

2.29 3.17 4.32 6.02 7.12 7.90 8.48 8.69 8.90

2.29 3.18 4.36 6.15 7.35 8.23 8.89 9.15 9.39

2.29 3.18 4.38 6.26 7.57 8.55 9.30 9.59 9.87

2.29 3.18 4.40 6.36 7.77 8.85 9.70 10.04 10.36

2.29 3.19 4.42 6.44 7.95 9.14 10.10 10.47 10.84

2.29 3.19 4.43 6.52 8.12 9.41 10.48 10.91 11.33

2.29 3.19 4.44 6.59 8.28 9.68 10.86 11.33 11.81

2.29 3.19 4.45 6.65 8.43 9.93 11.22 11.75 12.29

2.29 3.19 4.46 6.70 8.56 10.17 11.58 12.17 12.77

2.29 3.19 4.46 6.74 8.68 10.39 11.93 12.58 13.25

2.29 3.19 4.47 6.78 8.79 10.61 12.27 12.99 13.73

2.29 3.19 4.47 6.81 8.89 10.81 12.61 13.39 14.20

10000 10000 10000 10000 10000 10000 10000 10000 10000

5 10 20 50 100 200 500 1000 10000

2.29 3.16 4.25 5.75 6.64 7.24 7.67 7.83 7.98

2.29 3.17 4.30 5.91 6.91 7.60 8.11 8.30 8.47

2.29 3.18 4.34 6.05 7.17 7.95 8.54 8.76 8.97

2.29 3.19 4.37 6.18 7.40 8.29 8.96 9.22 9.46

2.29 3.19 4.40 6.29 7.62 8.61 9.38 9.67 9.96

2.29 3.19 4.42 6.39 7.82 8.92 9.78 10.12 10.45


2.29 3.19 4.43 6.48 8.01 9.21 10.18 10.57 10.94

2.29 3.20 4.45 6.56 8.18 9.49 10.57 11.01 11.44

2.29 3.20 4.46 6.62 8.34 9.76 10.96 11.44 11.93

2.29 3.20 4.46 6.68 8.49 10.01 11.33 11.88 12.42

2.29 3.20 4.47 6.73 8.62 10.26 11.70 12.30 12.91

2.29 3.20 4.48 6.78 8.74 10.49 12.06 12.72 13.40

2.29 3.20 4.48 6.82 8.86 10.71 12.41 13.14 13.89

2.29 3.20 4.48 6.85 8.96 10.91 12.75 13.55 14.38

5-10

National Frame Builders Association

Post-Frame Building Design Manual

Table 5.2. Sidesway Restraining Force Modifier (mD)

ke / k

Ch / k

Number of frames (endwalls counted as frames)

3 4 5 6 7 8 9 10 11 12 13 14 15 16

5 5 0.75 0.64 0.52 0.43 0.34 0.28 0.22 0.18 0.14 0.12 0.09 0.08 0.06 0.05 5 10 0.78
0.69 0.59 0.52 0.44 0.39 0.33 0.28 0.24 0.21 0.18 0.15 0.13 0.11 5 20 0.80 0.72
0.64 0.58 0.51 0.46 0.41 0.37 0.33 0.30 0.26 0.24 0.21 0.19 5 50 0.81 0.74 0.67
0.62 0.56 0.52 0.48 0.44 0.41 0.38 0.35 0.32 0.30 0.28 5 100 0.81 0.74 0.68 0.63
0.58 0.54 0.50 0.47 0.44 0.41 0.38 0.36 0.34 0.32 5 200 0.82 0.75 0.69 0.64 0.59
0.55 0.52 0.48 0.46 0.43 0.41 0.38 0.36 0.35 5 500 0.82 0.75 0.69 0.64 0.60 0.56
0.52 0.49 0.47 0.44 0.42 0.40 0.38 0.36 5 1000 0.82 0.75 0.69 0.64 0.60 0.56 0.53
0.50 0.47 0.45 0.42 0.40 0.39 0.37 5 10000 0.82 0.75 0.69 0.64 0.60 0.56 0.53 0.50
0.47 0.45 0.43 0.41 0.39 0.37

10 5 0.83 0.73 0.60 0.51 0.41 0.34 0.27 0.22 0.17 0.14 0.11 0.09 0.07 0.06 10 10
0.86 0.79 0.70 0.63 0.54 0.48 0.41 0.36 0.30 0.26 0.22 0.19 0.16 0.14 10 20 0.88
0.83 0.76 0.70 0.64 0.58 0.52 0.48 0.43 0.39 0.35 0.31 0.28 0.25 10 50 0.90 0.85
0.80 0.75 0.71 0.66 0.62 0.58 0.55 0.51 0.48 0.45 0.42 0.39 10 100 0.90 0.86 0.81
0.77 0.73 0.70 0.66 0.63 0.60 0.57 0.54 0.51 0.49 0.46 10 200 0.90 0.86 0.82 0.78
0.75 0.71 0.68 0.65 0.63 0.60 0.57 0.55 0.53 0.51 10 500 0.90 0.86 0.82 0.79 0.75
0.72 0.70 0.67 0.64 0.62 0.60 0.58 0.56 0.54 10 1000 0.90 0.86 0.83 0.79 0.76 0.73
0.70 0.67 0.65 0.63 0.61 0.59 0.57 0.55 10 10000 0.91 0.86 0.83 0.79 0.76 0.73 0.70
0.68 0.66 0.63 0.61 0.59 0.58 0.56

20 5 0.87 0.78 0.65 0.56 0.45 0.38 0.30 0.25 0.19 0.16 0.13 0.10 0.08 0.07 20 10
0.91 0.85 0.76 0.69 0.60 0.54 0.46 0.41 0.35 0.30 0.26 0.22 0.19 0.16 20 20 0.93
0.89 0.83 0.78 0.72 0.66 0.60 0.55 0.50 0.46 0.41 0.37 0.33 0.30 20 50 0.94 0.91
0.87 0.84 0.80 0.76 0.72 0.69 0.65 0.62 0.58 0.55 0.51 0.48 20 100 0.95 0.92 0.89
0.86 0.83 0.80 0.77 0.75 0.72 0.69 0.66 0.64 0.61 0.58 20 200 0.95 0.92 0.90 0.87
0.85 0.83 0.80 0.78 0.76 0.73 0.71 0.69 0.67 0.65 20 500 0.95 0.93 0.90 0.88 0.86
0.84 0.82 0.80 0.78 0.76 0.74 0.72 0.71 0.69 20 1000 0.95 0.93 0.91 0.88 0.86 0.84
0.82 0.81 0.79 0.77 0.75 0.74 0.72 0.71 20 10000 0.95 0.93 0.91 0.89 0.87 0.85 0.83
0.81 0.80 0.78 0.76 0.75 0.73 0.72

50 5 0.89 0.81 0.68 0.59 0.48 0.40 0.32 0.26 0.21 0.17 0.13 0.11 0.09 0.07 50 10
0.93 0.88 0.80 0.73 0.65 0.58 0.50 0.44 0.38 0.33 0.28 0.24 0.21 0.18 50 20 0.96
0.93 0.88 0.83 0.77 0.72 0.66 0.61 0.55 0.51 0.46 0.41 0.37 0.34 50 50 0.97 0.95
0.93 0.90 0.87 0.84 0.80 0.77 0.73 0.70 0.66 0.63 0.59 0.56 50 100 0.98 0.96 0.94
0.93 0.90 0.88 0.86 0.84 0.81 0.79 0.76 0.74 0.71 0.69 50 200 0.98 0.97 0.95 0.94
0.92 0.91 0.89 0.88 0.86 0.84 0.82 0.81 0.79 0.77 50 500 0.98 0.97 0.96 0.95 0.94
0.92 0.91 0.90 0.89 0.88 0.86 0.85 0.84 0.83 50 1000 0.98 0.97 0.96 0.95 0.94 0.93
0.92 0.91 0.90 0.89 0.88 0.87 0.86 0.85 50 10000 0.98 0.97 0.96 0.95 0.94 0.93 0.93
0.92 0.91 0.90 0.89 0.88 0.87 0.87

100 5 0.90 0.82 0.69 0.60 0.48 0.41 0.32 0.27 0.21 0.17 0.14 0.11 0.09 0.07 100 10
0.94 0.90 0.82 0.75 0.66 0.59 0.51 0.45 0.39 0.34 0.29 0.25 0.21 0.18 100 20 0.97
0.94 0.89 0.85 0.79 0.74 0.68 0.63 0.57 0.52 0.47 0.43 0.39 0.35 100 50 0.98 0.97
0.94 0.92 0.89 0.86 0.83 0.80 0.76 0.73 0.69 0.66 0.62 0.59 100 100 0.99 0.98 0.96
0.95 0.93 0.91 0.89 0.87 0.85 0.83 0.80 0.78 0.75 0.73 100 200 0.99 0.98 0.97 0.96
0.95 0.94 0.93 0.91 0.90 0.88 0.87 0.85 0.84 0.82 100 500 0.99 0.98 0.98 0.97 0.96
0.96 0.95 0.94 0.93 0.92 0.91 0.90 0.89 0.88 100 1000 0.99 0.98 0.98 0.97 0.97 0.96
0.95 0.95 0.94 0.93 0.93 0.92 0.91 0.91 100 10000 0.99 0.99 0.98 0.98 0.97 0.97
0.96 0.96 0.95 0.95 0.94 0.94 0.93 0.93

1000 1000 1000 1000 1000 1000 1000 1000 1000

5 10 20 50 100 200 500 1000 10000

0.91 0.95 0.98 0.99 0.99 1.00 1.00 1.00 1.00

0.83 0.91 0.95 0.98 0.99 0.99 1.00 1.00 1.00

0.70 0.83 0.91 0.96 0.98 0.99 0.99 1.00 1.00

0.61 0.76 0.87 0.94 0.97 0.98 0.99 1.00 1.00

0.49 0.67 0.81 0.91 0.95 0.98 0.99 0.99 1.00

0.41 0.60 0.76 0.89 0.94 0.97 0.99 0.99 1.00

0.33 0.52 0.70 0.86 0.92 0.96 0.98 0.99 1.00

0.27 0.46 0.65 0.83 0.90 0.95 0.98 0.99 1.00

0.22 0.40 0.59 0.79 0.88 0.94 0.97 0.98 0.99

0.18 0.35 0.54 0.76 0.86 0.93 0.97 0.98 0.99

0.14 0.30 0.49 0.72 0.84 0.91 0.96 0.98 0.99

0.11 0.26 0.45 0.69 0.82 0.90 0.95 0.97 0.99

0.09 0.22 0.40 0.65 0.79 0.88 0.95 0.97 0.99

0.07 0.19 0.36 0.62 0.77 0.87 0.94 0.97 0.99

10000 10000 10000 10000 10000 10000 10000 10000 10000

5 10 20 50 100 200 500 1000 10000


0.91 0.83 0.70 0.61 0.49 0.42 0.33 0.27 0.22 0.18 0.14 0.11 0.09 0.95 0.91 0.83
0.76 0.68 0.61 0.53 0.46 0.40 0.35 0.30 0.26 0.22 0.98 0.95 0.91 0.87 0.81 0.76
0.70 0.65 0.59 0.54 0.49 0.45 0.40 0.99 0.98 0.96 0.94 0.92 0.89 0.86 0.83 0.79
0.76 0.72 0.69 0.65 1.00 0.99 0.98 0.97 0.96 0.94 0.93 0.91 0.89 0.87 0.84 0.82
0.80 1.00 1.00 0.99 0.99 0.98 0.97 0.96 0.95 0.94 0.93 0.92 0.90 0.89 1.00 1.00
1.00 0.99 0.99 0.99 0.98 0.98 0.98 0.97 0.96 0.96 0.95 1.00 1.00 1.00 1.00 1.00
0.99 0.99 0.99 0.99 0.99 0.98 0.98 0.98 1.00 1.00 1.00 1.00 1.00 1.00 1.00 1.00
1.00 1.00 1.00 1.00 1.00

0.07 0.19 0.37 0.62 0.77 0.87 0.95 0.97 1.00

5-11

National Frame Builders Association

Post-Frame Building Design Manual

Table 5.2. Sidesway Restraining Force Modifier (mD), cont.

ke / k

Ch / k

Number of frames (endwalls counted as frames)

17 18 19 20 21 22 23 24 25 26 27 28 29 30

5 5 0.04 0.03 0.02 0.02 0.02 0.01 0.01 0.01 0.01 0.01 0.00 0.00 0.00 0.00 5 10 0.09
0.08 0.07 0.06 0.05 0.04 0.04 0.03 0.03 0.02 0.02 0.02 0.01 0.01 5 20 0.17 0.15
0.13 0.12 0.11 0.10 0.09 0.08 0.07 0.06 0.06 0.05 0.04 0.04 5 50 0.26 0.24 0.22
0.21 0.19 0.18 0.17 0.16 0.14 0.13 0.12 0.12 0.11 0.10 5 100 0.30 0.29 0.27 0.26
0.24 0.23 0.22 0.20 0.19 0.18 0.17 0.17 0.16 0.15 5 200 0.33 0.31 0.30 0.29 0.27
0.26 0.25 0.24 0.23 0.22 0.21 0.20 0.19 0.19 5 500 0.35 0.33 0.32 0.31 0.29 0.28
0.27 0.26 0.25 0.25 0.24 0.23 0.22 0.21 5 1000 0.35 0.34 0.33 0.31 0.30 0.29 0.28
0.27 0.26 0.25 0.25 0.24 0.23 0.23 5 10000 0.36 0.35 0.33 0.32 0.31 0.30 0.29 0.28
0.27 0.26 0.26 0.25 0.24 0.24

10 5 0.05 0.04 0.03 0.02 0.02 0.02 0.01 0.01 0.01 0.01 0.01 0.00 0.00 0.00 10 10
0.12 0.10 0.09 0.08 0.06 0.06 0.05 0.04 0.03 0.03 0.03 0.02 0.02 0.02 10 20 0.23
0.20 0.18 0.16 0.15 0.13 0.12 0.11 0.09 0.08 0.08 0.07 0.06 0.05 10 50 0.36 0.34
0.32 0.30 0.28 0.26 0.24 0.23 0.21 0.20 0.18 0.17 0.16 0.15 10 100 0.44 0.42 0.40
0.38 0.36 0.34 0.33 0.31 0.29 0.28 0.27 0.25 0.24 0.23 10 200 0.49 0.47 0.45 0.43
0.42 0.40 0.39 0.37 0.36 0.34 0.33 0.32 0.31 0.30 10 500 0.52 0.50 0.49 0.47 0.46
0.44 0.43 0.42 0.40 0.39 0.38 0.37 0.36 0.35 10 1000 0.53 0.52 0.50 0.49 0.47 0.46
0.45 0.43 0.42 0.41 0.40 0.39 0.38 0.37 10 10000 0.54 0.53 0.51 0.50 0.49 0.47 0.46
0.45 0.44 0.43 0.42 0.41 0.40 0.39

20 5 0.05 0.04 0.03 0.03 0.02 0.02 0.01 0.01 0.01 0.01 0.01 0.01 0.00 0.00 20 10
0.14 0.12 0.10 0.09 0.07 0.06 0.05 0.05 0.04 0.03 0.03 0.03 0.02 0.02 20 20 0.27
0.24 0.22 0.20 0.17 0.16 0.14 0.13 0.11 0.10 0.09 0.08 0.07 0.06 20 50 0.45 0.42
0.40 0.37 0.35 0.33 0.30 0.28 0.27 0.25 0.23 0.22 0.20 0.19 20 100 0.56 0.53 0.51
0.49 0.47 0.45 0.43 0.41 0.39 0.37 0.35 0.34 0.32 0.31 20 200 0.63 0.61 0.59 0.57
0.55 0.53 0.52 0.50 0.48 0.47 0.45 0.44 0.42 0.41 20 500 0.67 0.66 0.64 0.63 0.61
0.60 0.59 0.57 0.56 0.55 0.53 0.52 0.51 0.50 20 1000 0.69 0.68 0.66 0.65 0.64 0.62
0.61 0.60 0.59 0.58 0.57 0.55 0.54 0.53 20 10000 0.71 0.69 0.68 0.67 0.66 0.65 0.64
0.63 0.62 0.61 0.60 0.59 0.58 0.57

50 5 0.06 0.05 0.04 0.03 0.02 0.02 0.02 0.01 0.01 0.01 0.01 0.01 0.00 0.00 50 10
0.15 0.13 0.11 0.10 0.08 0.07 0.06 0.05 0.04 0.04 0.03 0.03 0.02 0.02 50 20 0.30
0.27 0.24 0.22 0.20 0.18 0.16 0.14 0.13 0.11 0.10 0.09 0.08 0.07 50 50 0.52 0.49
0.46 0.44 0.41 0.38 0.36 0.34 0.31 0.29 0.27 0.26 0.24 0.22 50 100 0.66 0.64 0.61
0.59 0.56 0.54 0.52 0.50 0.47 0.45 0.43 0.41 0.40 0.38 50 200 0.75 0.73 0.71 0.69
0.68 0.66 0.64 0.62 0.60 0.59 0.57 0.55 0.54 0.52 50 500 0.81 0.80 0.79 0.78 0.76
0.75 0.74 0.73 0.71 0.70 0.69 0.68 0.67 0.65 50 1000 0.84 0.83 0.82 0.81 0.80 0.79
0.78 0.77 0.76 0.75 0.74 0.73 0.72 0.71 50 10000 0.86 0.85 0.84 0.84 0.83 0.82 0.81
0.81 0.80 0.79 0.79 0.78 0.77 0.77

100 5 0.06 0.05 0.04 0.03 0.02 0.02 0.02 0.01 0.01 0.01 0.01 0.01 0.00 0.00 100 10
0.16 0.13 0.11 0.10 0.08 0.07 0.06 0.05 0.04 0.04 0.03 0.03 0.02 0.02 100 20 0.31
0.28 0.25 0.23 0.20 0.18 0.16 0.15 0.13 0.12 0.11 0.09 0.08 0.08 100 50 0.55 0.52
0.49 0.46 0.43 0.41 0.38 0.36 0.33 0.31 0.29 0.27 0.25 0.24 100 100 0.70 0.68 0.65
0.63 0.60 0.58 0.56 0.53 0.51 0.49 0.47 0.45 0.43 0.41 100 200 0.80 0.78 0.77 0.75
0.73 0.71 0.69 0.68 0.66 0.64 0.62 0.61 0.59 0.57 100 500 0.87 0.86 0.85 0.84 0.83
0.82 0.81 0.80 0.79 0.77 0.76 0.75 0.74 0.73 100 1000 0.90 0.89 0.88 0.88 0.87 0.86
0.85 0.84 0.84 0.83 0.82 0.81 0.80 0.80 100 10000 0.92 0.92 0.91 0.91 0.90 0.90
0.90 0.89 0.89 0.88 0.88 0.87 0.87 0.86

1000 1000 1000 1000 1000 1000 1000 1000 1000

5 10 20 50 100 200 500 1000 10000

0.06 0.16 0.33 0.58 0.74 0.85 0.93 0.96 0.99

0.05 0.14 0.29 0.55 0.72 0.84 0.93 0.96 0.99

0.04 0.12 0.26 0.52 0.69 0.82 0.92 0.95 0.99

0.03 0.10 0.24 0.49 0.67 0.80 0.91 0.95 0.99

0.02 0.09 0.21 0.46 0.64 0.79 0.90 0.94 0.99

0.02 0.07 0.19 0.43 0.62 0.77 0.89 0.94 0.98

0.02 0.06 0.17 0.40 0.60 0.75 0.88 0.93 0.98

0.01 0.05 0.15 0.38 0.57 0.74 0.87 0.93 0.98

0.01 0.05 0.14 0.35 0.55 0.72 0.86 0.92 0.98

0.01 0.04 0.12 0.33 0.53 0.70 0.85 0.92 0.98

0.01 0.03 0.11 0.31 0.50 0.68 0.84 0.91 0.98

0.01 0.03 0.10 0.29 0.48 0.66 0.83 0.90 0.98

0.00 0.02 0.09 0.27 0.46 0.65 0.82 0.90 0.98

0.00 0.02 0.08 0.25 0.44 0.63 0.81 0.89 0.98

10000 10000 10000 10000 10000 10000 10000 10000 10000

5 10 20 50 100 200 500 1000 10000

0.06 0.05 0.04 0.03 0.02 0.02 0.02 0.01 0.01 0.01 0.01 0.01 0.00 0.16 0.14 0.12
0.10 0.09 0.07 0.06 0.05 0.05 0.04 0.03 0.03 0.02 0.33 0.30 0.26 0.24 0.21 0.19
0.17 0.15 0.14 0.12 0.11 0.10 0.09 0.58 0.55 0.52 0.49 0.46 0.43 0.40 0.38 0.36
0.33 0.31 0.29 0.27 0.75 0.72 0.70 0.67 0.65 0.62 0.60 0.58 0.55 0.53 0.51 0.49
0.47 0.86 0.84 0.83 0.81 0.79 0.78 0.76 0.74 0.72 0.71 0.69 0.67 0.65 0.94 0.93
0.92 0.92 0.91 0.90 0.89 0.88 0.87 0.86 0.85 0.84 0.83 0.97 0.96 0.96 0.96 0.95
0.95 0.94 0.94 0.93 0.93 0.92 0.91 0.91 1.00 1.00 1.00 1.00 0.99 0.99 0.99 0.99
0.99 0.99 0.99 0.99 0.99

0.00 0.02 0.08 0.25 0.45 0.64 0.82 0.90 0.99

5-12

National Frame Builders Association

Post-Frame Building Design Manual

FRAME FRAME

APPLIED HORIZONTAL LOAD RESISTED FRACTION OF

NUMBER STIFFNESS LOAD DISPLACEMENT BY FRAME APPLIED LOAD

---------------------------------------------------------------------

1 10000.00 500.0 .1946696

1946.7

3.8934

2 100.00 1000.0 .3393392

33.9

.0339

3 100.00 1000.0 .3874022

38.7

.0387

4 100.00 1000.0 .3393392

33.9

.0339

5 10000.00 500.0 .1946696

1946.7

3.8934

DIAPHRAGM DIAPHRAGM

SHEAR

SHEAR

NUMBER
STIFFNESS DISPLACEMENT LOAD

--------------------------------------------

10000.00

.1446696 1446.7

10000.00

.0480630

480.6

10000.00

.0480630

480.6

10000.00

.1446696 1446.7

Figure 5.9. Sample output from computer program DAFI.

