Resources

Structural Design of Unreinforced Composite Masonry

INTRODUCTION

Concrete masonry offers many textures, colors and sizes, along with choices in bond patterns and joint treatment making it an excellent choice for exterior and interior walls in residential, commercial and public buildings. Concrete brick can be used in both structural and veneer applications and is economical, durable, easy to maintain, fire resistant, and reduces sound transmission.

Multi-wythe masonry walls are classified as either composite or noncomposite depending on how the wythes interact. Connections between wythes of composite walls are designed to transfer stresses between the wythes, allowing the wythes to act as a single member in resisting loads. In contrast, for noncomposite or cavity walls each wythe individually resists the loads imposed on it. Concrete brick are used both in composite walls and as nonloadbearing veneer in cavity wall construction. Requirements for concrete brick veneers are summarized in Concrete Masonry Veneers, TEK 03-06C (ref. 1).

Standard Specification for Concrete Building Brick, ASTM C 55 (ref. 2), governs concrete brick and similar solid units. C 55 requirements are summarized in Concrete Masonry Unit Shapes, Sizes, Properties, and Specifications, CMUTEC-001-23 (ref. 3). 

 

STRUCTURAL DESIGN METHODS

Composite wall structural design requirements are contained in Building Code Requirements for Masonry Structures (ref. 4) and the International Building Code (ref. 5).

Allowable stress design of unreinforced composite walls is typically governed by the flexural tensile capacity of the masonry system (see Table 1), although compression and shear must also be checked. Shear stress in the plane of interface between wythes and collar joints is limited to 5 psi (34.5 kPa) for mortared collar joints; 10 psi (68.9 kPa) for grouted collar joints; and the square root of the unit compressive strength of the header (over the net area of the header) for headers.

Tables 2 through 13, for lateral loads with or without concentric axial loads (see Figure 1), are based on Chapter 2, Allowable Stress Design, of Building Code Requirements for Masonry Structures (ref. 4) and the following:

  1. specified compressive strength of masonry, f’m = 1500 psi (10.3 MPa),
  2. section modulus based on the minimum net area of the composite wall cross section,
  3. faceshell and web dimensions based on ASTM C 90 (ref. 6) minimum requirements for hollow units,
  4. loads include increase in allowable stress for load combinations including wind or seismic (where increase does not apply, multiply the Table values by ¾), and
  5. allowable tensile stress in masonry, Ft, for hollow ungrouted concrete masonry normal to the bed joints is as noted in Table footnotes.

CONSTRUCTION

Concrete brick walls and wythes of concrete brick should be laid with full head and bed mortar joints. For composite construction, the collar joint (the vertical longitudinal joint between wythes of masonry) is filled with grout or mortar to allow structural interaction between the wythes.

In composite walls, Building Code Requirements for Masonry Structures (ref. 4) requires that concrete brick be bonded to the backup wythe using either masonry headers or wall tie and grout or mortar. These minimum requirements, described below, help ensure that composite action is present between the wythes.

When bonded using masonry headers, the headers must make up at least 4 percent of the wall surface and extend at least 3 in. (76 mm) into the backing. The shear stress developed in the masonry header is limited to the square root of the unit compressive strength of the header (in psi (MPa) over the net area of the header).

Figure 2 illustrates wall tie spacing requirements for composite walls bonded with corrosion resistant ties or wire and collar joints filled with mortar or grout. Cross wires of joint reinforcement and rectangular ties are commonly used as wall ties for composite walls. Z-ties, however, are not permitted with ungrouted hollow masonry (ref. 7).

For cavity wall construction, the following construction recommendations apply:

  • keep cavity substantially clean to allow free water drainage,
  • install weep holes at 32 in. (813 mm) o. c.,
  • install granular fill, mesh or other mortar collection device in bottom of cavity to prevent mortar droppings from blocking weep holes, and
  • embed wall ties at least 1 ½ in. (38 mm) into the mortar bed of solid units.

REFERENCES

  1. Concrete Masonry Veneers, TEK 03-06C, Concrete Masonry & Hardscapes Association, 2012.
  2. Standard Specification for Concrete Brick, ASTM C 55-01. American Society for Testing and Materials, 2001.
  3. Concrete Masonry Unit Shapes, Sizes, Properties, and Specifications, CMU-TEC-001-23, Concrete Masonry & Hardscapes Association, 2023.
  4. Building Code Requirements for Masonry Structures, ACI 530-99/ASCE 5-99/TMS 402-99. Reported by the Masonry Standards Joint Committee, 1999.
  5. International Building Code. International Code Council, 2000.
  6. Standard Specification for Loadbearing Concrete Masonry Units, ASTM C 90-01. American Society for Testing and Materials, 2001.
  7. Specification for Masonry Structures, ACI 530.1-99/ ASCE 6-99/TMS 602-99. Reported by the Masonry Standards Joint Committee, 1999.