When requirements for use of tables 5.1 and 5.2 are met, equation 5-8 can be used
to calculate the total sidesway resisting force, Q. In all other cases, analysis
tools such as DAFI must be used to obtain Q. A copy of output from program DAFI for
a 4-bay building with Ch fixed at 10000, ke at 10000, k at 100, and R at 1000 is
shown in figure 5.9. Although the sidesway resisting force for each frame is not
given in the DAFI output, it is numerically equal to the difference between the
load applied to the frame, and the load resisted by the frame two values that are
listed in the programs output. For example, Q for the critical middle frame (frame
3 in figure 5.9), would be equal to the difference between 1000.0 and 38.7 or
961.3.
Since diaphragm construction typically doesnt change from one side of a frame to
the other side of the frame, Ch and ch,i values associated with either of the two
diaphragm elements (that are adjacent to the frame) can be used in equation 5-9.
Horizontal restraining forces calculated for the three diaphragms in figure 5.2b,
are graphically illustrated in figure 5.10a. For post-frame component stress
analysis, these restraining forces should be applied as in-plane forces as shown in
figure 10b. In-plane forces are calculated from the horizontal forces as follows:

Qp,i = Q i / (cos i)

or

q p,i = Q i / (d i cos i)

(5-11)
where:

Q pi = Qi = i = q p,i =
di =

in-plane force applied to frame by diaphragm i, lbf (N) sidesway resisting force
due to diaphragm i, lbf (N)
slope of diaphragm i in-plane force applied to the frame per unit length of
diaphragm i, lbf/ft (N/m) slope length of diaphragm i, ft (m)

5.6.6 Simple Beam Analogy Equations. McGuire (1998) presented the concept of
modeling the diaphragm as a simple beam with an applied load inversely proportional
to deflection. This analogy resulted in the following equations for calculating
diaphragm shear forces and lateral displacements for the special case when: (1) all
diaphragm elements have the same stiffness Ch, (2) all interior frame elements have
the same stiffness, k, (3) both exterior frame elements (i.e., the two elements
representing the endwalls) have the same stiffness, ke, and (4) eave load, R, is
the same at each interior frame.

5-13

National Frame Builders Association

Post-Frame Building Design Manual

Roof Gravity Loads

s x qwr

s x qlr

Qa Qc

Qb

Ceiling Gravity Loads

s xqlw

s xqww

(a) Roof Gravity Loads

s x qwr

s x qlr

q p,a

q p,c

q p,b

Ceiling Gravity Loads

s xqww

(b)
Figure 5.10 (a) Frame with diaphragm resisting forces. (b) Resisting forces applied
as uniformly distributed in-plane loads for frame component stress analysis.

Vh = Chs[A sinh( x) + B cosh( x)] (5-12)

y = A cosh( x) + B sinh( x) + R/k (5-13)

ye = R / [ k (1 D)]

(5-14)

where:

Vh = diaphragm shear force, lbf (N) x = distance from endwall, in. (mm)
R = eave load, lbf (N)

s xqlw

s= y=
ye = k= ke = Ch = L= sinh = cosh =

frame spacing, in. (mm) lateral displacement of diaphragm at a distance x from the
endwall, in. (mm) lateral displacement of the endwall, in. (mm) stiffness of
interior frames, lbf/in. (N/mm) stiffness of endwall frames (or shearwalls),
lbf/in. (N/mm) horizontal shear stiffness for a width s of the diaphragm, lbf/in.
(N/mm) Distance between endwalls, in. (mm) hyperbolic sine hyperbolic cosine

( k / Ch )1/2 s

A = ye R/k

B=

A ( 1 cosh( L)) sinh( L)

D=

ke sinh( L) Ch s (1 - cosh( L))

5.7 Component Design

5.7.1 General. All building components must be checked to ensure that actual loads
do not exceed allowable design values. In this section, special attention is given
to components that are involved in load transfer by diaphragm action.

5.7.2 Diaphragms. The maximum shear in a


diaphragm section, Vp,i, cannot exceed the allowable shear strength of the section,
va,i, multiplied by the diaphragm length.

Vp,i < va,i d i where:

(5-15)

Vp,i = va,i =
di =

in-plane shear force in diaphragm section i from equation 5-9 lbf (N) allowable in-
plane shear strength of diaphragm i (see Section 6.3.3), lbf/ft (N/m) slope length
of diaphragm i, ft (m)

5-14

National Frame Builders Association

Post-Frame Building Design Manual

5.7.3 Diaphragm Chords. In addition to shear forces, a roof/ceiling diaphragm


assembly must also resist bending moment. The magnitude of this bending moment is
dependent on a number of factors. For design, this bending moment is assumed to be
no greater than:

Md = Vh L / 4

(5-16)

where:

Md = Vh = L=

diaphragm bending moment, lbf-ft (N m) maximum total shear in roof/ceiling


diaphragm assembly, lbf (N) distance between shearwalls, ft (m)

Equation 5-16 treats the roof/ceiling assembly as a uniformly loaded beam that is
simple supported by two shearwalls spaced a distance L apart. Each shearwall is
assumed to be subjected to a force that is equal to the maximum total shear in the
roof/ceiling assembly, Vh. The maximum total shear in the roof/ceiling assembly,
Vh, can be obtained from computer output (e.g. figure 5.9), or equation 5-7 or 5-12
if applicable. The uniform load on the roof/ceiling assembly (w in figure 5.11a) is
set equal to 2Vh/L. This quantity is multiplied by L2/8 to obtain Md.

The bending moment applied to a roof/ceiling diaphragm assembly is resisted by


axial forces (a.k.a. chord forces) in members orientated perpendicular to
trusses/rafters. This includes roof purlins and analogous framing members in the
ceiling diaphragm. For bending moment calculations, these members are referred to
as diaphragm chords (figure 5.11a). Any connection in the chords, either between
intermediate chord members or where they are connected to the endwalls, must be
designed to resist the calculated axial force.

If the roof/ceiling assembly behaves as a single beam in resisting bending moment,


the maximum chord force (which is located in the edge chords) can be calculated as:

Pe = Md / b where:

(5-17)

Pe = Md =
=
b=

axial force in edge chord, lbf (N) diaphragm bending moment from equation 5-16,
lbf-ft (N m)
reduction factor dependent on chord force distribution horizontal distance between
edge chords, ft (m)

Vh Vh

Shearwall

Chords (a)

Trusses/rafters

(b)

(c) (d)

Figure 5.11. (a) Plan view of a diaphragm under a uniform load, w. Chord force
distributions when (b) moment resisted by edge chords only, (b) chord force
distribution is linear, and (c) chord force distribution is linear, but diaphragm
halves assumed to act independently in resisting moment.

5-15

National Frame Builders Association

Post-Frame Building Design Manual

The axial force in an edge chord is dependent on chord force distribution as


indicated by the presence of in equation 5-17. The current ASAE EP484 diaphragm
design procedure (ASAE, 1999a) assumes that edge chords act alone in resisting
bending moment (figure 5.11b). For this case, is numerically equal to one (1).
This is a conservative approach. Alternatively, many engineers assume a linear
distribution of chord forces as shown in figure 5.11c. When a linear distribution
is assumed, the reduction factor is a function of chord location. If chords are
evenly spaced, then is given as:

(n 1 )2 = n/2
(n 2 i + 1)2
i =1

when n is even

(n 1 )2 = (n-1)/2
(n 2 i + 1)2
i =1

when n is odd

where:

= reduction factor when chords are evenly spaced and chord forces are linearly
distributed
n = number of chord rows, including the two rows of edge chords

The preceding equations were used to calculate the values given in table 5.3.

If a linear distribution of chord force is assumed (figure 5.11c), and interior


chords are evenly spaced, the load in an interior chord, Pi, is given as:

Pi = 2 Pe x i / b where:
(5-18)

Pi = Pe =
b=
xi =

axial force for chord in row i, lbf (N) axial force in edge chord from equation 5-
17, lbf (N) horizontal distance between edge chords, ft (m) horizontal distance
from center of diaphragm to chord row i.

Table 5.3. Reduction Factor, , for Axial

Force in Edge Chords

n*

n*

2 1.000

22 0.249

3 1.000

23 0.239

4 0.900

24 0.230

5 0.800

25 0.222

6 0.714

26 0.214

7 0.643

27 0.206

8 0.583

28 0.200

9 0.533

29 0.193

10 0.491

30 0.187

11 0.455

31 0.181

12 0.423
32 0.176

13 0.396

33 0.171

14 0.371

34 0.166

15 0.350

35 0.162

16 0.335

36 0.158

17 0.314

37 0.154

18 0.298

38 0.150

19 0.284

39 0.146

20 0.271

40 0.143

21 0.260

41 0.139

* n is the number of chord rows, including the

two rows of edge chords

Technical Note
Chord Forces
The axial force induced in an individual chord by applied building loads is a
function of many complex, interacting design variables. For this reason, designers
have had to rely on simplifying assumptions in order to approximate chord forces.
One common assumption is that the roof/ceiling assembly acts as a large deep beam
that is simply supported by two end shearwalls. This assumption is used to
calculate the maximum inplane bending moment to which a diaphragm is subjected.
This assumption is conservative in that it neglects the resistance to in-plane
bending contributed by sidewalls. Sidewalls help resist (and thereby reduce) in-
plane bending moments in two ways. First they brace endwalls

5-16

National Frame Builders Association


Post-Frame Building Design Manual

and other shearwalls, which limits rotation of the diaphragm at these shearwalls.
Second, they resist a change in eave length (and hence changes in eave chord
forces) by virtue of their own in-plane shear stiffness.
Because of the influence of sidewalls, the distribution of in-plane bending moment
will not follow that for a typical simple supported beam (i.e., zero moment at the
supports, and maximum moment at midspan). For this reason, Pollock and others
(1996) recommend modeling the roof/ceiling assembly as a deep beam with fixed
supports.
Because of uncertainty surrounding variation in in-plane bending moment with
building length, some designers will assign the maximum calculated in-plane bending
moment (Md from equation 5-16) to every location along the length of the building.
This is obviously a conservative approach.
Another major assumption that a designer must make involves the distribution of
chord forces across a building. Three different chord force distributions are shown
in figure 5.11b, 5.11c, and 5.11d. Whether or not edge chords resist virtually all
of the in-plane bending moment (figure 5.11b), or a linear distribution of axial
forces exists in chords between edge chords (figure 5.11c) is a question that is at
the heart of ongoing research. In reality, the distribution of chord forces lies
somewhere in between these two extremes, exactly where being dependent on specifics
of the design and on the magnitude of the applied load (Note: at higher load
levels, load distributions change due to geometric and material nonlinearities).
Presently, there is very little research data to support one specific design
procedure/assumption. The most extensive investigation of chord forces was by Niu
and Gebremedhin (1997) who strain gauged purlins in a full-scale building and in a
diaphragm test assembly. The data collected in this study does not strongly support
any particular hypotheses regarding chord force distribution. The only other
research of significance to chord force distribution was the comprehensive finite
element analyses of diaphragm assemblies by Wright (1992) and Williams (1999). Both
of these researchers found that in-plane bending

moment in their models was resisted almost entirely by the edge purlins. Bohnhoff
and others (1999) showed that as the shear stiffness of cladding is increased,
interior purlins get more involved in resisting in-plane bending moments.
Chord force distribution has also been shown to depend on the degree of interaction
between individual diaphragms. Figure 5.11d illustrates the distribution of chord
forces when there is no interaction between individual diaphragms on both sides of
a ridge. Note that interaction between individual diaphragms on opposites sides of
a ridge is highly dependent on: (1) the spacing between ridge purlins, and (2) the
rigidity of the ridge cap and other elements joining the two diaphragms.

5.7.4 Shearwalls. End and intermediate shearwalls must have sufficient strength to
transmit forces from roof and ceiling diaphragms to the foundation system. In
equation form:

va > Vs / (W DT)

(5-19)

where:

va =
Vs = W= DT =

allowable shear capacity of shearwall, lbf/ft (N/m) force induced in shearwall, lbf
(N) building width, ft (m) total width of door and window openings in the
shearwall, ft (m)
The allowable shear capacity of end and intermediate shearwalls, va, is obtained
from validated structural models, or from tests as outlined in ASAE EP558 (see
Section 6.5). The total force in the shear wall, Vs, is obtained from computer
output (e.g. figure 5.8), or equation 57 or equation 5-12 if applicable.

The total width of door and window openings, DT, generally varies with height as
shown in figure 5.12. At locations where DT is the greatest (section b-b in figure
5.12) additional reinforcing may be required to ensure that the allowable shear
stress is not exceeded.

The structural framing over a door or window opening will act as a drag strut
transferring

5-17

National Frame Builders Association

Post-Frame Building Design Manual

shear across the opening. The header over the opening shall be designed to carry
the force in tension and/or compression across the opening.
aa
bb
cWc
Figure 5.12. Shearwall showing variations in opening width, DT, with height.
Shearwall strength can easily be increased when the applied load exceeds shearwall
capacity. For example, the density of stitch screws can be increased and additional
fasteners can be added in panel flats (on both sides of each major rib is the most
effective). If only one side of the wall has been sheathed, add wood paneling or
metal cladding to the other side. Metal diagonal braces can also be added beneath
any wood paneling or corrugated metal siding.
5.7.5 Shearwall Connections. Connections that fasten (1) roof and ceiling
diaphragms to a shearwall, and (2) shearwalls to the foundation system, must be
designed to carry the appropriate amount of shear load. The design of these
connections may be proved by tests of a typical connection detail or by an
appropriate calculation method.
At end shearwalls it is not uncommon to use the truss top chord to transfer load
from roof cladding to endwall cladding. Sidewall steel is fastened directly to the
truss chord, as is the roof steel when purlins are inset. In buildings with top-
running purlins, roof cladding can not be fastened directly to the truss. In such
cases, blocking equal in depth to the purlins is placed between the purlins and
fastened to the truss. Roof cladding is then attached directly to this blocking.

5.7.6 Shearwall Overturning. Diaphragm loading produces overturning moment in


shearwalls. This moment induces vertical forces in shearwall-to-foundation
connections that must be added to vertical forces resulting from tributary loads.
In the case of embedded posts, increases in uplift forces may require an increase
in embedment depth, and increases in downward force may require an increase in
footing size (see Chapter 8).

5.8 Rigid Roof Design

5.8.1 General. When diaphragm stiffness is considerably greater than the stiffness
of interior post frames, the designer may want to assume that the diaphragm and
shearwalls are infinitely stiff. Under this assumption, 100% of the applied eave
load, R, is transferred by the diaphragm to shearwalls, and none of the applied
eave load is resisted by the frames. Because all eave load is assumed to be
transferred to shearwalls, no special analysis tools or design tables are required
to determine load distribution between diaphragms and post-frames. This simplifies
the entire diaphragm design process. This simplified procedure is referred to as
rigid roof design (Bender and others, 1991).

5.8.2 Calculation. When (1) the shearwalls and roof/ceiling diaphragm assembly are
assumed to be infinitely rigid, (2) the only applied loads with horizontal
components are due to wind, and (3) wind pressure is uniformly distributed on each
wall and roof surface, then the maximum shear force in the diaphragm assembly is
given as:

Vh = L (hwr qwr hlr qlr + hww fw qww hlw fl qlw) / 2 (5-20)

where:

Vh =
L= hwr = hlr = hww = hlw =

maximum diaphragm element shear force, lbf (N) building length, ft (m) windward
roof height, ft (m) leeward roof height, ft (m) windward wall height, ft (m)
leeward wall height, ft (m)

5-18

National Frame Builders Association

Post-Frame Building Design Manual

qwr = qlr = qww = qlw = fw = fl =

design windward roof pressure, lbf/ft2 (N/m2) design leeward roof pressure, lbf/ft2
(N/m2) design windward wall pressure, lbf/ft2 (N/m2) design leeward wall pressure
lbf/ft2 (N/m2) frame-base fixity factor, windward post frame-base fixity factor,
leeward post

Inward acting wind pressures have positive signs, outward acting pressures are
negative (figure 5.8). As previously noted, frame-base fixity factors, fw and fl,
determine how much of the total wall load is transferred to the eave, and how much
is transferred directly to the ground. The greater the resistance to rotation at
the base of a wall, the more load will be attracted directly to the base of the
wall. For substantial fixity against rotation at the groundline, set the frame-base
fixity factor(s) equal to 3/8. For all other cases, set the frame-base fixity
factor(s) equal to 1/2.

For symmetrical base restraint and frame geometry, equation 5-20 reduces to:

Vh = L [hr (qwr qlr) + hw f (qww qlw)] / 2 (5-21) where:

hr = hw = f=

roof height, ft (m) wall height, ft (m) frame-base fixity factor for both leeward
and windward posts

5.8.3 Application. The Vh value calculated using equation 5-20 (or 5-21) is always
a conservative estimate of the actual maximum shear force (due to wind) in a
diaphragm assembly. This estimate becomes increasingly conservative as the amount
of load resisted by interior post-frames increases. It follows that equations 5-20
and 5-21 are most accurate when diaphragm stiffness is considerably greater than
interior post-frame stiffness. This tends to be the case in buildings that are
relatively wide and/or high, and in buildings where individual posts offer no
resistance to rotation (i.e., the posts are more-or less pin-connected at both the
floor and eave lines).