Structural Testing of Concrete Masonry Assemblages

STRUCTURAL TESTING OF CONCRETE MASONRY ASSEMBLAGES

INTRODUCTION

A considerable amount of research has been conducted on concrete masonry units and assemblages in order to develop design stresses for building codes and to evaluate existing building systems. The properties of concrete masonry which are considered most important and which have been the subject of research on assemblages of concrete masonry units include: structural, fi re resistance, thermal insulation, noise insulation, and resistance to moisture penetration. This TEK is concerned with testing structural properties and reviews the kinds of loads and stresses which concrete masonry walls may be subjected to in service and the principal details of ASTM Test Methods used to investigate the structural behavior of masonry walls.

TYPES OF LOADS

Loads acting on masonry walls may be classified as vertical dead and live loads, sometimes called gravity loads, and lateral loads, these being due to wind, earthquake, earth or water pressure, etc. Vertical loads may be more or less uniformly distributed along the length of the wall or may include one or several concentrated loads which are transmitted to small areas of the wall section. Vertical loads may be centered in the same plane as the centroidal axis of the wall (concentric or axial loading) or at some distance away from this axis (eccentric loading). Eccentric loads produce bending as well as direct compressive stresses and consequently are more severe than concentric loading. Lateral loads may be uniformly distributed on the vertical surface as in the case of wind, or nonuniform according to some function of other factors, as in the case of earthquake or fluid pressure loads. Lateral loads may be concentrated and their direction may be normal or parallel to or at any intermediate angle with the wall surface. Loads may be gradually or quickly applied (impact), permanent (dead loads) or transient (wind load). Vertical and lateral loads may act simultaneously to produce a combination of axial compression and flexural stress in the masonry.

TYPES OF LOADING USED IN TESTS

Despite the variety of load types and stress conditions, masonry walls can be safety designed provided their ultimate strength in direct compression, flexure and shear is known. In developing this information the test procedures employed in the past often have varied as to specimen size, loading method and other details, but when properly presented and interpreted the results have proved to be applicable and useful. Beginning in the 1930’s, the National Bureau of Standards adopted standard methods for use in their own investigations of building constructions, these methods later forming the basis for ASTM Standard E 72, Standard Test Method for Conducting Strength Tests of Panels for Building Construction (ref. 2).

Figures 1 through 6 show schematically the general procedures used in conducting compression, flexural, and racking.

COMPRESSION

Test Methods

According to ASTM E 72, compressive strength tests are made on specimens having a height equal to the height of the wall in actual use and having a nominal width of 4 ft (1.2 m). Generally, story height walls (nominal height = 8 ft) are typical of those tested but compressive strength tests have been conducted on masonry walls over 20 ft high.

Referring to Figure 1, note that compression tests are made with the load line located a distance one-third the wall thickness from the inside face of the wall (eccentric loading) or at the central plane of the wall (axial loading). Eccentric loading is prescribed in ASTM E 72 and for many years has been preferred over axial loading by many investigators since it approximates more closely the loading condition of walls in buildings.

Loading at the central plane or centerline of the wall is true axial loading only when the wall section is geometrically and elastically symmetrical with respect to the center line.

With the advent of engineered loadbearing masonry design, simpler and less expensive test methods for determining compressive strength properties of masonry have come into wide usage: ASTM E 447, Standard Test Method for Compressive Strength of Masonry Prisms (ref. 1), and ASTM C 1314, Standard Test Method for Constructing and Testing Masonry Prisms Used to Determine Compliance with Specified Compressive Strength of Masonry (ref. 3). These tests prescribe methods for testing short compression prisms made of the same masonry units, mortar, and workmanship to be used in the construction. Although the test methods are similar, ASTM E 447 is intended for research purposes only (not for construction quality assurance purposes as is C 1314), and requires collection of additional detailed information associated with research tests.

Stresses Due to Applied Loads

The type of loading largely determines the general shape of the stress distribution diagram for the wall section. For solid walls, Figure 2, axial loading results in rectangular stress diagram, the fiber stresses being uniform over the entire cross section and equal to P/A. If the vertical load is applied eccentrically or off-center by a distance of one-sixth the wall thickness (t/6), the unit stress varies from a maximum of 2P/A at the wall face nearest the load line, to zero stress at the opposite face, Figure 2b. Eccentricity greater than t/6 would produce tensile stresses at the opposite face and a stress diagram which would show zero stress at some point between the wall faces. Eccentricity less than t/6 would result in some compression at the opposite face and a stress diagram of trapezoid shape.

In Figures 2a and 2b, the average compressive stress is the same in each case, assuming the same vertical load, but as noted, the maximum fiber stress for the eccentrically loaded wall is twice that for the axially loaded wall. A logical deduction is that a given wall will support a greater axial than eccentric load. This is borne out by tests which indicate that bending stress due to eccentric location of vertical load or other causes, reduces the ultimate vertical load capacity of masonry below its axial load strength.