Output from a DAFI analysis of a building with relatively high diaphragm and
shearwall stiffness values is presented in figure 5.9. This output shows less than
3% of the total horizontal eave load being resisted by the interior frames.
Although rigid roof design expedites calculation of maximum diaphragm shear forces,
the design procedure does not provide estimates of sidesway restraining force for
interior post-frame design.
5.9 References
Anderson, G.A., D.S. Bundy and N.F. Meador. 1989. The force distribution method:
procedure and application to the analysis of buildings with diaphragm action.
Transactions of the ASAE 32(5):1781-1786.
ASAE. 1999a. EP484.2 Diaphragm design of metal-clad wood-frame rectangular
buildings. ASAE Standards, 46th Ed., ASAE, St. Joseph, MI.
ASAE. 1999b. EP558.1 Load tests for metalclad wood-frame diaphragms. ASAE
Standards, 46th Ed., ASAE, St. Joseph, MI.
Bender, D. A., T. D. Skaggs and F. E. Woeste. 1991. Rigid roof design for post-
frame buildings. Applied Engineering in Agriculture 7(6):755-760.
Bohnhoff, D. R., P. A. Boor, and G. A. Anderson. 1999. Thoughts on metal-clad wood-
frame diaphragm action and a full-scale building test. ASAE Paper No. 994202, ASAE,
St. Joseph, MI.
Bohnhoff, D. R. 1992. Expanding diaphragm analysis for post-frame buildings.
Applied Engineering in Agriculture 8(4):509-517.
Gebremedhin, K.G. 1987a. SOLVER: An interactive structures analyzer for
microcomputers. (Version 2). Northeast Regional Agricultural Engineering Service.
Cornell University, Ithaca, NY.
Gebremedhin, K.G. 1987b. METCLAD: Diaphragm design of metal-clad post-frame
buildings using microcomputers. Northeast Regional Agricultural Engineering
Service. Cornell University, Ithaca, NY.

5-19

National Frame Builders Association


McGuire, P.M. 1998. One equation for compatible eave deflections. Frame Building
News 10(4):39-44.
Meader, N.F. 1997. Mathematical models for lateral resistance of post foundations.
Trans of ASAE, 40(1):191-201.
Niu, K.T. and K.G. Gebremedhin. 1997. Evaluation of interaction of wood framing and
metalcladding in roof diaphragms. Transactions of the ASAE 40(2):465-476.
Pollock, D. G., D. A. Bender and K. G. Gebremedhin. 1996. Designing for chord
forces in post-frame roof diaphragms. Frame Building News 8(5):40-44.
Purdue Research Foundation. 1986. Purdue plane structures analyzer. (Version 3.0).
Department of Forestry and Natural Resources. Purdue University, West Layfette, IN.
Williams, G. D. 1999. Modeling metal-clad wood-framed diaphragm assemblies. Ph.D.
diss., University of Wisconsin-Madison, Madison, WI.
Wright, B.W. 1992. Modeling timber-framed, metal-clad diaphragm performance. Ph.D.
diss. The Pennsylvania State University, University Park, PA.

Post-Frame Building Design Manual

5-20

National Frame Builders Association

Post-Frame Building Design Manual

Chapter 6: METAL-CLAD WOOD-FRAME DIAPHRAGM PROPERTIES

6.1 Introduction
6.1.1 General. One of the first steps in diaphragm design is to establish in-plane
shear strength and stiffness values for each identified diaphragm section. In most
post-frame buildings, these diaphragm sections consist of corrugated metal panels
that have been screwed or nailed to wood framing. Behavior of these metalclad wood-
frame (MCWF) diaphragms is complex, and consequently, has been the subject of
considerable research during the past 20 years. In addition to improving overall
design, this research has led to improved methods for predicting metal-clad wood-
frame diaphragm strength and stiffness.
6.1.2 Predicting Diaphragm Behavior. There are essentially three procedures for
predicting the strength and stiffness of a building diaphragm. First, an exact
replica of the building diaphragm (a.k.a. a full-size diaphragm) can be built and
tested to failure. Second, a smaller, representative section of the building
diaphragm can be built and laboratory tested. The strength and stiffness of this
test assembly are then extrapolated to obtain strength and stiffness values for the
building diaphragm. Lastly, diaphragm behavior can be predicted using finite
element analysis software. The latter requires that the strength and stiffness
properties of individual component (e.g., wood framing, mechanical connections,
cladding) be known.
Of the three procedures for predicting metal-clad wood-frame diaphragm properties,
only the second one extrapolation of diaphragm test assembly data - is commonly
used. This is because testing full-size diaphragms is simply not practical (a new
test would have to be conducted every time overall dimensions changed), and finite
element analysis of MCWF diaphragms is, for practical purposes, still in a
developmental stage. The later can be attributed to the fact that the large number
of variables affecting diaphragm structural properties, as well as the nonlinear
behavior of some variables, has thus far precluded the development of a quick and
reasonably accurate closed-form approxi-

mation of diaphragm strength and stiffness.


6.1.3 ASAE EP558 and EP484. Construction specifications and testing procedures for
diaphragm test assemblies are given in ASAE EP558 Load Test for Metal-Clad Wood-
Frame Diaphragms (ASAE, 1999b). EP558 also gives equations for calculating
diaphragm test assembly strength and stiffness. These calculations along with
construction specifications and testing procedures from EP558 are outlined in
Section 6.3: Diaphragm Assembly Tests. For additional details and further
explanation of testing procedures, readers are referred to the ASAE EP558
Commentary (ASAE, 1999b).
ASAE EP484, which was introduced in detail in Chapter 5, contains the equations for
extrapolating diaphragm test assembly properties for use in building design. These
calculations are presented in Section 6.4: Building Diaphragm Properties.
6.2 Design Variables
6.2.1 General. Many variables affect the shear stiffness and strength of a
structural diaphragm, including: overall geometry, cladding characteristics, wood
properties, fastener type and location, and blocking. A short description of each
of these variables follows.
6.2.2. Geometry. Geometric variables include: spacing between secondary framing
members (e.g. purlins), spacing between primary framing members (e.g.,
trusses/rafters), and overall dimensions. With respect to overall dimensions,
diaphragm depth is measured parallel to primary frames, diaphragm length is
measured perpendicular to primary frames. In most structures, the overall length of
a roof diaphragm is equal to the length of the building.
6.2.3 Cladding. Cladding type (e.g., wood, metal, fiberglass, etc.) is a
significant design variable. Coverage (and examples) in this design manual is
limited to corrugated metal cladding. Important design characteristics of this type
of cladding include: base metal (e.g., steel,

6-1

National Frame Builders Association


Post-Frame Building Design Manual

aluminum), base metal thickness, panel profile, and individual sheet width and
length.
6.2.4 Wood Framing. The species, moisture content and specific gravity of wood used
in the framing system will not only affect the structural properties of the wood
members, but also the shear stiffness and strength of mechanical connections
between wood members and between wood members and cladding.
6.2.5 Mechanical Connections. Type (screw or nail), size, and relative location of
mechanical fasteners used to join components significantly impact diaphragm
properties. Fasteners are primarily defined by what they connect. Major categories
include purlin-to-rafter, sheet-topurlin, and sheet-to-sheet (see figure 6.1).
Sheet-to-sheet fasteners are more commonly referred to as stitch or seam fasteners.
Removing stitch fasteners can dramatically reduce the

shear strength and stiffness of a diaphragm. Sheet-to-purlin fasteners are also


defined by their location (i.e., end, edge, and field). A sheet-to-purlin fastener
may be located in a rib or in the flat of a corrugated metal panel. Locating
fasteners in the flat generally produces stronger and stiffer diaphragms. The
nonlinear nature of fastener performance is one of the more complex variables
affecting diaphragm stiffness.
6.2.6 Blocking. When secondary framing members are installed above primary framing
(e.g. top running purlins) or below primary framing (e.g. bottom-running ceiling
framing), cladding can only be fastened directly to the secondary framing (see
figure 6.1). In such cases, blocking is often placed between the cladding and
primary framing to increase shear transfer between the components. This is commonly
done at locations where diaphragms and shearwalls intersect.

Sheet-to-Purlin Fasteners (Field)

Stitch Fastener

Purlin-to-Rafter Fastener

Sheet-to-Purlin Fastener (Edge) Sidelap Seam

Sheet-to-Purlin Fasteners (End) Corrugated Metal Cladding

Rake Board

Blocking between purlins

Purlin Rafter/Truss Top Chord

Figure 6.1. Components of a metal-clad wood-frame roof diaphragm.

6-2

National Frame Builders Association

Post-Frame Building Design Manual

F Rafter / Truss chord 2

Applied force, P
G

Purlins
Notes:

Cladding

1. Force P may be alternately applied at point H 2. Locate gages 2 and 4 on the


edge purlins 3. Locate gages 1 and 3 on the rafter / truss chord

Direction of Corrugations b = Test assembly length

4 E

a = Test assembly width

13
(a)

Deflection gage location and direction of measured


deflection (typ.)

3a = Test assembly width

F Rafter / Truss chord

Applied force, P/2 I

Purlins

Applied force, P/2

Cladding

b = Test assembly length Direction of corrugations

E JL
23 1 Deflection gage location
and direction of measured deflection (typ.)
Notes: 1. The applied forces may alternately be applied at points J and L 2. Locate
gages 1, 2, 3 and 4 on the rafters/ truss chords

H 4

(b)
Figure 6.2. (a) Cantilever test configuration, and (b) Simple beam test
configuration for diaphragm test assemblies.

6-3

National Frame Builders Association

Post-Frame Building Design Manual

6.3 Diaphragm Test Assemblies


6.3.1 Construction. With the exception of overall length and width, a diaphragm
test assembly is required to be identical to the diaphragm in the building being
designed. Specifically, frame members must be of identical size, spacing, species
and grade; metal cladding must be identical in composition, profile and thickness;
and fastener type and location must be the same. ASAE EP558 has established minimum
sizes for diaphragm test assemblies to ensure that there is not too great a
difference between the size of a diaphragm test assembly and the actual building
diaphragm.

6.3.2 Test Configurations. ASAE EP558 allows for two different testing
configurations: a cantilever test and a simple beam test (figures 6.2a and 6.2b,
respectively). In both figures 6.2a and 6.2b, variable a represents the spacing
between rafters/trusses (a.k.a. the frame spacing). This spacing should be equal
to, or a multiple of, the frame spacing in the building being designed.

6.3.3 Shear Strength. The allowable design shear strength, of a diaphragm test
assembly is equal to 40% of the ultimate strength of the assembly. In equation
form:

Cantilever test: va = 0.40 Pu / b


Simple beam test: va = 0.40 Pu / (2b)
where:

(6-1) (6-2)

va =
Pu = =
b=

allowable design shear strength, lbf/ft (N/m) ultimate strength, lbf (N) total
applied load at failure assembly length, ft (m) (see figure 6.2)

If one or more of the test assembly failures were initiated by lumber breakage or
by failure of the fastenings in the wood, then the allowable design shear stress
must be adjusted to account for test duration. To adjust from a total elapsed
testing time of 10 minutes to a normal load duration of ten years, divide va by a
factor of 1.6.

When this reduction is not applied (as would be the case when test assembly failure
is not initiated by wood failure), the NDS load duration factor, CD, can not be
used to increase the allowable design shear strength during building design.
Completely separate of the load duration factor adjustment is the 30% increase in
allowable strengths allowed by most codes for wind loadings (see Section 3.9.4).

6.3.4 Shear Stiffness. The procedure for determining the effective shear modulus of
a test assembly begins with calculation of the adjusted load-point deflection, DT.
This value takes into account rigid body rotation/translation during assembly test
and is calculated as follows:

Cantilever test: DT = D3 D1 (a/b) (D2 + D4)


Simple beam test: DT = (D2 + D3 D1 D4) / 2
where:

(6-3) (6-4)

DT = adjusted load point deflection, in. (mm)


D1, D2, D3, and D4 = deflection measurements, in. (mm) (see figure 6.2)
a = assembly width, ft (m) b = assembly length, ft (m)
The effective in-plane shear stiffness, c, for a diaphragm test assembly is defined
as the ratio of applied load to adjusted load point deflection at 40% of ultimate
load. In equation form:

Cantilever test: c = 0.4 Pu / DT,d


Simple beam test: c = 0.2 Pu / DT,d
where:

(6-5) (6-6)

c= DT,d =

effective in-plane shear stiffness, lbf/in. (N/mm) adjusted load-point deflection,


DT, at 0.4 Pu, in. (mm)

The in-plane shear stiffness for the diaphragm test assembly, c, is converted to an
effective shear modulus for the test assembly, G, as:

6-4

National Frame Builders Association

Post-Frame Building Design Manual

G = c (a/b)

(6-7)

where:

G = effective shear modulus of the test assembly, lbf/in (N/mm)

6.4 Building Diaphragm Properties

6.4.1 General. As described in Chapter 5, each building diaphragm is sectioned for


analysis. Each of these sections must be assigned a horizontal stiffness value, ch,
and an allowable load, va.

6.4.2 Shear Strength The allowable design shear strength of a building diaphragm is
equal to that calculated for the diaphragm test assembly. Consequently, to
calculate the total in-plane shear load that a building diaphragm can sustain,
simply multiply the allowable design shear strength, va, by the slope length of the
building diaphragm.

6.4.3 In-Plane Shear Stiffness. The in-plane shear stiffness, cp, of a building
diaphragm section is calculated from the effective shear modulus, G, of the
diaphragm test assembly using the following equation:

cp = or
cp =

G bs s
G bh s cos()

(6-8) (6-9)

where:
G= bS = s= bh = =

effective shear stiffness of test assembly, lbf/in (N/mm) slope length of building
diaphragm section being modeled, ft (m) width of the building diaphragm section
being modeled, ft (m) horizontal span length of building diaphragm section, ft (m)
slope of the building diaphragm section, degrees

Implicit in equation 6-8 is the assumption that the total shear stiffness of a
building diaphragm is a linear function of length.

6.4.4 Horizontal Shear Stiffness. The horizontal shear stiffness, ch, of a building
diaphragm section is related to its in-plane shear stiffness as follows:

ch = cp cos2()

(6-10)

or

ch = G bh cos() / s

(6-11)

6.5 Building Shearwall Properties

6.5.1 General. The same procedure used to determine the strength and stiffness of
building diaphragms is used to determine the strength and stiffness of building
shearwalls. That is, representative test assemblies are loaded to failure, to
determine their shear strength and stiffness. These properties are then linearly
extrapolated to obtain strength and stiffness values for the building shearwall(s).
6.5.2 Shearwall Test Assemblies. ASAE EP558 also contains guidelines for
construction and testing of shearwall test assemblies. With the exception of
overall length and width, a shearwall test assembly is required to be identical to
the shearwall in the building being designed. Specifically, frame members must be
of identical size, spacing, species and grade; cladding must be identical; and
fastener type and location must be the same.

6.6 Tabulated Data

6.6.1 Sources. Testing replicate samples of diaphragm test assemblies can get
expensive. For this reason, a designer may choose not to conduct his/her own
diaphragm tests, relying instead on designs that have been previously tested by
others. Information on many tested designs is available in the public domain.
Cladding manufacturers may have additional test information on assemblies that
feature their own products.
6.6.2 Example Tabulated Data. Table 6.1 contains design details and engineering
properties for roof diaphragm tests assemblies. The information in this table
represents a small percentage of available data.

6-5

National Frame Builders Association

Post-Frame Building Design Manual

Table 6.1. Steel-Clad Roof Diaphragm Assembly Test Data

Test Assembly Number Test Configuration Cladding


Manufacturer/Trade Name
Base Metal Thickness Gauge Major Rib Spacing, inches Major Rib Height, inches Major
Rib Base Width, inches Major Rib Top Width, inches Yield Strength, ksi Overall
Design Width, feet Length, b , feet Purlin Spacing, feet Rafter Spacing, feet
Purlin Location Purlin Orientation Number of Internal Seams Wood Properties Purlin
Size Purlin Species and Grade Rafter Species and Grade Stitch Fastener Type Length,
inches Diameter On Center Spacing, inches Sheet-to-Purlin Fasteners Type Length,
inches Diameter Location in Field Location on End Avg. On-Center Spacing in Field,
in. Avg. On-Center Spacing on End, in.
Purlin-to-Rafter Fastener
Engineering Properties Ultimate Strength, Pu, lbf. Allowable Shear Strength, va,
lbf/ft Effective In-Plane Stiffness, c ,lbf/in Effective Shear Modulus, G, lbf/in
Reference

1 Cantilever

2 Cantilever

3 Cantilever

4 Cantilever

Wick Agri Panel 28 12 0.75 1.25 0.375 50

Wick Agri Panel 28 12 0.75 1.25 0.375 50

Wick Agri Panel 29 12 0.75 1.25 0.375 80

Midwest Manufacturing. 29 12 1.0 2.5 0.5 80

9 12 2 9 Top running On edge 2

9 12 2 9 Top running On edge 2

9 12 2 9 Top running On edge 2

6 12 2 6 Top running On edge 2

2- by 4-inch No.1 & 2 SPF


No. 1 SYP

2- by 4-inch No.1 & 2 SPF


No. 1 SYP

2- by 4-inch No.1 & 2 SPF


No. 1 SYP

2- by 4-inch No.2 SYP No. 1 SYP

None

Screw 1.0 #10 24

Screw 1.0 #10 24

EZ Seal Nail 2.5 8d 24

Screw 1.0 #10


In Flat In Flat
12 6 60d Threaded Hardened Nail
Screw 1.0 #10
In Flat In Flat
12 6 60d Threaded Hardened Nail

Screw 1.0 #10


In Flat In Flat
12 6 60d Threaded Hardened Nail

EZ Seal Nail 2.5 8d


Major Rib In Flat 12 12
60d Threaded Hardened Nail

2140 71
1625 1220 Anderson, 1989

3390 113 2720 2040 Anderson, 1989

3220 107 2720 2040 Anderson, 1989

1930 64
1590 795 Wee & Anderson, 1990

6-6

National Frame Builders Association

Post-Frame Building Design Manual

Table 6.1. cont., Steel-Clad Roof Diaphragm Assembly Test Data

Test Assembly Number Test Configuration Cladding


Manufacturer/Trade Name
Base Metal Thickness Gauge Major Rib Spacing, inches Major Rib Height, inches Major
Rib Base Width, inches Major Rib Top Width, inches Yield Strength, ksi Overall
Design Width, feet Length, b , feet Purlin Spacing, feet Rafter Spacing, feet
Purlin Location Purlin Orientation Number of Internal Seams Wood Properties Purlin
Size Purlin Species and Grade Rafter Species and Grade Stitch Fastener Type Length,
inches Diameter On Center Spacing, inches Sheet-to-Purlin Fasteners Type Length,
inches Diameter Location in Field Location on End Avg. On-Center Spacing in Field,
in. Avg. On-Center Spacing on End, in.
Purlin-to-Rafter Fastener
Engineering Properties Ultimate Strength, Pu, lbf. Allowable Shear Strength, va,
lbf/ft Effective In-Plane Stiffness, c ,lbf/in Effective Shear Modulus, G, lbf/in
Reference

5 Cantilever

6 Cantilever

7 Cantilever

8 Cantilever

Midwest Manufacturing 29 12 1.0 2.5 0.5 80

Grandrib 3
29 12 0.75 1.75 0.5 80
Grandrib 3
29 12 0.75 1.75 0.5 80

Walters STR-28
28 12 0.94
80

6 12 2 6 Top running On edge 2

9 12 2 9 Top running On edge 2

9 12 2 9 Top running On edge 2

9 16 2 9 Top running On edge 2

2- by 4-inch No.2 SYP No. 1 SYP

2- by 4-inch No.2 DFL No. 2 DFL

2- by 4-inch No.2 SPF No. 2 SPF

2- by 4-inch No.2 SYP 1950f1.7E SYP

EZ Seal Nail 2.5 8d 24

None

None

Screw 1.5 #10 24

Screw 0.75 #12 In Flat In Flat


6 6 60d Threaded Hardened Nail

Screw 1.0 #10


In Flat In Flat
12 6 1-60d Spike + 2-10d Toenails

Screw 1.0 #10


In Flat In Flat
12 6 1-60d Spike + 2-10d Toenails

Screw 1.5 #10


In Flat In Flat 12 and 18
12 60d Threaded Hardened Nail

3995 133 2980 1490 Wee & Anderson, 1990

3300 110 2920 2190 Lukens & Bundy, 1987

2775 93
2950 2210 Lukens & Bundy, 1987

4884 122 3890 2190 Bohnhoff and others, 1991

6-7

National Frame Builders Association


Post-Frame Building Design Manual

Table 6.1. cont., Steel-Clad Roof Diaphragm Assembly Test Data

Test Assembly Number

9 10 11 12

Test Configuration

Simple Beam

Cladding

Type

Regular Leg Extended Leg Regular Leg Extended Leg

Base Metal Thickness Gauge

29

Major Rib Spacing, inches

Major Rib Height, inches

0.62

Major Rib Base Width, inches

1.75

Major Rib Top Width, inches

0.75

Yield Strength, ksi

80

Overall Design

Width, feet

36

Length, b , feet

12

Purlin Spacing, feet

Rafter Spacing

Pair of rafters every 12 feet (each pair spaced 6 in. apart)


Purlin Location

Top running and lapped

Inset

Purlin length, ft

13.2 and 12.0

11.25

Purlin Attachment

To special blocking nailed between each pair of rafters

To joist hanger attached to rafters

Purlin Orientation

On edge

Number of Internal Seams

11

Wood Properties

Purlin Size

2- by 6-inch

Purlin Species and Grade

No.2 DFL and 1650f DFL

Rafter Species and Grade

No. 2 DFL

Stitch Fastener*

Type

None

Screw*

None

Screw*

Length, inches

1.5 1.5

Diameter
#10 #10

On Center Spacing, inches

24 24

Sheet-to-Purlin Fasteners

Type

Screw

Length, inches

1.5

Diameter

#10

Location in Field

In Flat

Location on End

In Flat

Avg. On-Center Spacing in Field, in.