FLEXURAL

The different types of loading used in testing masonry walls for flexural or transverse strength are shown in Figure 3, together with the moment and shear diagrams and formulas for calculating maximum shear, moment and deflection assuming simply supported spans. ASTM E 72 specifies a specimen width of 4 ft and either line loading applied at the outer 1/4 points of the span as shown in sketch (b) of the figure, or uniformly distributed loading shown in sketch (c). Uniform transverse loading of upright specimens such as masonry walls has become practical and more commonly used with the development of the “bag method” in which a plastic or rubberized fabric bag, interposed between the wall face and a backboard, is inflated with air to give increments of pressure against the wall until failure. Comparing the deflection formulas for the three loading methods, it will be noted that 1/4-point loading causes the greatest deflection, assuming a given moment and wall section, and from this standpoint is more severe than the other two methods. Also, any line load method produces combinations of maximum shear and moment in the same region which generally results in a lower indicated strength than would be obtained with uniform loading where the regions of maximum moment and of maximum shear are widely separated. It also appears that a test loading which does not include concentrated loads more nearly simulates the more common loads considered in design, such as wind and fluid or earth pressure.

Flexural strength of unreinforced masonry assemblages can also be measured by other ASTM methods. Method E 518, Standard Test Method for Flexural Bond Strength of Masonry (ref. 5), is intended to provide simplified and economical means for gathering comparative research data on the flexural bond strength developed with different types of masonry units and mortar or for the purpose of checking job quality control as regards materials and workmanship. Unlike ASTM E 72, Method E 518 is typically not intended for use in establishing design stresses. Specimens are small prisms laid up in stack bond and tested in a horizontal position by applying load at the third-points or by applying a uniform load by means of an inflated air bag.

Method C 1072, Standard Test Method for Measurement of Masonry Flexural Bond Strength (ref. 6) covers physical testing of each joint of masonry prisms using a bond wrench test apparatus. This method permits the measurement of multiple joints in a prism rather than the single joint tests of E 82 and E 518, making statistical evaluations easier and more cost effective. The results are used to: determine compatibility of mortars and masonry units; determine the effect on flexural bond strength of factors such as mortar properties and workmanship; and predict the flexural strength of a wall. The flexural bond strength determined using C 1072 is not typically used to predict the flexural bond strength of a wall constructed of the same material unless testing is performed to document the difference between the two; nor to determine extent of bond for a water permeance evaluation.

RACKING

Racking strength tests are performed according to the general schemes shown in Figure 4 and give an indication of the resistance of the construction to the horizontal component of shearing forces acting parallel to the wall. In the method prescribed in ASTM E 72, shown in the upper sketch, the horizontal component is equal to the load P, and the principal stress, compression acting in a diagonal line between the load point and lower right reaction, is the resultant of load P and vertical reaction Rv.

In the scheme shown in the lower sketch, the load is applied diagonally downward and the horizontal component or longitudinal shear is equal to P cosφ , approximately 0.7P since φ is usually about 45o. This alternative method is not addressed in ASTM E 72, but has often been used because it eliminates the need for tie-down rods.

Results of racking tests of masonry walls generally are given in terms of the maximum horizontal component, pounds per linear foot of wall, although the total load P may also be reported. While failure is considered to be in shear, it actually is caused by a combination of shear and secondary tensile stresses, the latter acting normal to the compressive stress. Although exceeding both the shearing and secondary tensile stresses in intensity, compressive stress is not sufficient to cause a compression-type failure.

As with compression and flexural testing, an ASTM method exists for testing shear on specimens smaller than required in ASTM E 72. In ASTM E 519, Standard Test Method for Diagonal Tension (Shear in Masonry Assemblages) (ref. 4), 4 ft by 4 ft masonry assemblages are positioned in a compression testing machine so that a compressive load is applied along one diagonal, causing a diagonal tension failure with the specimen splitting apart normal to the direction of load (see Figure 5). This method also avoids the need for a hold down force to prevent rotation of the specimen as used in the E 72 method, simplifying the analysis of the state of stress in the specimen.

IMPACT

The impact test of ASTM E 695 (ref. 7) affords a qualitative measure of the capacity of wall, floor and roof panels to resist impact loading. The impact force is obtained from the free-fall of a bag of lead shot through a path which causes it to strike the center of the panel at an angle normal to the surface. The essential details of the method as adapted to the testing of upright masonry wall panels are shown in Figure 6. Panels are typically 4 ft long and are simply supported on a span six inches less than the height of the specimen. The height of drop is increased in increments until failure occurs, but not exceeding 10 ft, and the maximum drop is the value reported.

The structural testing of masonry walls and assemblages encompasses much more than merely determining the ultimate load at failure. At each load increment, strains and deflections are carefully measured with precision instruments. at various locations on the specimen. In some procedures a load increment is applied and measurements are taken after which the load is released and measurements again taken to determine the residual strain or deflection. The specimen is examined and notes made of any cracking, crushing or other visible distress. This process is repeated at each increase in load so that when the test has been concluded the research engineer has accumulated the data needed to give a clear picture of the structural behavior of the specimen through all stages of loading.