Avg. OnCenter Spacing on End, in.

Engineering Properties

Ultimate Strength, Pu, lbf. Allowable Shear Strength, va, lbf/ft Effective In-Plane
Stiffness, c ,lbf/in

6950 116 4700

7850 131 7500

6400 107 3700

6950 116 4400

Effective Shear Modulus, G, lbf/in

4700

7500

3700

4400
Reference

NFBA, 1996

* Because of the extended leg, screws installed in the flat at overlapping seams
function as stitch fasteners.

6-8

National Frame Builders Association

Post-Frame Building Design Manual

Table 6.1. cont., Steel-Clad Roof Diaphragm Assembly Test Data

Test Assembly Number Test Configuration Cladding


Manufacturer/Trade Name
Base Metal Thickness Gauge Major Rib Spacing, inches Major Rib Height, inches Major
Rib Base Width, inches Major Rib Top Width, inches Yield Strength, ksi Overall
Design Width, feet Length, b , feet Purlin Spacing, feet
Rafter Spacing, feet
Purlin Location Purlin Orientation Number of Internal Seams Wood Properties Purlin
Size
Purlin Species and Grade
Rafter Species and Grade Stitch Fastener
Type Length, inches Diameter On Center Spacing, inches Sheet-to-Purlin Fasteners
Type Length, inches
Diameter
Location in Field Location on End Avg. On-Center Spacing in Field, in. Avg. On-
Center Spacing on End, in.
Purlin-to-Rafter Fastener
Engineering Properties Ultimate Strength, Pu, lbf. Allowable Shear Strength, va,
lbf/ft Effective In-Plane Stiffness, c ,lbf/in Effective Shear Modulus, G, lbf/in
Reference

13 Simple Beam

14 Simple Beam

15 Simple Beam

Metal Sales Pro Panel II 30


9.0

Metal Sales Pro Panel II 30


9.0

McElroy Metal Max Rib 29 9.0 0.75 1.75

104 104 80

24
12
2.33 Pair of rafters every 12 feet (each pair spaced 6 in. apart)
Top running
On edge
8

24
12
2.33 Pair of rafters every 12 feet (each pair spaced 6 in. apart)
Top running
On edge
8

24 12 2
8
Top running NA 7

2- by 6-inch 1650f 1.5E SPF 1650f 1.5E SPF

2- by 6-inch 1650f 1.5E SPF 1650f 1.5E SPF

Mac-Girt steel hat section: 1.5 in. tall, 3.2 in. wide, 18 ga.
2250f 1.9E SP

Screw 0.625 #12


9

None

None

Screw 1.5
#10
In Flat In Flat
9 4.5

Screw 1.5
#10 in field #14 in ends
In Flat In Flat
9 4.5

Screw 1.0
#14
In Flat In Flat 18 (3 screws/sheet) 9 (4 screws/sheet) Two - #12 x 1.6 in.
screws/joint

9600 160 7680 7680 Townsend, 1992

6600 110 7100 7100 Townsend, 1992

8645 144 10700 7130 Myers, 1994

6-9

National Frame Builders Association

Post-Frame Building Design Manual

6.7 Example Calculations


A designer wishes to find ch and va for roof diaphragm sections in a gable-roofed
building with roof slopes of 4-in-12. Distance between eaves is 36 feet, and post-
frame spacing, s, is 10 feet.
A cantilever test of a representative diaphragm test assembly with a width, a, of
10 feet and a length, b, of 12 feet, yields an ultimate strength, Pu of 3900 lbf
and an effective inplane stiffness, c, of 4000 lbf/in. The test assembly failure
was not wood related, therefore the ultimate strength was not adjusted for load
duration.
Equation 6-1: va (test assembly) = 0.40 Pu / b
va (test assembly) = 0.40 (3900 lbf) /12 ft = 130 lbf/ft
Equation 6-7: G = c (a/b)
G = (4000 lbf/in) (10 ft/12 ft) = 3333 lbf/in.
Equation 6-11: ch = G bh cos() / s
ch = (3333 lbf/in) (36 ft / 2) (cos 18.4) / 10 ft = 5690 lbf/in.
The horizontal stiffness, ch of 5690 lbf/in represents a single diaphragm section
that runs from eave to ridge and has a width of 10 feet.
va (diaphragm) = 1.30 va (test diaphragm) = 1.3 (130 lbf/ft) = 169 lbf/ft
As described in Section 3.9.4, the allowable strength of a diaphragm can generally
be increased by 30% when wind or seismic loads are acting in combination with other
loads.

6.8 References
Anderson, G.A. 1989. Effect of fasteners on the stiffness and strength of timber-
framed metalclad roof sections. ASAE Paper No. MCR89501. ASAE, St. Joseph, MI.
ASAE. 1999a. EP484.2: Diaphragm design of metal-clad, wood-frame rectangular
buildings. ASAE Standards, 46th Edition. St. Joseph, MI.
ASAE. 1999b. ASAE EP558: Load tests for metal-clad wood-frame diaphragms. ASAE
Standards, 46th edition. ASAE, St. Joseph, MI.
Anderson, and P.A. Boor. 1991. Influence of insulation on the behavior of steel-
clad wood frame diaphragms. Applied Engineering in Agriculture 7(6):748-754.

Lukens, A.D., and D.S. Bundy. 1987. Strength and stiffnesses of post-frame building
roof panels. ASAE Paper No. 874056. ASAE, St. Joseph, MI.
Myers, N.C. 1994. McElroy Metal Post Frame Roof Diaphragm Test. Test Report 94-418.
Progressive Engineering, Inc., Goshen, IN.
NFBA. 1996. 1996 Diaphragm Test. National Frame Builders Association, Inc.,
Lawrence, KS.
Townsend, M. 1992. Alumax test report: diaphragm loading on roofs and end wall
sections. Alumax Building Products, Perris, CA.
Wee, C.L. and G.A. Anderson. 1990. Strength and stiffness of metal clad roof
section. ASAE Paper No. 904029. ASAE, St. Joseph, MI.

6-10

National Frame Builders Association

Post-Frame Building Design Manual

Chapter 7: POST PROPERTIES

7.1 Introduction
7.1.1 Types. Several different post types are currently used in post-frame
construction. The most common of these are laminated lumber posts. Solid-sawn posts
are still used by most builders, but not to the extent they were a decade ago.
Parallel strand lumber (PSL) and laminated veneer lumber (LVL) products are gaining
in popularity. Use of these and other engineered lumber products as posts in
postframe buildings can only be expected to increase as the relative cost of these
products decreases.
7.1.2 Preservative Treatment. If posts are to be embedded, they must be
preservative treated to avoid decay. General issues of preservative treatment have
already been presented in Chapter 4. Discussion in this chapter will focus on the
structural aspects of post selection and design.
7.2 Solid-Sawn Posts
7.2.1 Size. Post size varies considerably with building geometry and design loads.
The most common sizes are 6- by 6-inch, 6- by 8-inch, and 4- by 6-inch. Although
both S4S (Surfaced on 4 Sides) and rough sawn posts are available, most rough sawn
posts are not graded and therefore are generally only used in code exempt
applications.
7.2.2 Wood Species. Species of wood used in posts depends on local availability and
on preservative treatment needs. Commonly used species includes Southern Pine,
Douglas Fir and Ponderosa Pine.
7.2.3 Design Properties. NDS design values for species and grades typically used in
postframe construction are given in table 7.1. These values have been adjusted for
conditions of use in which wood moisture content exceeds 19% for extended time
periods, as is the case for embedded posts. To apply the values in table

7.1, a post must be graded by an approved grading agency and stamped accordingly.
7.2.4 Current Demand. Solid-sawn post use in post-frame construction is on the
decline, primarily because posts of acceptable size, length and quality are
increasingly difficult to obtain. The scarcity of long posts in structural sizes
has made laminated posts more price competitive. Additionally, laminated post
prices are typically constant on a per-foot basis regardless of length, while the
cost of solid-sawn posts increases exponentially with length.
7.3 Laminated Lumber Posts
7.3.1 General. Laminated lumber posts are posts that are fabricated by joining
together individual pieces of dimension lumber, most commonly 2- by 6-inch, 2- by
8-inch and 2- by 10-inch members. Structural properties of the finished product
vary significantly depending on the means of lamination and the presence or absence
of joints in individual layers. Laminates are either glued together or joined
together with mechanical fasteners (i.e., nails, screws, bolts, shear transfer
plates, metal plate connectors).
7.3.2 Advantages. By combining individual laminates to build up a desired cross-
section, the statistical probability that a strength-reducing characteristic of
wood (such as a knot) would exist through the entire cross section is greatly
diminished. Consequently, laminated posts have more uniform strength and stiffness
properties than solid-sawn posts. This increased reliability results in higher
allowable design values.
7.3.3 Laminate Orientation. Laminated post strength is dependent on orientation of
individual laminates with respect to the principal load direction. If a post is
designed (and positioned within the structure) to resist loads acting on the edge,
or narrow face, of the laminates, the post is said to be vertically-laminated
(figure 7.1a). If a post is oriented such that the applied load acts on the wide
face of the laminates, the post is said to be horizontally-laminated (figure 7.1b).

7-1

National Frame Builders Association

Post-Frame Building Design Manual

Table 7.1. Design Stresses for Selected Species and Grades of Solid-Sawn Posts *

Species and Grade

Bending, Fb

Design Values in Pounds per Square Inch (psi)

Tension Parallel to Grain, Ft

Shear Parallel to Grain, Fv

Compression Perpendicular to Grain, Fc

Compression Parallel to Grain, Fc


Modulus of Elasticity, E

Douglas Fir-Larch

Sel Str

1500

1000

85

420

1045

1,600,000

No. 1

1200

825

85

420

910 1,600,000

No. 2

750 475

85

420

430 1,300,000

Northern Pine Sel Str No. 1 No. 2

1150 950 500

800 650 375

65 65 65

290 290 290

820 1,300,000 730 1,300,000 340 1,000,000

Ponderosa Pine Sel Str No. 1 No. 2

1000 825 475

675 550 325


65 65 65

360 360 360

730 1,100,000 635 1,100,000 295 900,000

Southern Pine Sel Str No. 1 No. 2

1500 1350 850

1000 900 550

110 110 100

375 375 375

950 1,500,000 825 1,500,000 525 1,200,000

* From the National Design Specifications (NDS) for wood under wet-use conditions,
AF&PA (1997b). Values are for lumber in the size category Posts and Timbers.

Load

Load

HV

V VH

HV (a) (b)
Figure 7.1. (a) Vertically laminated, and (b) horizontally laminated post cross-
sections.

7.4 Glued-Laminated (Glulam) Posts


7.4.1 Advantages. For a given species and grade of lumber, glued-laminated posts
have higher allowable design values than solid-sawn posts and most spliced
mechanically-laminated posts (see Section 7.6). Glued-laminated posts exhibit
complete composite action, that is, the glue interface is of sufficient integrity
that it is assumed that there is no slip between laminates regardless of load
level. With no slip between layers, glued-laminated posts behave much like solid-
sawn posts, and are very effective in carrying biaxial bending loads.
7.4.2 Vertical Lamination. Glued-laminated posts that have a rather square cross-
section are typically designed as vertically-laminated components; that is, they
are designed to resist primary bending moments about an axis perpendicular to the
wide faces of individual laminations (Axis V-V, figure 7.1b). This class of posts

7-2

National Frame Builders Association

Post-Frame Building Design Manual

(cross-sectional aspect ratios less than 1.5) are commonly used as posts in post-
frame buildings.
7.4.3 Horizontal Lamination. In contrast to the glued-laminated posts commonly used
in postframe construction, deep glulam beams (e.g. door headers) are generally
designed as horizontally laminated components (figure 7.1a). Lumber is used more
efficiently in these assemblies by placing higher grade lumber in outer laminates
where bending stresses are higher, and using lower grade lumber near the center
where bending stresses are low. In addition, horizontal lamination facilitates the
manufacture

of curved members.
7.4.4 Design Properties. Design properties for both horizontally- and vertically-
laminated glulams are published by American Institute of Timber Construction (AITC,
1985) and AF&PA (1997b). Values for selected vertically-laminated assemblies are
listed in table 7.2. These values are for dry-use conditions and normal load
duration. In actual application, glulam design values must be adjusted by
applicable factors involving curvature, volume, beam stability and column
stability. These factors (and direction regarding their application) can also be
found in the two references cited in this paragraph.

Table 7.2. Design Values for Vertically Glued Laminated Posts a

AITC Combin
ation Symbol

Lumber Grade

MOE, million
psi

Extreme Fiber in Bending, psi

Bending about V-V Axis. 4 or


3 Lams More Lams

Bending about
H-H Axis.

Tension Parallel to Grain,


psi

Compression Parallel to Grain, psi

2 or 3 Lams

4 or More Lams

Douglas Fir- Larch 13 Dense Sel Str 12 Sel Str 11 No. 1 Dense 10 No. 1 9 No. 2
Dense 8 No. 2

2.0 2300 2400 1.8 1950 2100 2.0 2100 2300 1.8 1750 1950 1.8 1800 1850 1.6 1550 1600

2200 1900 2100 1750 1600 1350

1600 1400 1500 1300 1150 1000

1950 1650 1700 1450 1350 1150

2300 1950 2300 1950 1800 1550

Hem-Fir 21 20 19

Sel Str No. 1 No. 2


1.6 1650 1750 1.6 1500 1550 1.4 1300 1350

1500 1350 1150

1100 975 850

1350 1250 975

1450 1450 1300

Southern Pine 52 Dense Sel Str 51 Sel Str 50 No. 1 Dense 49 No. 1 48 No. 2 Dense 47
No. 2

1.9 2300 2400

2100

1500

1.7 1950 2100

1750

1300

1.9 2100 2100 b 1800 b 1550

1.7

1750

1850 b

1550 b

1350

1.7

1800

1850 b

1600 b

1400

1.4

1550

1600 b

1350 b

1200

1850 1600 1700 1450 1350 1150


2200 1900 2300 2100 2200 1900

Wet Service Factor, CM c 0.833 0.80

0.80

0.80

0.80

0.73

0.73

a From the National Design Specifications (NDS), AF&PA (1997b). b Values reflect
the removal of the more restrictive slope-of-grain requirements. c The tabulated
values are applicable when in-service moisture content is less than 16%. To obtain

wet-use values, multiply the tabulated values by the factors shown.

7-3

National Frame Builders Association

Post-Frame Building Design Manual

7.4.5 Manufacturing Requirements. For glulam design values apply, tight quality
control must be maintained during the laminating process. The AITC has published
standards for the design (AITC, 1985) and manufacturing (AITC, 1988) of glued-
laminated members. Fabrication procedures for the members must conform to an
additional standard (AITC, 1983), which covers physical construction issues as well
as quality control, testing and marking procedures. The rigorous requirements for
construction, as well as the planing that must be performed (individual laminates
prior to lamination, and the finished member after lamination completion), combine
to essentially eliminate the possibility of on-site fabrication. These factors also
increase product price, however, for many applications, higher design properties
justify the higher cost.
7.4.6 End Joints. Posts of any length can be created by end-joining individual
laminates. The most common glued end joint is the finger joint. Although finger
joining is a common manufacturing process, only a few manufacturing facilities have
the capability of producing finger joints that meet AITC quality standards for
structural joints (i.e., the type of joints required in glulams). Joints that do
not meet criteria established for structural joints are likely to fail when
subjected to design level stresses.
7.4.7 Glulams for Post-Frame Buildings. A handful of companies now manufacture and
market glulams specifically for use in post-frame buildings. These posts are
intended for soil embedment, with pressure preservative treated wood on one end,
and non-treated wood on the other. Fabrication of such posts requires special
resins and procedures for joining and laminating treated wood to non-treated wood.
7.5 Unspliced MechanicallyLaminated Posts
7.5.1 General. The majority of posts used in post-frame construction with an
overall length less than 18 feet are unspliced, mechanicallylaminated posts. An
unspliced post is any laminated post that does not contain end joints. This means
that each layer is comprised of a single uncut piece of dimension lumber.
7.5.2 Fasteners. As previously noted, a mechanically laminated post is a laminated
post in

which nails, screws, bolts, and/or shear transfer plates (STPs) have been used to
join individual laminates. Nails are the most commonly used mechanical fastener and
posts that only feature nails are often referred to as nail-laminated posts. STPs
are medium-gage metal plates that are stamped such that teeth protrude from both
surfaces.
Mechanical fasteners that connect preservative treated lumber should be AISI type
304 or 316 stainless steel, silicon bronze, copper, hotdipped galvanized (zinc-
coated) steel nails or hot-tumbled galvanized nails.
7.5.3 Advantages. Unspliced mechanicallylaminated posts generally cost less than
solidsawn posts, and they are stronger than similarly sized solid-sawn posts when
bent around axis V-V (figure 7.1a). As previously noted, this is due to the fact
that strength reducing defects are spread out in laminated assemblies. Also,
pressure preservative treatment retention is more uniform in the narrower laminates
of a mechanically-laminated post than it is in wide solid-sawn posts.
7.5.4 Disadvantages. When mechanically-laminated posts are bent around axis H-H
(figure 7.1b), there can be considerable slip between laminates. For this reason,
the bending strength and stiffness of mechanically-laminated assemblies bent about
axis H-H is relatively low. To compensate for this weakness, mechanicallylaminated
posts are generally only used where: (1) there is adequate weak axis support (i.e.,
the posts are part of a sheathed wall), (2) cover plates can be added to increase
bending strength and stiffness about axis H-H (figure 7.2), or (3) the bending
moment about axis H-H is relatively low or non-existent.
Figure 7.2. Cover plates used to increase the bending capacity of a mechanically
laminated post about axis H-H.