REFERENCES

  1. Standard Test Method for Compressive Strength for Laboratory Constructed Masonry Prisms, ASTM E 447-97. American Society for Testing and Materials, 1997.
  2. Standard Test Method for Conducting Strength Tests of Panels for Building Construction, ASTM E 72-95. American Society for Testing and Materials, 1995.
  3. Standard Test Method for Constructing and Testing Masonry Prisms Used to Determine Compliance with Specified Compressive Strength of Masonry, ASTM C 1314-97. American Society for Testing and Materials, 1997.
  4. Standard Test Method for Diagonal Tension (Shear) in Masonry Assemblages, ASTM E 519-81(1993)e1. American Society for Testing and Materials, 1993.
  5. Standard Test Method for Flexural Bond Strength of Masonry, ASTM E 518 -80(1993)e1. American Society for Testing and Materials, 1993.
  6. Standard Test Method for Measurement of Masonry Flexural Bond Strength, ASTM C 1072-94. American Society for Testing and Materials, 1994.
  7. Standard Test Method for Measuring Relative Resistance of Wall, Floor, and Roof Construction to Impact Loading, ASTM E 695-79(1997)e1. American Society for Testing and Materials, 1997.

Precast Concrete Lintels for Concrete Masonry Construction

INTRODUCTION

Lintels function as beams to support the wall weight and other loads over an opening, and to transfer these loads to the adjacent masonry. Because of their rigidity, strength, durability, fire resistance and aesthetics, the most common types of lintels for concrete masonry construction are those manufactured of precast reinforced concrete or reinforced concrete masonry units (ref. 3). The color and surface texture of these lintels can be used as an accent or to duplicate the surrounding masonry.

LINTEL DIMENSIONS

Precast lintel dimensions are illustrated in Figure 1. Precast concrete lintels are manufactured to modular sizes, having specified dimensions corresponding to the concrete masonry units being used in the construction.

A modular lintel length should be specified, with a minimum length of the clear span plus 8 in. (203 mm), to provide at least 4 in. (102 mm) bearing at each end (ref. 1). Additionally, if lintels are subjected to tensile stresses during storage, transportation, handling, or placement, it is recommended that steel reinforcement be provided in both the top and bottom to prevent cracking. Minimum concrete cover over the steel should be 1 ½ in. (13 mm). The lintel width, or width of the combination of side-by-side lintels, should equal the width of the supported masonry wythe.

Lintels should be clearly marked on the top whenever possible to prevent the possibility of improper installation in the wall. In the event the top of the lintel is not marked and may be installed upside down, the same size bars should be used in both the top and bottom.

LINTEL DESIGN

Precast concrete lintels are designed using the strength design provisions of Building Code Requirements for Structural Concrete, ACI 318-99 (ref. 2). In strength design, service loads are increased to account for variations in anticipated loads, becoming factored loads. The lintel is then sized to provide sufficient design strength. Further information on determining design loads for lintels is included in ASD of CM Lintels Based on 2012 IBC/2011 MSJC, TEK 17-01D (ref. 3).

Nominal lintel strength is determined based on the strength design provisions of ACI 318 and then reduced by strength reduction factors, called phi (Φ) factors. These factors account for any variability in materials and construction practices. The resulting capacity needs to equal or exceed the factored loads. Precast concrete strength reduction factors are 0.9 and 0.85 for flexure and shear, respectively (ref. 2).

Tables 1 through 4 list design moment and shear strengths for various precast lintel sizes and concrete strengths, based on the following criteria (ref. 2).

Flexural strength:

Shear strength, no shear reinforcement:

ACI 318 contains requirements for minimum and maximum reinforcing steel areas to ensure a minimum level of performance. Minimum reinforcement area for lintels is As min = 3(f’c)½bd/fy but not less than 200bd/fy. In addition, the reinforcement ratio is limited to 75% of the balanced reinforcement ratio, ρmax = 0.75ρb.

Deflection criteria for lintels is based on controlling cracking in the masonry being supported. Consequently, less deflection is allowed when the lintel supports unreinforced masonry. In this case, lintel deflection is limited to the effective span of the lintel (measured in inches) divided by 600 (L/600) (ref. 1). In addition, ACI 318 limits precast lintel deflection to L/240 when the element supported by the lintel is not likely to be damaged by large deflections, and L/480 when the element supported by the lintel is likely to be damaged by large deflections. Lintel deflection is calculated based on the effective moment of inertia, Ie, as follows (ref. 2, Section 9.5.2.3).

Shrinkage and creep due to sustained loads cause additional long-term deflections over and above those occurring when loads are first applied. ACI 318 requires that deflections due to shrinkage and creep are included, and provides an expression to estimate this additional deflection (ACI 318 Section 9.5.2.5):

λ = ξ/(1+50ρ’)

where ξ = 2.0 for exposures of 5 years or more.

DESIGN EXAMPLE

The residential basement wall shown in Figure 3 needs a lintel over the window opening. The floor live load is 400 lb (1.8 kN) per joist and the floor dead load is 100 lb (0.44 kN) per joist. Consider the floor joist loads, spaced at 16 in. (406 mm) on center, as uniformly distributed. Use a lintel self-weight of 61 lb/ ft (0.89 kN/m) and weight of 77.9 lb/ft2 (3.73 kPa) for the bond beam at the top of the wall over the lintel.

Determine effective depth, d: Assuming an 8 in. (203 mm) high lintel with two No. 4 (13M) bars,
d = 7.625 in. – 1.5 in. – 0.5/2 in.
= 5.88 in. (149 mm)

Check for arching action: The effective span length, L = 96 + 5.88 = 101.9 in. (2588 mm). Since the height of masonry above the opening is less than L/2, arching of the masonry over the opening cannot be assumed (see ref. 4 for detailed information about determining arching action).

Determine design loads:
LL = (400 lb)(12/16 in.) = 300 lb/ft (4.4 kN/m)
Dead loads include floor, wall, and lintel self-weight.
Dfloor = 100 lb (12/16 in.) = 75 lb/ft (1.1 kN/m)
Dlintel = 61 lb/ft (0.89 kN/m)
Db beam = (77.9lb/ft²)(7.625/12 ft)= 50 lb/ft (0.31 kN/m)
Dtotal = (75 + 61 + 50) = 186 lb/ft (3.2 kN/m)

For deflection calculations use loads as given above. For strength design multiply live loads by 1.7 and dead loads by 1.4. Maximum moment and shear for strength design:

Mmax = wL²/8
= {[(1.7)(300)+(1.4)( 186 ) lb/ft](101.9 in.)²/8}(ft/12 in.)
= 83,328 in.-lb (9.4 kN m)

Vmax = wL/2 (at distance “d” from support) (ref.2)
= [(1.7)(300)+(1.4)(186 lb/ft)](101.9/2-5.88 in.)(ft/12 in.)
= 2,893 lb (12.9 kN)

From Table 3, an 8 x 8 in. (203 x 203 mm) lintel with two No. 4 (13M) bars and f ‘c = 4000 psi (20.7 MPa) has sufficient strength.

Check deflection: Deflection is determined using the effective moment of inertia of the lintel, Ie, calculated as follows (ref. 2).

Ec = wc1.533(f’c)½ = (150 pcf)1.533(4000 psi)½
= 3,834,000 psi (26,400 MPa)
fr = 7.5(f’c)½ = 474 psi (3.3 MPa)
yt = 7.625 in./2 = 3.81 in. (97 mm)
Ig = bh³/12 = (7.625 in.)(7.625 in.)³/12
= 282 in.4 (11,725 cm4)
Mcr = frIg/yt = 474 psi(282 psi)/3.81 in.
= 35,083 in.-lb (4.0 kN⋅m)
Mmax uf = wL²/8 = [(300+186 lb/ft)(101.9 in.)²/8](ft/12 in.)
= 52567 in.-lb (5.9 kN⋅m)
(Mcr/Mmax uf)³ = (35,083/52567)³ = 0.297
n = Es/Ec = 29,000,000/3,834,000 = 7.6
ρ = As/bd = 0.40 in.²/(7.625 in.)(5.88 in.) = 0.00892
= 7.6(0.00892) = 0.0678
c = nρd[(1 + 2/)½ – 1]
= 0.0678(5.88 in.)[(1+ 2/0.0678)½-1] = 1.80 in. (45 mm)
Icr = bc³/3 + nAs (dc
= 7.625 in.(1.8 in.)³/3 + 7.6(0.4 in.²)(5.88 – 1.8)²
= 65.4 in.4 (2714 cm4)
Ie = (Mcr/Mmax ufIg + [1- (Mcr/Mmax uf)³]Icr
= 0.297(282) + [1-0.297]65.4 in.4
= 130 in.4 (5411 cm4) < Ig OK

For a simply supported beam under uniform load,

max = 5wL4/384EcIe
= 5(300 + 186 lb/ft)(101.9 in.)4/[384(3,834,000 psi)(130 in.4)]/(12 in./ft)
= 0.114 in. (2.9 mm)

Long-term deflection multiplier,
λ = ξ/(1+50ρ’) = 2/[1 + 50(0)] = 2

Long-term deflection,
LT = λ∆max = 2(0.114 in.) = 0.228 in. (5.8 mm)

Total deflection,
tot = max + LT = 0.114 + 0.228 = 0.342 in. (8.7 mm)

Deflection limit for this case is L/240 = 101.9 in./240
= 0.42 in. (10.7 mm) > 0.342 in. (8.7 mm) OK