7-4

National Frame Builders Association

Post-Frame Building Design Manual

7.5.5 Bending About Axis V-V. Allowable design stresses for bending of unspliced
mechanically-laminated posts about axis V-V are calculated in accordance with
ANSI/ASAE EP559 Design Requirements and Bending Properties for Mechanically
Laminated Columns (ASAE, 1999). The procedure outlined in ANSI/ASAE EP559 is
identical to procedures outlined in the NDS (AF&PA, 1997a) with the exception of
two adjustment factors: the repetitive member factor, Cr, and the beam stability
factor, CL.
7.5.5.1 Repetitive Member Factor. ANSI/ ASAE EP559 allows the use of the repetitive
member factors in Table 7.3 when: (1) each lamination is between 1.5 and 2.0
inches, (2) all laminations have the same depth (face width), (3) faces of adjacent
laminations are in contact, (4) the centroid of each lamination is located on the
centroidal axis of the post (axis V-V in figure 7.1a), that is, no laminations are
offset, (5) all laminations are the same grade and species of lumber, (6)
concentrated loads are distributed to the individual laminations by a load
distributing element, and (7) the mechanical fasteners joining the individual
layers meet the criteria in table 7.4. Note that if one or more of these criteria
are not met, the NDS repetitive member factor of 1.15 should be used if it applies.
7.5.5.2 Beam Stability Factor. The beam stability factor, CL, is a function of the
slenderness ratio, RB, which in turn, is a function of: beam thickness, b; depth,
d; and effective span length, Le. ANSI/ASAE EP559 states that for mechanically-
laminated posts being bent about axis V-V, thickness, b, shall be equated to 60% of
the actual post thickness, and depth, d, to the actual face width of a lamination.
The effective span length, Le, is a function of the unsupported length, Lu. The
unsupported length shall be set equal to the on-center spacing of bracing that
keeps the post from buckling laterally.
7.5.5.3 Design Values. Tables 7.5a and 7.5b contain design values for assemblies
fabricated from visually graded and machine stress rated dimension lumber,
respectively. The design bending stresses have been adjusted for repetitive member
use. They must be further adjusted to account for stability,
wet use, load duration, temperature, and in certain cases, special preservative and
fire treatments.

Table 7.3. Repetitive Member Factors* Number of laminations

34

Visually graded

1.35

1.40

Mechanically graded

1.25

1.30

* For mechanically-laminated dimension lumber assemblies with minimum interlayer


shear capacities as specified in Table 7.4. From ANSI/ASAE EP559 (ASAE, 1999).

Table 7.4. Minimum Required Interlayer

Shear Capacities*

Nominal face width of lamina-


tions, inches

Minimum required interlayer shear capacity


per interface per unit length of post, lb/in.

6 12

8 15

10 19

12 24

* For unspliced mechanically-laminated posts. From ANSI/ASAE EP559 (ASAE, 1999).

7.5.6 Bending About Axis H-H. When all laminates are the same size, species and
grade of lumber, the allowable design bending strength about axis H-H is
conservatively taken as the sum of the bending strengths of the individual layers.
The bending strength of an individual layer is equated to the product of the
flatwise section modulus of an individual laminate and the NDS adjusted design
bending stress. For flatwise bending, the NDS adjusted design bending stress, Fb,
is equal to tabulated design bending stress, Fb, multiplied by the appropriate flat
use factor, a repetitive member factor of 1.15, and all other applicable factors.
Note that the beam stability factor is equal to 1.0 for flatwise bending.

7-5

National Frame Builders Association

Post-Frame Building Design Manual

Table 7.5a Design Values for Unspliced Mechanically-Laminated Posts in Bending


About Axis V-V.

Extreme Fiber Bending Stress*, psi

Grade

Nominal Width of Individual Layers, inches 6 8 10 12


Number of laminations 3. 4. 3. 4. 3. 4. 3. 4.

Modulus of
Elasticity, x 106 psi

Douglas Fir-Larch

Sel Str

2540 2640 2350 2440 2150 2230 1960 2030

1.9

No. 1 & Better 2020 2090 1860 1930 1710 1770 1550 1610

1.8

No. 1

1760 1820 1620 1680 1490 1540 1350 1400

1.7

No. 2

1540 1590 1420 1470 1300 1350 1180 1230

1.6

Hem Fir Sel Str No. 1 & Better No. 1 No. 2

2460 1840 1670 1490

2550 1910 1730 1550

2270 1700 1540 1380

2350 1760 1600 1430

2080 1560 1410 1260

2160 1620 1460 1310

1890 1420 1280 1150

1960 1470 1330 1190

1.6 1.5 1.5 1.3

Southern Pine

Dense Sel Str 3650 3780 3310 3430 2900 3010 2770 2870
1.9

Sel Str

3440 3570 3110 3220 2770 2870 2570 2660

1.8

Non-Dense SS 3170 3290 2840 2940 2500 2590 2360 2450

1.7

Dense No. 1

2360 2450 2230 2310 1960 2030 1820 1890

1.8

No. 1

2230 2310 2030 2100 1760 1820 1690 1750

1.7

Non-Den. No. 1 2030 2100 1820 1890 1620 1680 1550 1610

1.6

Dense No. 2

1960 2030 1790 1960 1620 1680 1550 1610

1.7

No. 2

1690 1750 1620 1690 1420 1470 1320 1370

1.6

Non-Den. No.2 1550 1610 1490 1540 1280 1330 1220 1260

1.4

* For dry posts under normal load duration. Size and repetitive member factors
applied. For other appli-

cable modification factors, see NDS (AF&PA, 1997a).

Table 7.5b Design Values for Unspliced Mechanically-Laminated Posts in Bending


About Axis V-V.

Grade

Extreme Fiber Bending Stress*, psi

3 Laminates
4 Laminates

Grade

Extreme Fiber Bending Stress*, psi

3 Laminates

4 Laminates

900f-1.0E

1130

1170

1950f-1.5E

2440

2540

900f-1.2E

1130

1170

1950f-1.7E

2440

2540

1200f-1.2E

1500

1560

2100f-1.8E

2630

2730

1200f-1.5E

1500

1560

2250f-1.6E

2810

2930

1350f-1.3E
1690

1760

2250f-1.9E

2810

2930

1350f-1.8E

1690

1760

2400f-1.7E

3000

3120

1450f-1.3E

1810

1890

2400f-2.0E

3000

3120

1500f-1.3E

1880

1950

2550f-2.1E

3190

3320

1500f-1.4E

1880

1950

2700f-2.2E

3380

3510
1500f-1.8E

1880

1950

2850f-2.3E

3560

3710

1650f-1.4E

2060

2150

3000f-2.4E

3750

3900

1650f-1.5E

2060

2150

3150f-2.5E

3940

4100

1800f-1.6E

2250

2340

3300f-2.6E

4130

4290

1800f-2.1E

2250

2340

* For dry posts under normal load duration. Repetitive member factors applied. For
other applicable

modification factors, see NDS (AF&PA, 1997a).


7-6

National Frame Builders Association

Post-Frame Building Design Manual

7.5.7 Flexural Rigidity. To calculate deflections due to bending requires that the
flexural rigidity of the member be known. The flexural rigidity of a solid-sawn
member is equal to its modulus of elasticity times its moment of inertia about the
axis it is being bent. The flexural rigidity of an unspliced laminated post when
bent around axis V-V is simply equal to the sum of the flexural rigidities of the
individual laminates about axis VV. In other words, the flexural rigidity about
axis V-V is not dependent on the properties of the mechanical fasteners. This is
not the case with respect to bending about axis H-H. The bending stiffness about
axis H-H axis is highly dependent on the shear stiffness of the mechanical
connections between the individual laminates. A high bound for flexural rigidity
about axis H-H is obtained by assuming complete composite action between layers (no
interlayer slip). A lower bound is obtained by assuming no composite action (no
interlayer connections). In the latter case, the total flexural rigidity is equal
to the sum of the flexural rigidities of the individual laminates. Special analysis
procedures, such as that developed by Bohnhoff (1992) are available for more
accurate estimates of deformation due to bending about axis H-H. Use of these
programs requires knowledge of the shear stiffness properties of the mechanical
connections.
7.5.8 Compressive Properties. The allowable compressive load for an unspliced
mechanically laminated post is typically calculated by treating the individual
laminates as discrete columns. This method conservatively assumes no composite
action between laminates. An allowable compressive stress is first calculated for
each laminate for buckling about axis V-V. This allowable stress is then multiplied
by the crosssectional area of the laminate to obtain an allowable load for buckling
about axis V-V. This calculation is repeated for each layer, and the resulting
individual laminate loads are summed to obtain a total allowable column load for
buckling about axis V-V. The entire process is repeated to obtain a total allowable
load for buckling about axis H-H.
The NDS (AF&PA, 1997a) presents methods for calculating a compressive load capacity
that accounts for some composite action; however, connectors used in fastening the
laminations must meet criteria outlined in the NDS.

7.5.9 Field Fabrication. A distinct advantage of mechanically-laminated posts is


that fabrication can be performed using tools and equipment readily available on
the job site. With unspliced posts that will be embedded in the ground, it is
common to construct the post so that an interior laminate is left shorter than the
surrounding laminates. When the post is installed with this feature located on the
top of the post, the truss can be set in the resulting pocket, enabling a double
shear connection between the post and truss. The interior laminate is generally
significantly shorter (approximately 1 foot) than needed to accommodate the truss.
This is done to compensate for varying depths of embedment. After posts are
installed, a spacer (or block) of the same cross-sectional size as the shortened
laminate is placed in between the shortened laminate and the truss. A schematic of
this procedure is shown in Figure 7.3.

Block Height

Block

1. Post set, bottom of 2. Truss set on

truss marked, and block block and bolted


height measured

into place.

3. Block nailed into place and top of outer layers cut off.

Figure 7.3. On-site truss placement in a mechanically laminated post.

7.6 Spliced Mechanically-Laminated Posts


7.6.1 Types. A spliced post is any post in which at least one laminate contains one
or more endjoints (i.e., is comprised of two or more individual pieces of lumber).
Major end-joint types used in spliced mechanically-laminated posts include: simple
butt joints, reinforced butt joints, and glued finger joints. Butt joints are
generally reinforced by pressing metal plate connectors into one or both sides of
each joint.

7-7

National Frame Builders Association

Post-Frame Building Design Manual

Level line of sight

(a) (b)

(c)

Figure 7.4. (a) Treated portions of 3-layer spliced posts are embedded in the soil.
(b) Top of treated portions cut so that tops at same elevation. (c) Untreated post
portions spliced to treated portions.

7.6.2 Use. Virtually all mechanically-laminated posts with overall lengths


exceeding 20 foot are spliced posts.
7.6.3 Advantages. Splicing enables the fabrication of long posts from shorter, less
expensive lengths of dimension lumber. Splicing also enables the construction of
posts with preservative treated lumber on only one end. This reduces the quantity
of treated lumber used in a building, which in turn reduces the number of special
corrosion-resistant fasteners needed to join treated lumber.
With simple butt joints, the attachment of nontreated lumber to treated lumber is
sometimes done in the field. This attachment is done after the treated pieces have
been laminated and embedded in the ground (figure 7.4a). Prior to attaching the
untreated top-portion of each post, the embedded treated portions are all cut so
that their tops are at the same elevations (note: because of differing depths-of-
embedment, the top of each embedded section is generally at a different height
above grade). With the embedded portions at the same elevation (figure 7.4b),

the upper portions will have the same overall length (figure 7.4c). This eliminates
cutting and blocking like that associated with the special construction shown in
figure 7.3.
7.6.4 Disadvantages. Spliced mechanicallylaminated posts have the same
disadvantages as unspliced mechanically-laminated posts (see Section 7.5.4). In
addition, a simple (nonreinforced) butt joint can significantly reduce bending
strength and stiffness in the vicinity of the joint. If a post contains a simple
butt joint in each laminate, and these joints are all located within 1 or 2 feet of
each other, engineers will often model that portion of the post as a hinge
connection.
7.6.5 Design Properties. Design properties for spliced mechanically-laminated posts
are highly dependent on the type and relative location of end joints, and on the
type and relative location of mechanical fasteners, especially those located in the
vicinity of end joints. Procedures for designing and determining the bending
strength and stiffness of spliced nail-laminated posts are outlined in ANSI/ASAE
EP559 (ASAE, 1999).

7-8

National Frame Builders Association

Post-Frame Building Design Manual

The design portion of EP559 includes requirements for joint arrangement, overall
splice length, nail strength, nail density, nail diameter, and nail location. If
these design requirements are followed, the bending strength and stiffness of the
nail-laminated post can be calculated using the equations in the EP. It is
important to note that the intent of the EP559 design requirements is to maximize
the bending strength of the splice region, while minimizing overall splice length.
Overall splice length is defined as the distance between the two farthest removed
end joints in a post that contains one end joint in each laminate. Reducing overall
splice length generally reduces the amount of preservative treated lumber used in a
post.
7.6.6 Laboratory Tests. Engineers must generally rely on laboratory tests to
determine design properties for spliced posts that do not meet the design
requirements of ANSI/ASAE EP559. In recognition of this, a laboratory test
procedure specifically for spliced mechanically laminated posts is outlined in
ANSI/ASAE EP559.
7.6.7 Computer Modeling. Discontinuities at butt joints result in a post with a
varying bending stiffness along its length. If the overall splice length is rather
short (i.e., all joints are located within a distance equal to 1/4th the post
length), the post is generally sectioned into three elements for computer frame
analysis: a middle element that contains all the joints, and two joint-free outer
elements. The joint-free elements are treated like unspliced mechanicallylaminated
posts with flexural rigidities calculated as described in Section 7.5.7. The
element containing the joints is assigned an effective flexural rigidity that will
cause it to deform like actual laboratory tested posts. A procedure for backing-
out an effective flexural rigidity from bending test data is given in ANSI/ASAE
EP559. The EP also contains an equation for calculating the flexural rigidity of
the splice region of any nail-laminated post that meets the design requirements of
the EP.
7.7 References
American Forest and Paper Association (AF&PA). 1997a. National Design
Specifications for Wood Construction (NDS). American Forest and Paper Association,
Washington, D.C.

American Forest and Paper Association (AF&PA). 1997b. NDS Supplement - Design
values for wood construction. American Forest and Paper Association, Washington,
D.C.
American Institute of Timber Construction (AITC). 1983. Structural glued laminated
timber. ANSI/AITC A190.1-1983. Englewood, CO.
American Institute of Timber Construction (AITC). 1985. Design standard
specifications for structural glued laminated timber of softwood species. AITC
117.85. Englewood, CO.
American Institute of Timber Construction (AITC). 1988. Manufacturing standard
specifications for structural glued laminated timber of softwood species. AITC
117.88. Englewood, CO.
ASAE. 1999. ANSI/ASAE EP559: Design requirements and bending properties for
mechanically laminated columns. ASAE Standards, 46th edition. ASAE, St. Joseph, MI.
Bohnhoff, D.R. 1992. Modeling horizontally naillaminated beams. ASCE Journal of
Strucutral Engineering 118(5):1393-1406.
7-9

National Frame Builders Association

Post-Frame Building Design Manual

7-10

National Frame Builders Association

Post-Frame Building Design Manual

Chapter 8 - POST FOUNDATION DESIGN

8.1 Introduction
8.1.1 General. A distinct advantage of postframe construction is the opportunity to
transfer structural loads to the soil via embedded posts, thereby eliminating the
need for a traditional foundation.
8.1.2 Post Loads. Post loads (i.e., structurally induced shear, bending moment and
axial loads) are obtained using procedures presented in Chapter 5. Most post
foundation design equations require that post loads be specified at the ground
surface.
8.1.3 Post Foundation Classification. Based on their depth, post foundations are
categorized as shallow foundations. Shallow foundations exhibit behavior quite
different from that of deeper systems such as pilings. Specifically, post
deformation below grade is relatively insignificant compared to the deformation of
the soil around the post. Soil deformation around a post is a three-dimensional
phenomena. Figure 8.1 shows the lines of constant soil pressure (in a horizontal
plane of soil) that form when a post moves laterally. The greater the distance
between two posts, the less influence one post will have on the soil pressure near
the other. For design purposes, individual embedded posts are considered isolated
foundations when post spacing is six times greater than post width. Higher
allowable lateral soil bearing pressures are justified for a foundation featuring
isolated posts instead of a continuous foundation wall.
8.1.4 Design Variables. Factors that influence the strength and stiffness of a post
foundation include: embedment depth, post constraint (Section 8.2), soil properties
(Section 8.3), footing size (Section 8.4), collar size (Section 8.5), backfill
properties (Section 8.6), and post dimensions (Section 8.7).
8.1.5 Design Guides. The first design manual for post foundations was originally
published by the American Wood Preservers Institute (Patterson, 1969). The basic
design approach and guidelines for post embedment analysis have been accepted by
several major building codes.

The most comprehensive current design guideline is ASAE EP486 (ASAE, 1999a). The
material in this chapter is largely based on this engineering practice.
CL B/2 B/2

2.0B 1.5B 1.0B


0.5B 1.0B 1.5B 2.0B 2.5B

q
0.9q 0.8q 0.7q 0.6q
0.5q 0.4q

Post 1.0B 1.5B 2.0B


0.1q

0.3q 0.2q
3.0B

3.5B
Figure 8.1. Constant Pressure Lines in a Horizontal Plane of Soil.

8.2 Post Constraint

8.2.1 Nonconstrained Post. The most basic type of post foundation consists of a
post simply embedded in the ground, with no attachments or additional support
(figure 8.2). If the rotation and lateral displacement of the post are resisted
solely by the soil, the post foundation is said to be non-constrained.
8.2.2 Constrained Post. If a post bears on (or is attached to) an additional
immovable structural element such that the lateral displacement at some point at
or above the ground surface is essentially equal to zero, the post foundation is
said to be constrained. An example of a constrained post foundation would be when
the post is installed immediately adjacent to a concrete slab floor in the building
(figure 8.3).

8-1

National Frame Builders Association

Post-Frame Building Design Manual

8.2.3 Varying Constraint. It is important to note that a single post can be both
constrained or non-constrained, depending on the load case. Using the previous
example of a slab floor, and assuming that the post is not attached to the slab, if
the wind loading was such that the post

Ground Level

Ma Va

was pushing on the slab, the post would be considered constrained. However, if the
wind were blowing in the opposite direction, the post would not be supported by the
slab; hence, the post would be analyzed for that load case as nonconstrained.

Ground Level

Ma

Va

do d

Post

Resultant Soil Force

do d

Resultant Soil Force

Rotation Axis

Rotation Axis

Soil Forces
Footing

Footing

LOAD CASE A

LOAD CASE B

Figure 8.2. Free body diagrams of non-constrained post foundations. Load Case A:
groundline shear and moment both cause clockwise rotation of embedded portion of
post. Load Case B: groundline shear and moment cause clockwise and counter
clockwise rotation, respectively, of embedded portion of post.

Ground Level

Ma Va

d
Resultant Soil Force

Floor Post

Soil Forces

Footing

Figure 8.3. Free body diagram of a constrained post foundation.