NOTATIONS

a             = depth of equivalent rectangular stress block, in. (mm)
As           = area of tension reinforcement, in.² (mm²)
b             = actual width of lintel, in. (mm)
c              = distance from extreme compression fiber to neutral axis, in. (mm)
C             = resultant compressive force in concrete, lb (kN)
d              = distance from extreme compression fiber to centroid of tension reinforcement, in. (mm)
Db beam   = dead load of bond beam, lb/ft (kN/m)
Dfloor      = dead load of floor, lb/ft (kN/m)
Dlintel      = dead load of lintel, lb/ft (kN/m)
Dtot         = total design dead load, lb/ft (kN/m)
Ec            = modulus of elasticity of concrete, psi (MPa)
f ‘c           = specified compressive strength of concrete, psi (MPa)
fr             = modulus of rupture of concrete, psi (MPa)
fy             = specified yield strength of reinforcement, psi (MPa) (60,000 psi, 413 MPa)
Icr            = moment of inertia of cracked section transformed to concrete, in.4 (cm4)
Ie              = effective moment of inertia, in.4 (cm4)
Ig              = moment of inertia of gross concrete section about centroidal axis, in.4 (cm4)
L               = effective length, clear span plus depth of member, not to exceed the distance between center of supports, in. (mm)
LL             = live load, lb/ft (kN/m)
Mcr           = cracking moment, in.-lb (kN⋅m)
Mmax        = maximum factored moment on section, in.-lb (kN⋅m)
Mmax uf     = maximum unfactored moment on section, in.-lb (kN⋅m)
Mn             = nominal moment strength, in.-lb/ft (kN⋅m/m)
n                = modular ratio, Es/Ec
T                = resultant tensile force in steel reinforcement, lb (kN)
Vmax          = maximum factored shear on section, lb (kN)
Vn               = nominal shear strength, lb (kN)
w                = uniform load, lb/in. (kN/m)
wc               = density of concrete, pcf (kN/m³)
yt                = distance from centroidal axis of gross section to extreme fiber in tension, in. (mm)
max          = maximum immediate deflection, in. (mm)
LT            = long-term deflection, in. (mm)
tot            = total deflection, in. (mm)
εc               = strain in concrete, in./in. (mm/mm)
εs               = strain in steel reinforcement, in./in. (mm/mm)
ξ                 = time-dependent factor for sustained load
λ                 = multiplier for additional long-term deflection
Φ                = strength reduction factor
ρ                 = reinforcement ratio, As/bd
ρ’                = reinforcement ratio for nonprestressed compression reinforcement, As/bd
ρb               = reinforcement ratio producing balanced strain conditions
ρmax           = limit on reinforcement ratio

REFERENCES

  1. Building Code Requirements for Masonry Structures, ACI 530-99/ASCE 5-99/TMS 402-99. Reported by the Masonry Standards Joint Committee, 1999.
  2. Building Code Requirements for Structural Concrete, ACI 318-99. American Concrete Institute, 1999.
  3. ASD of CM Lintels Based on 2012 IBC/2011 MSJC, TEK 17-01D, Concrete Masonry & Hardscapes Association, 2011.

Strength Design of Reinforced Concrete Masonry Foundation Walls

INTRODUCTION

Although concrete masonry foundation walls can be constructed without reinforcing steel, reinforcement may be required for walls supporting large soil backfill loads. The strength design provisions found in Chapter 3 of Building Code Requirements for Masonry Structures (ref. 1) typically provides increased economy over the allowable stress design method, as thinner walls or larger reinforcing bar spacings often result from a strength design analysis. Strength design criteria are presented in detail in TEK 14-04B, Strength Design Provisions for Concrete Masonry (ref. 2).

DESIGN LOADS

Soil imparts lateral loads on foundation walls. The load is assumed to increase linearly with depth, resulting in a triangular load distribution on the wall. This lateral soil load is expressed as an equivalent fluid pressure, with units of pounds per square foot per foot of depth (kN/m²/m). For strength design analysis, this lateral soil pressure is increased by multiplying by a load factor, which provides a factor of safety against overload conditions. The maximum moment on the wall depends on the total wall height, the soil backfill height, the wall support conditions, the factored soil load, the existence of any surcharges on the soil and the presence of saturated soils.

Foundation walls also provide support for the structure above the foundation, transferring vertical loads to the footing. Vertical compression counteracts flexural tension, increasing the wall’s resistance to flexure. In low-rise construction, these vertical loads are typically small in relation to the compressive strength of the concrete masonry. Vertical load effects are not addressed in this TEK.

DESIGN TABLES

Tables 1 through 4 present reinforcement schedules for 6, 8, 10 and 12-in. (152, 203, 254 and 305-mm) walls, respectively. Additional reinforcement alternatives may be appropriate, and can be verified with an engineering analysis. Walls from 8 to 16 ft (2.4 to 4.9 m) high and soil pressures of 30, 45 and 60 psf/ft (4.7, 7.0, and 9.4 kN/m²/m) are included.

The effective reinforcement depth, d, assumed for the analyses are practical values, taking into account variations in face shell thickness, a range of reinforcing bar sizes, minimum required grout cover and construction tolerances for placing the reinforcement.

The following assumptions also apply to the values in Tables 1 through 4:

  1. there are no surcharges on the soil adjacent to the wall,
  2. there are negligible axial loads on the wall,
  3. the wall is simply supported at top and bottom,
  4. the wall is grouted at cells containing reinforcement (although solid grouting is acceptable),
  5. section properties are based on minimum face shell and web thickness requirements of ASTM C 90 (ref. 3),
  6. the specified compressive strength of masonry, f’m, is 1500 psi (10.3 MPa),
  7. Grade 60 (413 MPa) reinforcement,
  8. reinforcement requirements listed account for a soil load factor of 1.6 (ref. 6),
  9. the maximum width of the compression zone is limited to six times the wall thickness, or a 72 in. (1,829 mm) vertical bar spacing, whichever is smaller,
  10. reinforcing steel is placed toward the tension (interior) face of the wall (as shown in Figure 1), and
  11. the soil is well drained to preclude the presence of saturated soil.

DESIGN EXAMPLE

Wall: 12-in. (305 mm) thick concrete masonry foundation wall, 12 ft (3.66 m) high

Soil: equivalent fluid pressure is 45 psf/ft (7.0 kN/m²/m) (excluding soil load factors), 10 ft (3.05 m) backfill height

Using Table 4, the wall can be adequately reinforced using No. 9 bars at 72 in. o.c. (M# 29 at 1,829 mm).

CONSTRUCTION ISSUES

This section discusses those issues which directly relate to structural design assumptions. See TEK 03-11, Concrete Masonry Basement Wall Construction and TEK 05-03A, Concrete Masonry Foundation Wall Details (refs. 4, 5) for more complete information on building concrete masonry foundation walls.

Figure 1 illustrates wall support conditions, drainage and protection from water. Before backfilling, the floor diaphragm must be in place, or the wall must be properly braced to resist the soil load. Ideally, the backfill should be free-draining granular material, free from expansive soils or other deleterious materials.

The assumption that there are no surcharges on the soil means that heavy equipment should not be operated directly adjacent to any basement wall system. In addition, the backfill materials should be placed and compacted in several lifts. Care should be taken when placing backfill materials to prevent damaging the drainage, waterproofing or exterior insulation systems.

REFERENCES

  1. Building Code Requirements for Masonry Structures, ACI 530-02/ASCE 5-02/TMS 402-02. Reported by the Masonry Standards Joint Committee, 2002.
  2. Strength Design Provisions for Concrete Masonry, TEK 14-04B, Concrete Masonry & Hardscapes Association, 2008.
  3. Standard Specification for Loadbearing Concrete Masonry Units, ASTM C 90-03. ASTM International, 2003.
  4. Concrete Masonry Basement Wall Construction, TEK 0311, Concrete Masonry & Hardscapes Association, 2001.
  5. Concrete Masonry Foundation Wall Details, TEK 05-03A, Concrete Masonry & Hardscapes Association, 2003.
  6. Minimum Design Loads for Buildings and Other Structures, ASCE 7-02. American Society of Civil Engineers, 2002.

 

Allowable Stress Design of Concrete Masonry Foundation Walls

INTRODUCTION

Basements provide: economical living, working and storage areas; convenient spaces for mechanical equipment; safe havens during tornadoes and other violent storms; and easy access to plumbing and ductwork. Concrete masonry is well suited to basement and foundation wall construction due to its inherent durability, compressive strength, economy, and resistance to fire, termites, and noise.

Traditionally, residential basement walls have been constructed of plain (unreinforced) concrete masonry, often designed empirically. Walls over 8 ft (2.4 m) high or with larger soil loads are typically designed using reinforced concrete masonry or using design tables included in building codes such as the International Building Code (ref. 4).

DESIGN LOADS

Soil imparts a lateral load on foundation walls. For design, the load is traditionally assumed to increase linearly with depth resulting in a triangular load distribution. This lateral soil load is expressed as an equivalent fluid pressure, with units of pounds per square foot per foot of depth (kPa/m). The maximum force on the wall depends on the total wall height, soil backfill height, wall support conditions, soil type, and the existence of any soil surcharges. For design, foundation walls are typically assumed to act as simple vertical beams laterally supported at the top and bottom.

Foundation walls also provide support for the structure above, transferring vertical loads to the footing. When foundations span vertically, this vertical compression counteracts flexural tension, increasing the wall’s resistance to flexure. In low-rise construction, these vertical loads are typically small in relation to the compressive strength of concrete masonry. Further, if the wall spans horizontally, vertical compression does not offset the flexural tension. Vertical load effects are not included in the tables and design example presented in this TEK (references 2 and 3 include vertical load effects).

EMPIRICAL DESIGN

The empirical design method uses historical experience to proportion and size masonry elements. Empirical design is often used to design concrete masonry foundation walls due to its simplicity and history of successful performance.