8-2

National Frame Builders Association

Post-Frame Building Design Manual

8.3 Soil Properties


8.3.1 General. The capability of a soil to handle loads transmitted to it by a post
depend on such characteristics as: particle size and size distribution (a.k.a. soil
classification), moisture content, density, and depth below grade. These soil
characteristics control the allowable vertical and lateral soil pressures.
8.3.2 Soil Classification. Soil is classified by the size of individual particles
and the distribution of sizes within the sample. There are four major particle
(grain) sizes: gravel, sand, silt, and clay. The most popular classification system
in the U.S. (i.e., the Unified Soil Classification (USC) system) classifies gravels
as grains between 0.2 and 3.0 inches, sands as particles between 0.003 and 0.2
inches, silts as grains between 0.003 and 0.00008 inches, and clays as all
particles finer than 0.00008 inches. The distribution of these particles within a
given soil has a major impact of soil behavior. A soil with a wide distribution of
particle sizes is referred to as a well-graded soil. A poorly graded soil is
comprised of similar sized particles. The best soils for foundation design are
gravels and sands, with well-graded gravels and sands, better than poorly graded
gravels and sands. Organic silt, peat and soft clay soils are not suitable for post
foundations, as they have neither the strength nor the stability to support
structural loads.
8.3.3 Soil Moisture Content. The effective shear strength of a soil can be reduced
significantly when soil is allowed to saturate with water. To avoid water
saturation of soils around posts, install rain gutters, and slope the finish grade
away from the building. A minimum 2% slope for a distance of at least 6 ft (2 m)
from the building walls is recommended.
8.3.4 Soil Density and Depth. Allowable vertical and lateral soil pressures
increase with increases in soil density and depth. This is because soil confinement
pressures increase as both of these variables increase.
8.3.5 Tabulated Design Values. Table 8.1

contains soil properties as tabulated in ASAE are referred to as presumptive values


and should only be used if there is no active building code in effect, and site-
specific soil properties are unavailable.
The vertical soil pressures given in table 8.1 are for the first foot (300 mm) of
footing width and first foot below grade. A twenty percent increase in allowable
soil pressure is allowed for each additional foot (300 mm) of foundation width or
depth, up to a maximum of three times the original value.
The lateral soil pressure values in table 8.1 are per unit depth. To obtain the
allowable lateral pressure at a point below grade, SL, multiple the lateral soil
pressure value, S, by the distance below grade of the point in question. For
example, the lateral pressure per unit depth, S, for a firm sandy gravel is 300
lbm/ft2 per foot of depth. This equates to an allowable pressure of 1200 lbf/ft2 (4
ft x 300 lbm/ft2 per ft x 1lbf/lbm) for points four feet below grade. [Note: use of
variable SL to represent S when adjusted for depth, is unique to this design
manual, and is done to avoid confusion between values that have and have not been
adjusted for depth. It is important to realize that SL and S have different units.]
8.3.6 Soil Tests. Site-specific soil test results are often used to determine
allowable soil pressures. Such calculations generally result in higher allowable
design values than would be obtained using table 8.1. This is because presumptive
values are the lowest values associated with a broad classification of soils, each
at their minimum strength conditions.
8.3.7 Soil Sampling. Soil samples should be gathered from the applicable location
in the soil profile: one-third the foundation depth for lateral soil pressure
calculations for non-constrained posts; and at footing depth for lateral soil
pressure calculations for constrained posts and for vertical soil pressure
calculations. From each soil sample, the cohesion, c, angle of internal friction ,
and bulk density, w, must be determined.

8-3

National Frame Builders Association

Post-Frame Building Design Manual

Table 8.1. Presumed Soil Properties for Post Foundation Design (ASAE, 1999). For
use in ab-

sence of codes or test.

Class of Material

Density or Consistency

Lateral Pressure

Per Unit Depth, S

lbf/ft2 kPa per


per ft

Lateral Sliding Coefficient

Vertical Pressure, Sv

lbf/ft2

kPa

Friction Angle, degrees

Density, w lbm/ft2 kg/m3

1. Massive crystalline bedrock

- 1200 180 0.79 4000 200

--

2. Sedimentary and foliated rock

- 400 60 0.35 2000 100 -

--

3. Sandy gravel and/or gravel (GW

firm

300

45

- 38 120 2000

and GP)

loose

200

30

0.35 2000 100

32

90 1500
4. Sand, silty sand, clayey sand,

firm 200

30

silty gravel and clayey gravel (SW, SP, SM, SC, GM, and GC)

loose

150

22.5

0.25 1500

75

30 105 1750 26 85 1400

5. Clay, sandy clay, silty clay and medium

clayey silt (CL, ML, MH and CH)

soft

130 100

20 15

--

1000

50

15 120 2000 10 90 1500

Firm consistency of class 4 and the medium consistency of class 5 can be molded
by strong finger pressure, and the firm consistency of class 3 is too compact to be
excavated with a shovel.
The hydrostatic increase in lateral pressure per unit depth has been included in
the equations of this chapter. Source: Table 29B UBC modified with the addition of
firm and medium values from Hough (1969).
Sliding resistance source: Table 29-B UBC. Allowable foundation pressures are
for footings at least 1 ft (300 mm) wide and 1 ft (300 mm) deep into natural grade.
Pressure
may be increased 20% for each additional 1 ft (300 mm) of width and/or depth to a
maximum of three times the tabulated value. Source: Table 29-B UBC.
Soil friction angle varies from soft to medium density for clay materials, and
from loose to firm for sand and gravel materials. Source: Merritt (1976).
Soil density varies from soft to medium density for clay materials, and from
loose to firm for sand and gravel materials. Source: Hough(1969).
Multiply an assumed lateral sliding resistance of 130 lbf/ft2 (6 kPa) by the
contact area. Use the lesser of the lateral sliding resistance and one-half the
dead load.

8.3.8 Allowable Vertical Soil Pressure From Soil Test Data. The allowable vertical
soil pressure for round or square footings, Sv, can be estimated from site-specific
soil test as:

Sv = SBC / FS where:

(8-1)

Sv =
FS = SBC =

allowable vertical soil pressure, lbf/ft2 (kPa) factor of safety (2.3 to 3.0)
ultimate soil bearing capacity, lbf/ft2 (kPa)

SBC = 0.6 g w b (Nq + 1) tan + (Nq - 1+ Nq tan )(g w y + c/tan) (8-2)


Nq = e tan tan2(/2 + 45)

c = soil cohesion, lbf/ft2 (Pa) = soil angle of internal friction, de-


grees w = soil bulk density, lbm/ft3 (kg/m3) g = gravitational constant, 1 lbf/lbm
(0.00981 kPa m2/kg) y = depth where soil allowable pressure
is calculated, ft (m) b = footing diameter or length of one
side, ft (m)
For shallow foundations, a factor of safety between 2.3 and 3.0 is typically
applied to vertical soil pressure (Whitlow, 1995). Equation 8.2 is a modified
Terzaghi-Meyerhoff equation taken from Whitlow (1995). Values compiled in table 8.2
can be used to facilitate calculation of the ultimate soil bearing capacity, SBC.

8-4

National Frame Builders Association

Post-Frame Building Design Manual

Table 8.2. Ultimate Bearing Capacity*

deg.

SBC = 0.6 w b T1 + Nq T2(w y + c T3)


T1 T2 T3

10 2.471 0.612 1.907 5.671

12 2.974 0.845 2.606 4.705

14 3.586 1.143 3.480 4.011

16 4.335 1.530 4.578 3.487

18 5.258 2.033 5.966 3.078

20 6.399 2.693 7.729 2.747


22 7.821 3.564 9.981 2.475

24 9.603 4.721 12.879 2.246

26 11.854 6.269 16.636 2.050

28 14.720 8.358 21.547 1.881

30 18.401 11.201 28.025 1.732

32 23.177 15.107 36.659 1.600

34 29.440 20.532 48.297 1.483

36 37.752 28.155 64.181 1.376

38 48.933 39.012 86.164 1.280

40 64.195 54.705 117.061 1.192

42 85.374 77.771 161.244 1.111

44 115.308 112.317 225.659 1.036

46 158.502 165.169 321.635 0.966

50 319.057 381.429 698.295 0.839

* See Equation 8.2 for variable descriptions.

8.3.9 Allowable Lateral Soil Pressure From Soil Test Data. The allowable lateral
pressure per foot of depth, S, can be estimated from sitespecific soil test data
as:

S = SRP / FS

(8-3)

where:

S=
FS = SRP =

allowable lateral soil pressure, lbf/ft2 per ft, (kPa per m) factor of safety (1.5
to 2.0) Rankine passive pressure for drained, cohesiveless soils, lbf/ft2 per ft,
(kPa per m).

SRP = w g tan2(45 + /2)

(8-4)

w = soil bulk density, lbm/ft3 (kg/m3) = soil angle of internal friction, de-
grees g = gravitational constant, 1 lbf/lbm
(0.00981 kPa m2/kg)

For lateral earth pressures in drained soils, a factor of safety between 1.5 and
2.0 is typical (Whitlow, 1995). Equation 8-2 assumes drained soils (i.e., the water
table is located below the top of the footing). Equation 8-2 does not account for
soil cohesion, therefore the equation is conservative for clays. Values for the
Rankine passive pressure are given in table 8.3.

Table 8.3. Rankine Passive Soil Pressures

for Drained, Cohesiveless Soils

deg.

SRP, lbf/ft2 per ft Soil Density, lbm/ft3


95 100 105 110 115

120

10 135 142 149 156 163 170

12 145 152 160 168 175 183

14 156 164 172 180 188 197

16 167 176 185 194 203 211

18 180 189 199 208 218 227

20 194 204 214 224 235 245

22 209 220 231 242 253 264

24 225 237 249 261 273 285

26 243 256 269 282 295 307

28 263 277 291 305 319 332

30 285 300 315 330 345 360

32 309 325 342 358 374 391

34 336 354 371 389 407 424

36 366 385 404 424 443 462

38 399 420 441 462 483 504

40 437 460 483 506 529 552

42 479 504 530 555 580 605

44 527 555 583 611 638 666

46 582 613 643 674 704 735

50 717 755 793 830 868 906

8.3.10 Adjustment to Allowable Vertical Pressure. Most codes allow for a 33%
increase in the allowable vertical pressure values, Sv, when post loads result from
wind and seismic forces acting alone or in combination with vertical forces (see
Section 3.9.4). This adjustment would apply directly to the Sv value from equation
8-1, and is cumulative with the adjustments described in Section 8.3.5 for the
presumptive Sv values listed in table 8.1. In this manual, a prime () will be used
to denote an allowable Sv value that has been adjusted (i.e., Sv Sv).
8.3.11 Adjustment to Allowable Lateral Pressure. In addition to the 33% increase
generally allowed when post loads result from wind

8-5

National Frame Builders Association

Post-Frame Building Design Manual

and seismic forces (acting alone or in combination with vertical forces), the
allowable lateral pressure, S, can be doubled when posts have a spacing at least
six times their width. This increase is due to the multi-dimensional nature of
pressure distribution in the soil around isolated posts as depicted in figure 8.1,
and described in Section 8.1.3. In this manual, a prime () will be used to denote
an allowable S value that has been adjusted (i.e., S S).
8.4 Footings
8.4.1 General. Typically, the soil is not able to resist applied vertical loads
when those loads are transferred through the post alone. Therefore, the post is set
on some type of footing, which is installed in the hole prior to post placement.
Footings in post-frame construction are usually poured concrete. This type of
footing is depicted in Figure 8.4. Generally there is no mechanical attachment of
the footing to the post.
8.4.2 Friction. A footing is assumed to only resist vertical loads; the friction
between the footing and the post is assumed to be negligible

when assessing the post lateral load resistance capabilities. Also, the friction
between the post (and/or collar) and the surrounding soil are assumed to be
negligible when assessing the vertical load-carrying capability of a given post
foundation design.
8.5 Collars
8.5.1 General. When lateral soil pressures exceed allowable values, additional
lateral surface area can be obtained by increasing post depth, or by adding a
structural element called a collar. A collar is typically either concrete cast
around the base of the post (and considered to be attached to the post) or built-up
wood attached to the post. These structural elements are represented in figure 8.4.
8.5.2 Location. The collar increases the lateral load resistance capability of the
post foundation by increasing the bearing area in the region of the post where
lateral soil capability is relatively high. Collars are typically not placed at the
top of the post foundation (at the surface of the ground) due to the possibility of
frost heave.

Ground level Post


Original excavated post hole and backfill region
Poured concrete collar Built-up wood collar Footing
(a) (b) Figure 8.4. Examples of common post foundation elements with (a) a poured
concrete collar, and (b) a built-up wood collar.
8-6

National Frame Builders Association

Post-Frame Building Design Manual

8.5.3 Attachment. Whether poured concrete or wood, the collar must be attached to
the post in a manner sufficient to carry the structural loads involved. As with any
wood structural element exposed directly to the soil, appropriate preservatives and
fastener systems must be employed to maintain structural integrity over the design
life of the building.
8.6 Backfilling
8.6.1 General. The details of backfilling are often overlooked by the designer, and
with potentially dire consequences. After the footing and post are installed (and
the collar, if required), the hole that was dug or drilled is backfilled.
Essentially, the material used for backfill is the medium through which some, if
not all, transverse loads are passed from the post to the virgin soil. Backfill
material is subjected to higher pressures than the surrounding virgin soil due to
its proximity to the post. Therefore, material used for backfill and its
installation are critically important for the successful performance of a post
foundation design.
8.6.2 Materials. Typical materials for backfill include concrete, well-graded
granular aggregate, gravel, sand, or soil initially excavated from the post hole.
These alternatives are listed in the order of decreasing stiffness.
8.6.3 Concrete. While concrete is the stiffest backfill material, it is also the
most expensive. Concrete backfill essentially increases post width, b. It must be
installed with attention to the possibility of frost heave (discussed later).
8.6.4 Excavated Soil. The most common backfill material is the excavated soil. If
used as backfill, it should be free of topsoil and organic matter. Silt- or clay-
based soils should be moist (not wet) and well packed.
8.6.5 Compaction. Backfill materials should be tamped or vibrated upon backfill in
maximum layers (a.k.a. lifts) of 8 inch (400 mm).
8.7 Post Dimensions
8.7.1 Effective Width. Design equations for lateral loading (Section 8.7) are a
function of

effective post width, b, which in turn, is a function of post size and shape.

For posts whose narrow face is pushing on the soil:

b = 1.4 B

(8-5)

where:

b = effective post width, ft (m) B = width of post face pushing on the


soil, ft (m)

For posts whose wide face is pushing on the soil, b is equal to the diagonal
dimension of the post.

For poles, the effective post width, b, is equal to the pole diameter.

8.7 Design for Lateral Loadings

8.7.1 General. Bending moments and post shears cause lateral movement of the post
foundation. Designers must insure that this movement does not induce soil stresses
that exceed allowable lateral soil pressures. If the allowable lateral soil
pressure is exceeded, the designer must increase the lateral soil bearing area by
adding a collar, by increasing embedment depth, d, and/or by increasing effective
post width, b. In the majority of cases, the most economical way to increase
bearing area is to increase post depth. For this reason, embedment depth, d, is the
dependent variable in most design equations. Occasionally a designer will add an
extra laminate to the embedded portion of a laminated post to increase effective
width. More often, designers will backfill all or a portion of the hole with
concrete, which is akin to adding a concrete collar.
8.7.2 Assumptions. Equations in 8.7.3 and 8.7.4 assume that only the post (and not
the footing) resists lateral loads. This is because variations in vertical post
loads make it impossible to rely on post-to-footing friction for lateral load
resistance.
8.7.3 Required Embedment Depth for NonConstrained Posts Without Collars. Two
different load cases for a non-constrained, noncollared post are shown in figure
8.2: The first

8-7

National Frame Builders Association

Post-Frame Building Design Manual

load case (a.k.a. Load Case A ) represents conditions where groundline shear and
groundline bending moment both cause the embedded portion of the post to rotate in
the same direction. Load Case B represents conditions where groundline bending
moment causes the embedded portion of the post to rotate in an opposite direction
than the rotation caused by groundline shear. Minimum post embedment depth, d, for
both Load Case A & B is calculated using one of the following equations.

From ASAE EP486 (1999a), AWPI (Patterson, 1969), and the UBC (ICBO, 1994):

d 2=

7.02 Va + 7.65 Ma / d S b

(8-6)

From ASAE EP486.2 (1999b):

d2=

6 Va + 8 Ma / d S b

(8-7)

where:

d= Va =
Ma =
S =
b=

minimum embedment depth, ft (m) shear force applied to foundation at ground


surface, lbf (N) bending moment applied to foundation at ground surface, ft-lbf (N-
m) adjusted allowable lateral soil pressure, lbf/ft per ft (kPa/m per m) effective
post width, ft (m), see Section 8.7.1

Equations 8-6 and 8-7 must be solved by iteration. For Load Case B, Va and Ma must
be input with opposing signs. Note that equation 8-6 is in a slightly different
form than appears in any of the three referenced documents. See the following
technical note on equation development for additional information.

Technical Note Non-Constrained Post Equations


Equation 8-6 for the embedment depth, d, of non-constrained, non-collared posts
appears in most code documents as:
d = 0.5 A [1 + (1 + 4.36 h / A)1/2] (T-1)

where:

A= b= d= P= h=
SL =
2.34 P / (SL b) effective post width, ft (m) post embedment depth, ft (m) applied
lateral force, lbf (N) distance from ground surface to point of application of
force P, ft (m) adjusted allowable lateral soil pressure at one-third the embedment
depth, lbf/ft2 (kPa)

Equation T-1 was developed for point-loaded posts that behave as pure cantilevered
beams. Unfortunately, posts in post-frame buildings are not point-loaded, and
embedded posts are supported in such a way that they behave more like propped
cantilevers.

To adjust equation T-1 so that it can be applied to posts subjected to a variety of


loadings and above-grade constraint conditions, load P is replaced with an
equivalent shear force and bending moment located at the ground surface. Using
predefined nomenclature: Va is substituted into equation T-1 for P, and Ma is
substituted for the product of P and h. In addition, the adjusted allowable lateral
soil pressure at onethird the embedment depth, SL, is replaced by the quantity S
d / 3. This substitution eliminates having to recalculate SL every time the
embedment depth changes. With these substitutions, equation T-1 appears in ASAE
EP486 (1999a) as:
d 2 = 3.51Va/(S b)[1+(1+(0.62 Ma S b d)/ Va2)1/2]

Because it is somewhat confusing, the ASAE EP486 equation was rewritten for this
design manual in the form of Equation 8-6.

The first major revision to ASAE EP486 (due for release in 2000) will contain
several changes, including a switch from equation 8-6 to equation 8-7. Equation 8-7
is based on five common assumptions: (1) only the post (and not the footing)
resists lateral loads, (2) the post behaves as a rigid body below grade (3) soil
type remains constant, (4) at a given depth, soil resisting pressure, q, is equal
to the product of soil stiffness, k, and lateral post movement at that depth, and
(5) soil stiffness, k, at a distance, y,

8-8

National Frame Builders Association

Post-Frame Building Design Manual

below grade is equal to the product of the horizontal subgrade reaction, nh, and
the distance below grade, y. In equation form, the soil resisting pressure, q, for
a non-constrained, noncollared post is given by Meador (1997) as:

q = nh (y y2/do)

(T-2)

where:

q=
nh =
=
y= do =

actual soil pressure at a depth y below grade, lbf/ft2 (kPa) constant of horizontal
subgrade reaction, lbf/ft4 (N/m4)
lateral post deflection at grade, ft (m) depth below grade, ft (m) distance from
surface to point of post rotation in soil, ft (m), (see figure 8.2)
Equation T-2 is a parabolic function that produces the soil pressure profile shown
in figure 8.2. If a summation of the horizontal forces in figure 8.2 is set equal
to zero, and the bending moment around any point is equated to zero, the following
two equations can be obtained for the grade deflection , and distance to post
rotation point, do.

= (24 Ma + 18 Va d)/(d 3 nh b)

(T-3)

do = (3 Va d + 4 Ma)/(4 Va + 6 Ma /d) (T-4)


Examination of equation T-4 shows that the point of post rotation is two-thirds the
embedment depth when there is no shear in the post at the ground surface (Va = 0).
When there is no moment in the post at the ground surface (Ma = 0), the point of
post rotation is located at threequarters of the embedment depth. If both Va and Ma
are positive and non-zero, the point of rotation is between two-thirds and three-
fourths of the embedment depth.

Substitution of equation T-3 into equation T-2 yields the following equation for
soil pressure:
q = (18 Va + 24 Ma/d)(y y 2/do)/(d 2 b) (T-5)

Typically, a designer selects a value for d, such that for all points below the
surface, the actual

soil pressure, q, does not exceed the adjusted allowable lateral soil pressure, SL
= S y. It can be shown that every time a designer does this, the depth at which
the actual soil pressure is closest to the allowable pressure is right at the
surface. In other words, for a non-constrained post, the designer does not need to
compare S y and q from equation T-5 at every value of y, instead, the designer
only needs to check it at y = 0. It follows that the embedment depth, d, needed to
ensure that the actual soil pressure does not exceed the allowable soil pressure at
the surface (or any point below the surface) is given as:

d 2 = (18 Va + 24 Ma / d )/(S b)

(T-6)

Equation T-6 is not used in practice as field and laboratory tests have shown that
it is extremely conservative for non-constrained posts. This is because when actual
soil pressures at the surface equal the allowable soil pressure, the actual soil
pressure at points below the surface are below (and in most cases substantially
below) allowable soil pressures. Consequently, nonconstrained post foundations are
no where near failure when allowable soil pressures near the surface are exceeded.
A more realistic embedment depth is obtained by replacing S in equation T-6 with
3S. The resulting equation is equation 8-7. Note that when this equation is used,
actual soil pressure will exceed allowable soil pressure for points between y = 0
and y = 2do/3, and for points deeper than y = 4do/3.

For an in-depth discussion and greater detail on non-constrained post foundation


equation development see Meador (1997).

8.7.4 Required Embedment Depth for Constrained Posts Without Collars. A free body
diagram of a constrained, non-collared post is shown in figure 8.3. Minimum post
embedment depth, d, for the constrained, non-collared case is calculated using one
of the following equations.

From ASAE EP486 (1999a), AWPI (Patterson, 1969), and the UBC (ICBO, 1994):
d=

4.25 Ma S b

1/3

(8-8)

8-9

National Frame Builders Association

Post-Frame Building Design Manual

From ASAE EP486.2 (1999b):

d=

4 Ma S b

1/3

(8-9)

where:

d= Ma =
S =
b=

minimum embedment depth, ft (m) bending moment applied to foundation at ground


surface, ft-lbf (N-m) adjusted allowable lateral soil pressure, lbf/ft per ft
(kPa/m per m) effective post width, ft (m), see Section 8.7.1

Note that equation 8-8 is in a slightly different form than appears in any of the
three referenced documents. See the following technical note on equation
development for additional information.

Technical Note Constrained Post Equations

Equation 8-8 for the embedment depth, d, of constrained, non-collared posts appears
in most code documents as:

d = [4.25 P h / (SL b)]1/2

(T-1)

where:

d= P= h=
SL =
b=

post embedment depth, ft (m) applied lateral force, lbf (N) distance from ground
surface to point of application of force P, ft (m) adjusted allowable lateral soil
pressure at the full embedment depth, lbf/ft2 (kPa) effective post width, ft (m)

Equation 8-8 is derived from equation T-1 by


substituting bending moment, Ma, for the product of P and h, and replacing SL with
the quantity Sd. This latter substitution eliminates having to
recalculate SL every time the embedment depth changes.

As described in the previous technical note on non-constrained posts, the first


major revision to ASAE EP486 will contain several changes. One

of these is a switch from equation 8-8 to equation 8-9. Equation 8-9 is based on
the same assumptions as described for equation 8-7. These assumptions result in the
following equation for actual soil resisting pressure, q, for a constrained, non-
collared post (Meador, 1997):

q = nh y2 / d

(T-2)

where:

q= nh = = y=

actual soil pressure at a depth y below grade, lbf/ft2 (kPa) constant of horizontal
subgrade reaction, lbf/ft4 (N/m4)
lateral movement of post at a depth y = d, ft (m) depth below grade, ft (m)

Equation T-2 is a parabolic function that produces the soil pressure profile shown
in figure 8.3. If the bending moment around any point in figure 8.3 is equated to
zero, the following equation is obtained for the lateral movement, , of the post
at a depth, d.

= 4 Ma /(d 3 nh b)

(T-3)

Substitution of equation T-3 into equation T-2 yields the following equation for
soil pressure:

q = 4 Ma y 2/(d 4 b)

(T-4)

The actual soil pressure increases at an increasing rate as y increases. The


allowable lateral soil pressure, SL, increases at a constant rate as y increases
(note: SL = S y). This means that if a designer ensures that the actual soil
pressure, q does not exceed the allowable pressure at a depth, y = d, then the
actual stress will be less than the allowable for all points between the ground
surface and y = d. In equation form:

S y > q = 4 Ma y 2/(d 4 b) for y = d (T-5)

Equation T-5 becomes equation 8.9 after it is rearranged so that d is the dependent
variable.

For an in-depth discussion and greater detail on constrained post foundation


equation development see Meador (1997).

8 - 10

National Frame Builders Association

Post-Frame Building Design Manual


8.7.5 Required Embedment Depth for Posts With Collars. This design manual does not
contain embedment equations for posts with collars. For such equations, see ASAE
EP486 (1999a, 1999b) and Meador (1997).
8.8 Design for Downward Loadings

8.8.1 Required Footing Area. Downward forces are resisted by the footing. The
footing area, A, required to resist these forces is:

A = P / SV

(8-10)

where:

A= P= SV =

required footing area, ft2 (m2) vertical foundation load, lbf (N) adjusted
allowable vertical soil pressure, lbf/ft2 (kPa) (see Section 8.3.10)

8.9 Design for Uplift Loadings

8.9.1 General. If the net vertical force acting on a post is upward, either the
footing or a collar must be attached to the post. When the footing or a collar is
attached to the post, upward movement of the post foundation cannot occur without
displacing a cone-shaped mass of soil. The mass of this of soil depends on
foundation depth, footing (or collar) size, soil density, and soil internal
friction angle.
8.9.2 Skin Friction. An attached footing or collar is required to resist uplift
forces because skin friction between a post and backfill cannot be relied on to
resist such forces.
8.9.3 Concrete Backfill. Concrete cast against undisturbed soil and mechanically
fastened to the post adds uplift resistance of both the concrete mass and the skin
friction between the concrete and soil. Note that this practice is not recommended
in soils with a high susceptibility to frost heave
8.9.4 Volume of Displaced Soil. The volume of soil that must be displaced when
pushed upward by a footing or collar is dependent on the shape of the footing or
collar. Figures 8.5 and 8.6 show configurations for circular and rectangular
foundation elements, respectively.

Ground Level
Post
Ap dT

r / tan

Collar Footing

2r

Figure 8.5. Schematic of relevant uplift resistance components for post foundation
with an attached circular collar.

Ground Level
Post
AP dT

Collar
l1 l2
Unattached Footing
Figure 8.6. Schematic of relevant uplift resistance components for post foundation
with an attached rectangular collar.

8 - 11

National Frame Builders Association

Post-Frame Building Design Manual

The volume of displaced soil, VS, is calculated using the following equations:

For circular footings and collars:

VS = dT [r 2 + dT r tan + dT 2 tan2 / 3] dT Ap

For rectangular footings and collars:

VS = dT (l1 l2 - Ap) + dT2 tan (l1 + l2) + dT 3 tan2 / 3

where: VS = dT =
r=
= Ap = l1, l2 =

volume of displaced soil, ft3 (m3) distance from ground surface to top of collar,
or to top of footing if collar is not present, ft (m) radius of collar, or footing
if collar is not present, ft (m)
angle of internal soil friction post cross-sectional area, ft2 (m2) length and
width of a rectangular collar or footing, ft (m)

8.9.5 Uplift Resistance, U. The resistance to uplift, U, is calculated as:

U = g ( MF+ w VS ) where:

(8-11)

U= MF =
w= dT =
VS = g=

uplift resisting force, lbf (N) mass of all foundation elements that are attached
to the post, lbm (kg) soil density, lbm/ft3 (kg/m3) distance from ground surface to
top of collar, or to top of footing if collar is not present, ft (m) volume of
displaced soil, ft3 (m3) gravitational constant, 1 lbf/lbm (9.81 N/kg)

8.10 Frost Heave Considerations

8.10.1 General. Freezing temperatures in the soil result in the formation of ice
lenses in the spaces (a.k.a. pores) between soil particles. Under the right
conditions, these ice lenses will continue to attract water and increase in size.
This expansion of the ice lenses increases soil volume. If this expansion occurs
under a footing, or alongside a foundation element with a rough surface, that
portion of the foundation will be

forced upward.
8.10.2 Problems. Frost heave can induce large differential movements in the
foundation. This differential movement can crack building finishes, and induce
significant stress in structural connections and components. When ice lenses thaw,
soil moisture content increases dramatically. The soil is generally in a saturated
state with reduced strength. As soil water drains from the soil, effective soil
stresses increase and the foundation will generally settle.
8.10.3 Minimizing Frost Heave. Frost heave can be minimized by: (1) avoiding clays
and silts, (2) extending footings below the frost line, and (3) providing good
drainage.
8.10.3.1 No Silts and Clays. Fine grained soils such as clays and silts are more
susceptible to frost heave because (1) water is drawn upward by the fine pores
which function as capillaries, and (2) there is much more surface area in a unit
volume of fine grained soil, and therefore more surface area for water adsoprtion.
8.10.3.2 Footing Depth. The most sure fire way to avoid frost heave problems is to
locate the footing where water never freeze. It is for this reason that codes
require foundations to be located below the frost line. Exceptions include footings
on rock and floating foundation systems. A floating foundation is reinforced so
that it can float as a monolithic unit as the soil swells and shrinks.
8.10.3.3 Water Drainage. Proper surface and subsurface drainage can reduce frost
heave. Drainage of surface waters from a builder is enhanced by installing rain
gutters, adequately sloping the finish grade away from the building, and raising
the building elevation to a level above that of the surrounding area. Subsurface
drainage is achieved with the placement of drain tile or coarse granular material
below the maximum frost depth, with drainage to an outlet. Such drainage lowers the
water table and interrupts the flow of water moving both vertically and
horizontally through the soil.

8 - 12

National Frame Builders Association

Post-Frame Building Design Manual

8.10.4 Concrete Floors. If the ground beneath a concrete floor can freeze, the
floor should be installed such that its vertical movement is not restricted by
embedded posts or by structural elements attached to embedded posts. While concrete
shrinkage may break bonds between a floor and surrounding components, more
proactive measures will ensure independent vertical behavior. For example, plastic
film can be placed against surrounding surfaces prior to pouring the floor.
8.10.5 Concrete Backfill. The use of poured concrete as a backfill material may
actually increase the likelihood of frost heave. The rough soil-to-concrete
backfill interface provides the potential for significant vertical uplift forces
due to frost heave. Also, the placement of concrete in holes that decrease in
diameter with depth provide additional risk for frost heave.
8.10.6 Top Collars. Although common in past years, placement of collars at the
ground surface (to increase lateral load resistance) has all but been abandoned due
to frost heave considerations.
8.11 References
ASAE. 1999a. ASAE EP486: Post and pole foundation design. Shallow post foundation
design. ASAE Standards, 46th Edition. ASAE,. St. Joseph, MI
ASAE. 1999b. ASAE EP486.2: Shallow post foundation design. In review. ASAE. St.
Joseph, MI.
Hough, B.K. 1969. Basic Soils Engineering, 2nd Edition. Ronald Press Co. Table 7-2,
p. 249.
International Conference of Building Officials (ICBO). 1994. Uniform Building Code,
1994 Edition. ICBO, Whittier, CA
McGuire, P. M. 1998. Overlooked assumption in nonconstrained post embedment. ASCE
Practice Periodic on Structural design and Construction, 3(1):19-24.
Meador, N.F. 1997. Mathematical models for lateral resistance of post foundations.
Trans of ASAE, 40(1):191-201.

Merritt, F.S. 1976. Standard Handbook for Civil Engineers, pp. 7-53.
Patterson, D. 1969. Pole Building Design. American Wood Preservers Institute
(AWPI), Washington D.C.
Whitlow, R. 1995. Basic Soil Mechanics. 3rd edition. John Wiley & Sons, Inc. New
York, NY

8 - 13

National Frame Builders Association

Post-Frame Building Design Manual

8 - 14

National Frame Builders Association

Post-Frame Building Design Manual

Chapter 9: DESIGN EXAMPLE

9.1 Introduction

9.1.1 General. Structurally efficient post-frame buildings utilize the roof as a


diaphragm to resist horizontal wind forces. This chapter presents an example of
diaphragm design following the five steps outlined in Section 5.1.4.

9.1.2 Building Specifications. Table 9.1 lists design parameters for the example
building.

Table 9.1. Example Building Specifications

Width (truss length)

36 ft

Length (along ridge)

60 ft

Height at post bearing

12 ft

Roof slope

4/12 (18.43)

Bay spacing

10 ft

Number of frames (including end walls)

Post embedment depth

4 ft

Post grade & species


No.2 S. Pine

Post size

Nom. 6- by 6-in.

Roof snow load

30 psf

Roof dead load

5 psf

Concrete slab?

Yes

Ceiling?

No

9.1.3 Wind Loads. It is assumed that the example building is located in a


jurisdiction that has adopted the 1994 Uniform Building Code. Design wind loads
calculated according to this code are presented in Table 9.2

Table 9.2. Wind Loads

Wind speed

80 mph

Exposure category

Windward wall, qww

8.13 psf

Leeward wall, qlw

-5.08 psf *

Windward roof, qwr

3.05 psf

Leeward roof, qlr

-7.12 psf *

* Negative loads act away from the surface in

question. Positive loads act toward the sur-

face in question.

9.2 Step 1: Modeling


9.2.1 General. The structural model for this example building follows that in
Section 5.2. The frames are numbered from one to seven beginning on the left end.
That portion of the roof diaphragm between each frame is broken into two discrete
segments labeled 1a, 1,b, 6a, 6b. See Figure 9.1.
123 456 7
1a 2a 3a 4a 5a 6a
1b 2b 3b 4b 5b 6b

Figure 9.1. Identification of frame elements and roof diaphragm segments.

9.3 Step 2: Stiffness Properties

9.3.1 Frame Stiffness, k. One reliable way to determine frame stiffness is to use a
planeframe analysis program such as the PPSA program mentioned in Section 5.3.2. In
this example, all posts will be considered fixed at the grade line and pin
connected to trusses (figure 5.5). Consequently, the stiffness of each embedded
post can be calculated using equation 53 which is given as:

kp = 3 E I / Hp3

For the nominal 6- by 6-inch No. 2 Southern Pine posts:

E=
I= Hp =

1.2 x 106 lbf/in.2 (No adjustment for wet conditions is necessary for Southern Pine
timbers. It is generally required for laminated posts.) 76.26 in.4 144 in.

Thus, kp = 91.9 lbf/in.

9-1

National Frame Builders Association

Post-Frame Building Design Manual

Frame stiffness, k, is obtained by summing individual post stiffness values


(equation 5-2). This summation yields:

k = 184 lbf/in.

9.3.2 Diaphragm Stiffness, Ch. The diaphragm assembly used in this example is Test
Assembly 11 in Table 6.1. Its properties are summarized in Table 9.3.

Table 9.3. Diaphragm Properties

Metal thickness

29 gage

Assembly width, 3 x a

36 ft

Assembly length, b

12 ft

Allowable shear strength, va Effective in-plane shear stiffness, c


107 lbf/ft 3700 lbf/in.

Effective shear modulus, G

3700 lbf/in.

In-plane shear stiffness for a single diaphragm section is calculated using


equation 6-9, which is given as.

cp =

G bh s cos()

Substitution of appropriate values yields:

cp =

(3700 lbf/in.)(18 ft) (10 ft)(cos(18.43))

cp = 7020 lbf/in.
The horizontal shear stiffness, ch, of a single diaphragm section is calculated
using equation 6-10 which is given as:
ch = cp cos2 ()
Substitution of appropriate values yields:
ch = (7020 lbf/in.) cos2(18.43) = 6320 lbf/in.
Total horizontal shear stiffness of a diaphragm element, Ch, is found by summing
the stiffness values of the two sections that comprise each diaphragm element (see
equation 5-4).
Ch = 6320 + 6320 = 12,640 lbf/in.

9.3.3 Shearwall Stiffness, ke. There are no large doors in the endwalls of the
example building. Lacking a specific tested endwall assembly, the 12 ft high
endwalls will be assumed to have the same shear stiffness as an 8 ft section of the
roof diaphragm; that is, ke will be set equal to Ch or 12,640 lbf/in.

9.4 Step 3: Eave Loads

9.4.1 Windward Roof Pressures. As noted in Section 9.1.3, this example uses wind
loads calculated in accordance with the 1994 UBC. Pressure coefficients (from UBC
table 16-H) for windward roof slopes between 2/12 and 9/12 are -0.9 (outward) and
0.3 (inward). It is important to recognize the significant impact that wind
direction (inward or outward) has on calculated eave loads. The 3.05 psf design
windward roof pressure listed in table 9.2 was calculated using the 0.3 pressure
coefficient. When combined with the negative pressure of 7.12 psf on the leeward
roof, the net lateral roof pressure is 10.17 psf. If the 0.9 pressure coefficient
would have been used, the net lateral roof pressure would have been 2.03 psf.

9.4.2 Fixity Factors, f. Based on the assumption of a post fixed at the groundline
(see Section 9.3.1), a fixity factor of 3/8 is appropriate for this example.

9.4.3 Eave Load, R. Since this example uses symmetrical base restraint and frame
geometry, equation 5-6 may be used.

R = s [hr (qwr qlr) + hw f (qww qlw)]

where:

hr = hw = s=
f=

(36 ft /2) (4/12) = 6 ft 12 ft 10 ft 0.375

or

R = 10 ft [6 ft (3.05 psf + 7.12 psf) + 12 ft (.375)(8.13 psf + 5.08 psf)]

R = 1205 lbf

9-2

National Frame Builders Association

Post-Frame Building Design Manual

For later calculations, it is convenient to calculate R in terms of its components


roof, windward wall and leeward wall.
RR = 10(6)(3.05 + 7.12) = 610.2 lbf RW = 10(12)(.375)(8.13) = 365.8 lbf RL = 10(12)
(.375)(-5.08) = -228.6 lbf RR + RW RL = 1205 lbf
9.5 Step 4: Load Distribution
9.5.1 Introduction. For this example problem, diaphragm shear stiffness, Ch, frame
stiffness, k, endwall stiffness, ke, and eave load, R, are all constant.
Consequently, in addition to analysis methods such as DAFI, load distribution can
be determined using the ANSI/ASAE EP484.2 tables (Section 5.6.3) and the simple
beam analogy equations (Section 5.5.6). For comparison purposes, all three methods
are demonstrated here (Sections 9.5.2 9.5.4). The information obtained from these
analyses is then used to determine the maximum diaphragm shear force (Section
9.5.6) and maximum post forces (Section 9.5.7).
9.5.2 ANSI/ASAE EP484.2 Tables. In this design manual, the ANSI/ASAE EP484.2 tables
are tables 5.1 and 5.2. Table 5.1 contains shear force modifiers or mS values.
The product of this modifier and eave load, R, is the maximum shear force in the
diaphragm, Vh. Table 5.2 contains sidesway restraining force modifiers or mD
values. The product of this modifier and eave load, R, is referred to as the
horizontal restraining force, Q, which is the amount of eave load transferred away
from the center postframe(s) by the diaphragm.
Use of tables 5.1 and 5.2 requires two ratio: ke/k and Ch/k. For this example
analysis, both ratios are equal to 69 (12640/184). Using linear interpolation, the
mS value from table 5.1 is 2.77, and the mD value from table 5.2 is 0.90.
The maximum diaphragm shear force, Vh, which occurs adjacent to each endwall, is
given as:
Vh = mS R = 2.77(1205 lbf) = 3340 lbf
The horizontal restraining force, Q, that must be applied to the center post frame
(i.e., frame 4 in figure 9.1) is given as:

Q = mD R = 0.90(1205 lbf) = 1085 lbf


The difference between eave load, R, and the horizontal restraining force, Q, is
the amount of the eave load that is transferred by the center post-frame to the
foundation.
R Q = 120 lbf
The eave deflection, , for a post-frame with stiffness, k, subjected to an eave
load, R, and horizontal restraining force, Q, is given as:
= (R Q) / k
Eave deflection for the center post-frame is given as:
= (1205 lbf 1085 lbf) / 184 lbf/in.
= 0.652 in.
9.5.3 Simple Beam Analogy Equations. As previously noted, the simple beam analogy
equations for diaphragm shear force, Vh, and diaphragm displacement, y, can be used
when R, k, ke and Ch are constant. These two equations are given in Section 5.6.6
as:
Vh = Ch s [A sinh( x) + B cosh( x)]
y = A cosh( x) + B sinh( x) + R/k
Input parameters and calculated equation constants for the simple beam analogy
equations have been compiled for this example analysis in Table 9.4.
Maximum diaphragm shear is calculated by setting x = 0 in., or:
Vh = 12,640 lbf/in.(1.0054x10-3 in.-1) (120 in.)[-6.286 in.(0) + 2.181 in.(1)]
Vh = 3326 lbf
Maximum diaphragm displacement is calculated by setting x = L/2 = 360 in. , or:
y = -6.286 in.( 1.0662) + 2.181 in.( 0.3699) + 1205 lbf/(184 lbf/in.)
y = 0.6535 in.

9-3

National Frame Builders Association

Post-Frame Building Design Manual

Table 9.4. Parameters for Simple Beam Analogy Equations

R 1205 lbf

s 120 in.

L 720 in.

ke 12,640 lbf/in.

k 184 lbf/in.

R/ k

6.549 in.

Ch 12,640 lbf/in. 1.0054x10-3 in.-1 *

L 0.7239

cosh( L)

1.2737

sinh( L)

0.7888

D -23.890 *

ye 0.2631 in. *

A -6.286 in. *

B 2.181 in. *

cosh(0)

sinh(0)
0

cosh( 360 in.)

1.0662

sinh( 360 in.)

0.3699

* Equations for calculation of these values are given in Section 5.6.6.

The force transferred to the foundation by the center frame (frame 4) is equal to
the product of eave displacement, y, and frame stiffness, k, or:
y k = 0.6535 in. (184 lbf/in.) = 120.2 lbf
The horizontal restraining force, Q, for the frame 4 is equal to the difference
between the eave load, R, and the 120.2 lbf, or
Q = 1205 lbf 120.2 lbf = 1084.8 lbf
Note that ye in table 9.4 is the eave displacement of the endwall.
9.5.4 DAFI. As previously mentioned, DAFI is a computer program specifically
written for determining load distribution between diaphragms and frames. DAFI can
be downloaded free from the NFBA web site (www.postframe.org).
The maximum shear force in the diaphragm. Vh, is numerically equal to the load
resisted by the endwall frame. In figure 9.2, this value is given as 3353.2 lbf.
Note that this value is more precise than the 3340 lbf value calculated from the mS
values in table 5.1 because the values in table 5.1 are only given to three
significant figures. It is important to note that the shear load

FRAME FRAME

APPLIED HORIZONTAL LOAD RESISTED FRACTION OF

NUMBER STIFFNESS LOAD DISPLACEMENT BY FRAME APPLIED LOAD

---------------------------------------------------------------------

1 12640.00 602.5 .2652868

3353.2

5.5655

2 184.00 1205.0 .4829074

88.9

.0737

3 184.00 1205.0 .6122254

112.6

.0935

4 184.00 1205.0 .6551232

120.5
.1000

5 184.00 1205.0 .6122254

112.6

.0935

6 184.00 1205.0 .4829074

88.9

.0737

7 12640.00 602.5 .2652867

3353.2

5.5655

DIAPHRAGM DIAPHRAGM

SHEAR

SHEAR

NUMBER

STIFFNESS DISPLACEMENT LOAD

--------------------------------------------

12640.00

.2176206 2750.7

12640.00

.1293180 1634.6

12640.00

.0428978

542.2

12640.00

.0428979

542.2
5

12640.00

.1293180 1634.6

12640.00

.2176206 2750.7

Figure 9.2. Output from computer program DAFI for example building.

9-4

National Frame Builders Association

Post-Frame Building Design Manual

listed for each diaphragm in the DAFI output is essentially an average shear load
in the diaphragm. For example, the average shear load in diaphragm 1 is listed as
2750.7 lbf. To calculate the maximum shear load in each diaphragm element, simply
add the quantity R/2 to the average value. For this example analysis, half the eave
load is 602.5 lbf. Adding this to the average shear load in diaphragm 1 yields the
expected maximum shear force in the diaphragm of 3353.2 lbf.
The amount of eave load transferred to the foundation by each frame is listed in
figure 9.2 under the column heading load resisted by frame. The difference
between this value and the eave load, R, is the horizontal restraining force, Q.
The load resisted by the most heavily loaded frame (i.e., frame 4) is 120.5 lbf.
This equates to a horizontal restraining force for frame 4 of 1084.5 lbf (1205 lbf
120.5 lbf).
9.5.5 Comparison of Methods. The ANSI/ ASAE EP484.2 tables (tables 5-1 and 5-2),
simple beam analogy equations, and program DAFI yield identical values for maximum
diaphragm shear, horizontal restraining force, and eave deflections. Again, it is
important to note that the ANSI/ASAE EP484.2 tables and the simple beam analogy
equations are restricted to designs with fixed values of Ch, k, R, and ke. Although
DAFI is more versatile, a DAFI analysis requires computer access. The simple beam
analog equations can be quickly solved with a hand calculator that supports
hyperbolic trigonometric functions.
9.5.6 Diaphragm Shear. The maximum inplane shear force, Vp, in a diaphragm section
is calculated from the maximum horizontal shear force, Vh, in the diaphragm
elements using equation 5-9 which is given as:
Vp,i = (ch,i / Ch) Vh / (cos i)
For this example analysis, all six diaphragm elements have the same Ch, and all
twelve of the diaphragm sections shown in figure 9.1 have the same horizontal
stiffness, ch and slope, , that is:
Ch = 12,640 lbf/in. ch,i = 6320 lbf/in. = 18.43

Diaphragm elements 1 and 6 are both subjected to the maximum horizontal shear, Vh,
of 3350 lbf. Consequently, the in-plane shear force in diaphagm sections 1a, 1b, 6a
and 6b is given as:

Vp =

6320 lbf/in (3350 lbf) 12,640 lbf/in (cos 18.43)

Vp = 1766 lbf
Dividing the in-plane shear force by the slope length of a diaphragm section yields
the in-plane shear force on a unit length basis, vp.

vp = 1766 lbf /(18 ft / cos (18.43)) vp = 93.1 lbf/ft

9.5.7. Post Forces. The most critical posts from a design perspective are those
associated with the most heavily loaded frame. In the example building this is the
center post-frame (a.k.a. frame 4).

There are two basic methods for determining post forces. The first is to analyze
the frame with a plane-frame structural analysis program, the second is to assume
the truss is rigid and then use a series of equations to calculate post forces.

A structural analog for a plane-frame structural analysis of frame 4 is shown in


figure 9.3a. Post forces obtained with this analog are given in figure 9.3b. For
this example analysis, the load combination of full dead + full wind + snow
was used, with a roof dead load of 5 psf and a roof snow load of 30 psf (Note: in
practice, the building designer must check all applicable load cases). The force
applied to the frame by the diaphragm, qp, was applied as a force of 30.12 lbf per
foot of top chord. This force was obtained by first combining equations 5-10 and 5-
11 into the following equation:

q p,i = Q (c h,i / Ch ) / b i

(9-1)

where: Q is the horizontal restraining force


(1084.5 lbf for frame 4); ch,i is the horizontal stiffness of diaphragm segment i
(6320 lbf/in); Ch is the horizontal stiffness of diaphragm element i
(12,640 lbf/in); and bi is the horizontal span of diaphragm segment i (18 ft).

9-5

National Frame Builders Association

Post-Frame Building Design Manual

s x (5 psf + 30 psf /2)

3820 lbf

2650 lbf

s x 3.05 psf 30.12 lbf/ft

s x 7.12 psf 30.12 lbf/ft

0 in.-lbf 662 lbf

313 lbf

0 in.-lbf

Windward post

Leeward post

449 lbf
25100 in.-lbf

161 lbf 20700 in.-lbf

3821 lbf

2646 lbf

(a) (b)
Figure 9.3. (a) Structural analog for frame 4 of the example building (s = 10 ft).
(b) Resulting forces on post ends. Lateral deflection at the top of the windward
and leeward posts were 0.572 and 0.735 inches, respectively.

s x 8.13 psf s x 5.08 psf

Roof dead + 1/2 snow = 7200 lbf

Vertical component of windward roof pressure = 549 lbf


+
Vertical component of diaphragm restraining force = 180.75 lbf

Horizontal component of windward roof pressure = 183 lbf


+
Horizontal component of diaphragm restraining force = -542.25 lbf

3 ft 3 ft Vtw

9 ft

9 ft

Vertical component of leeward roof pressure = -1281.6 lbf


+
Vertical component of diaphragm restraining force = -180.75 lbf

9 ft 9 ft

Horizontal component of leeward roof 3 ft pressure = -427.2 lbf


+ 3 ft Horizontal component of diaphragm
restraining force = 542.25 lbf Vtl

Pw = 3821 lbf

Pl = 2646 lbf

Figure 9.4. Resultant of forces applied to truss of frame 4.

In lieu of a computer analysis, post axial forces for a two-post frame can be
obtained by drawing a free-body diagram of the truss and summing forces about each
truss-to-post connection. Such a free-body diagram for frame 4 of the example
building is shown in figure 9.4. The axial forces (Pw and Pl) obtained in this
manner are identical to those obtained via the computer analysis (figure 9.3).

To obtain post shears and bending moments without reliance on a computer is a


straight forward process if the truss is assumed to be completely rigid. When this
assumption is made, the lateral movement, , of all posts at their truss attachment
point will be equal to that obtained using the methods outlined in Sections 9.5.2,
9.5.3 and 9.5.4. Post shear and post bending
9-6

National Frame Builders Association

Post-Frame Building Design Manual

moment can then be calculated using the following equations which assume zero
bending moment at the top of the post.

Vy = kp R i + s q (Hp y)

(9-2)

My = (s q / 2)(Hp y)2 + Vt (Hp y) (9-3)

Mmax = - Vt2 / (2 s q)

(9-4)

where:

Vy =
kp = =
Ri =
= = s= q= Hp = y= My =
Vt = Mmax =

post shear at distance y from base, lbf (N) post stiffness, lbf/in. (N/mm)
lateral movement of post top, in. (mm) contribution of wall pressure to eave load,
lbf (N) RW for windward wall RL for leeward wall frame spacing wall pressure,
lbf/ft2 (N/m2) post height, ft (m) distance from post base, ft (m) bending moment
in post at distance y from base, lbf-ft (N-m) Vy at y = Hp, lbf (N) bending moment
at y = Hp + Vt /(sq) (i.e., at the point of zero post shear)

Positive sign conventions for the preceding variables are illustrated in figure
9.5.

s xq

y
Hp Vy

Vy My

Figure 9.5. Positive sign convention for variables used in equations 9-2 and 9-3.

Using equation 9-2, the shears at the top, Vt, and bottom, Vb, of the windward post
of frame 4 are:
Vt = (91.9 lbf/in.)(0.655 in.) 365.8 lbf + (10 ft)(8.13 psf) (12 ft 12 ft)
Vt = 305.6 lbf
Vb = (91.9 lbf/in.)(0.655 in.) 365.8 lbf + (10 ft)(8.13 psf) (12 ft 0 ft)
Vb = 670.0 lbf
and the shears at the top, Vt, and bottom, Vb, of the leeward post of frame 4 are:
Vt = (91.9 lbf/in.)(0.655 in.) 228.6 lbf + (10 ft)(5.08 psf) (12 ft 12 ft)
Vt = 168.4 lbf
Vb = (91.9 lbf/in.)(0.655 in.) 228.6 lbf + (10 ft)(5.08 psf) (12 ft 0 ft)
Vb = 441.2 lbf
Equation 9-3 yields bending moments at the base of the windward and leeward posts
of 26200 and 19640 lbf-in., respectively. The difference between these values and
those in figure 9.3b are due to the rigid truss assumption.
According to equation 9-4, bending moments at the point of zero shear in the
windward and leeward posts are 6890 and 3350 lbf-in., respectively.
9.6 Step 5: Check Allowable Values
9.6.1 Diaphragm Shear. The actual maximum diaphragm shear stress of 93.1 lbf/ft
(Section 9.5.6) is less than the allowable of 107 lbf/ft (table 9.3) so the
diaphragm has sufficient strength.
9.6.2 Windward Post Stresses. Posts are subject to combined bending and compression
and must be checked per the requirements of the 1997 NDS Section 3.9.2 and NDS
equation 3.93. This equation, simplified for uniaxial bending is:

9-7

National Frame Builders Association

Post-Frame Building Design Manual

CSI = ( fc / Fc )2 + f b / {Fb [ 1 ( fc / FcE)]} < 1.0 (9-5)


where:

CSI = fc = fb = Fc = = Fb = = FcE = =

combined stress index actual compressive stress actual bending stress allowable
compressive stress
Fc CD CM CF Ci CP allowable bending stress
Fb CD CM CF Ci CLCr Cf CV critical buckling design stress K E I / ( le / d
)2

and:

Fc = Fb = CD = CM = CF = Ci = CP = CL = Cr = Cf = CV =
E =
I=
le /d = K=
=

tabulated compressive stress tabulated bending stress load duration factor wet
service factor size factor incising factor column stability factor beam stability
factor repetitive member factor form factor volume factor
E CM Ci moment of inertia slenderness ratio 0.3 for visually graded lumber 0.384
for machine evaluated lumber

Actual stresses for the windward post are: fc = PW / A = 3821 lbf / (30.25 in.2) =
126 lbf/in.2 f b = M / S = 26200 lbf-in. / (27.73 in.3) = 945 lbf/in.2 (at the
base) f b = 6890 lbf-in. / (27.73 in.3) = 248 lbf/in.2 (at point of zero shear)

For No. 2 Southern Pine timber, the tabulated compression and bending stresses and
modulus of elasticity are:

Fb = Fc = E=

850 lbf/in.2 525 lbf/in.2 1,200,000 lbf/in.2

Applicable adjustment factors are:

CD = CM =
CF = Ci =
CL = Cr =
Cf = CV = CP =
CP =

1.60 since the shortest duration load in the combination of loads is wind 1.00 for
modulus of elasticity, compression and bending of Southern Pine timber regardless
of moisture content 1.00 for nominal 6- by 6-inch No.2 Southern Pine 1.00 since
Southern Pine does not need to be incised for pressure treatments 1.00 since post
is square 1.00 because post spacing exceeds 24 inches. Note that this value is non-
zero for mechanically laminated posts 1.00 since posts are rectangular 1.00 since
posts are not gluedlaminated 1.00 at the base of the post where support is provided
in both directions is less than 1.00 at locations removed from supports that keep
the post from buckling. For such cases, CP is calculated using NDS equation 3.7-1.

It follows that at the base of both the windward and leeward posts:

Fc = Fb = FcE =

( 525 lbf/in.2)(1.60) = 840 lbf/in.2 ( 850 lbf/in.2)(1.60) = 1360 lbf/in.2 A very


large number if the effective buckling length, le, is assumed to be very small
because of support at the base. As a result, the ratio fc / FcE in equation 9-5 is
assumed to equal zero.

and at the base of the windward post:


CSI = ( 126 / 840 )2 + ( 945 / 1360 ) = 0.02 + 0.70 = 0.72 < 1.0 OK

The other critical location to check the combined stress index (CSI) is at the
point of maximum bending moment (point of zero shear) in the upper portion of the
post. At this location, the column stability factor is generally based on an
effective column buckling length of 0.8 Hp (see

9-8

National Frame Builders Association

Post-Frame Building Design Manual

NDS Appendix G), which results in the following slenderness ratio:

le / d = 0.8 (144 in.) / 5.5 in. = 20.9 thus:


FcE = 0.3 (1200000 lbf/in.2) / (20.9)2 = 820 lbf/in.2

The ratio of FcE / Fc is 0.976. This yields a Cp of 0.682, resulting in the


following allowable compressive stress, Fc.
Fc = ( 525 lbf/in.2)(1.60)(0.682) = 573 lbf/in.2

The CSI at the point of maximum moment in the upper portion of the post is:
CSI = ( 126 / 573 )2 + 248 / [1360 (1 - 126/820)]
= 0.05 + 0.22 = 0.27 < 1.0 OK

9.6.3 Windward Post Embedment. The windward post is constrained by the floor slab.
Since our example building is in an UBC jurisdiction, embedment depth will be
checked using equation 8-8 which is given as:

d=

4.25 Ma S b
1/3

For this example, the soil is assumed to be a firm silty sand which puts it in
class 4 (firm) of Table 8.1 a soil with a tabulated lateral soil pressure of 200
lbf/ft per foot of depth. In accordance with the UBC, the tabulated lateral
pressure can be adjusted for wind loading by a factor of 1.33. Since the posts are
more than six diameters apart, the allowable lateral pressure can also be doubled
for isolated conditions. Thus, the allowable lateral soil bearing pressure is:
S' = (200 lbf / ft2 / ft)(1.33)(2) = 532 lbf / ft2 / ft

As previously calculated, the moment at grade is 26200 lbf-in or 2180 lbf-ft.

The effective width of the post, b, is:

b = (1.4)(5.5 in)/12 = 0.64 ft

The minimum embedment depth, d, is:

d=

4.25 (2180 lbf-ft) 1/3 (532 lbf/ft3) (0.64 ft)

d = 3.00 ft < 4 ft OK

9.6.4 Leeward Post Stresses. Because (1) the axial force and maximum bending
moments associated with the leeward post are all less than those for the windward
post, (2) the windward and leeward posts are similarly supported, and (3) the
windward post is not overstressed, there is not need to check stresses in the
leeward post.

9.6.5 Leeward Post Embedment. Unless the post-frame designer makes special
provisions to tie the base of the leeward post to the floor slab, it will be non-
constrained. Since this is a UBC jurisdiction, embedment depth will be checked
using equation 8-6, which is given as follows:

d2 =

7.02 Va + 7.65 Ma / d S b

Solution of this equation is an iterative process. The values for S and b are as
determined for the windward post. Leeward post base shear and bending moment where
previously calculated as 441 lbf and 1640 lbf-ft, respectively

d2 =

7.02(441 lbf) + 7.65(1640 lbf-ft)/d (532 lbf/ft3) (0.64 ft)

d = 4.22 ft > 4 ft

At this point, the post-frame designer must apply engineering judgement. It is


important to remember that the analogs in this example produce conservative values
for base moments and shears, especially for the non-constrained case. The designer
must also consider what is known about the soil type and its variability on the
building site. If an embedment of 4 ft rather than 4.22 ft satisfies uplift
requirements as calculated elsewhere (not included in this example) an experienced
post-frame designer could validly judge that an embedment of 4 ft. is OK.

9-9
National Frame Builders Association
9.7 Example Summary
There are many items that the post-frame designer must still check. These include
but are not limited to:
The interconnection between diaphragms and shearwalls
Diaphragm chords Footings for gravity loads Uplift checks for embedded posts
All secondary members and headers The connections of all members, especially
truss to post End wall posts Diaphragms and shearwalls for wind against
the endwall
This example has focused solely on those items that are unique to post-frame. The
post-frame designer should be able to perform the remaining checks and designs
using commonly accepted practices and techniques.

Post-Frame Building Design Manual

9-10

You might also like