Table 1 lists the allowable backfill heights for 8, 10 and 12-inch (203, 254 and 305 mm) concrete masonry foundation walls. Table 1 may be used for foundation walls up to 8 feet (2.4 m) high under the following conditions (ref. 1):

  1. terrain surrounding the foundation wall is graded to drain surface water away from foundation walls,
  2. backfill is drained to remove ground water away from foundation walls,
  3. tops of foundation walls are laterally supported prior to backfilling,
  4. the length of foundation walls between perpendicular masonry walls or pilasters is a maximum of 3 times the foundation wall height,
  5. the backfill is granular and soil conditions in the area are non-expansive,
  6. masonry is laid in running bond using Type M or S mortar, and
  7. units meet the requirements of ASTM C 90 (ref. 6).

Where these conditions cannot be met, the wall must be engineered using either an allowable stress design (see following section) or strength design procedure (see ref. 5).

WALL DESIGN

Tables 2 through 4 of this TEK have been rationally designed in accordance with the allowable stress design provisions of Building Code Requirements for Masonry Structures (ref. 1) and therefore meet the requirements of the International Building Code even though the latter limits reinforcment spacing to 72 in. (1829 mm) when using their tables. Additional reinforcement alternatives may be appropriate and can be verified with an engineering analysis.

Tables 2, 3 and 4 list reinforcement options for 8, 10 and 12-in. (203, 254 and 305-mm) thick walls, respectively. The effective depths of reinforcement, d, (see Table notes) used are practical values, taking into account variations in face shell thickness, a range of bar sizes, minimum required grout cover, and construction tolerances for placing the reinforcing bars.

Tables 2 through 4 are based on the following:

  1. no surcharges on the soil adjacent to the wall and no hydrostatic pressure,
  2. negligible axial loads on the wall,
  3. wall is simply supported at top and bottom,
  4. wall is grouted only at reinforced cells,
  5. section properties are based on minimum face shell and web thicknesses in ASTM C 90 (ref. 6),
  6. specified compressive strength of masonry, f’m, is 1,500 psi (10.3 MPa),
  7. reinforcement yield strength, fy, is 60,000 psi (414 MPa),
  8. modulus of elasticity of masonry, Em, is 1,350,000 psi (9,308 MPa),
  9. modulus of elasticity of steel, Es, is 29,000,000 psi (200,000 MPa),
  10. maximum width of compression zone is six times the wall thickness (where reinforcement spacing exceeds this distance, the ability of the plain masonry outside the compression zone to distribute loads horizontally to the reinforced section was verified assuming two-way plate action),
  11. allowable tensile stress in reinforcement, Fs, is 24,000 psi (165 MPa),
  12. allowable compressive stress in masonry, Fb, is f’m (500 psi, 3.4 MPa),
  13. grout complies with ASTM C 476 (2,000 psi (14 MPa) if property spec is used) (ref. 7), and
  14. masonry is laid in running bond using Type M or S mortar and face shell mortar bedding.

DESIGN EXAMPLE

Wall: 12-inch (305 mm) thick, 12 feet (3.7 m) high.

Loads: equivalent fluid pressure of soil is 45 pcf (7.07 kPa/ m), 10 foot (3.1 m) backfill height. No axial, seismic, or other loads.

Using Table 4, #8 bars at 40 in. (M 25 at 1016 mm) o.c. are sufficient.

CONSTRUCTION ISSUES

This section is not a complete construction guide, but rather discusses those issues directly related to structural design assumptions. Figures 1 and 2 illustrate typical wall support conditions, drainage, and water protection.

Before backfilling, the floor diaphragm must be in place or the wall must be properly braced to resist the soil load. In addition to the absence of additional dead or live loads following construction, the assumption that there are no surcharges on the soil also means that heavy equipment should not be operated close to basement wall systems that are not designed to carry the additional load. In addition, the backfill materials should be placed and compacted in several lifts, taking care to prevent wall damage. Care should also be taken to prevent damaging the drainage, waterproofing, or exterior insulation systems, if present.

REFERENCES

  1. Building Code Requirements for Masonry Structures, ACI 530-99/ASCE 5-99/TMS 402-99. Reported by the Masonry Standards Joint Committee, 1999.
  2. International Building Code. International Code Council, 2000.
  3. Strength Design of Reinforced CM Foundation Walls, TEK 15-02B, Concrete Masonry & Hardscapes Association, 2004.
  4. Standard Specification for Loadbearing Concrete Masonry Units, ASTM C 90-01. American Society for Testing and Materials, 2001.
  5. Standard Specification for Grout Masonry, ASTM C476- 01. American Society for Testing and Materials, 2001.

Post-Tensioned Concrete Masonry Wall Design

INTRODUCTION

The 1999 Building Code Requirements for Masonry Structures, ACI 530/ASCE 5/TMS 402 (ref. 1), was the first masonry code in the United States to include general design provisions for prestressed masonry. Prestressing masonry is a process whereby internal compressive stresses are introduced to counteract tensile stresses resulting from applied loads. Compressive stresses are developed within the masonry by tensioning a steel tendon, which is anchored to the top and bottom of the masonry element (see Figure 1). Post-tensioning is the primary method of prestressing, where the tendons are stressed after the masonry has been placed. This TEK focuses on the design of concrete masonry walls constructed with vertical post-tensioned tendons.

Advantages

Prestressing has the potential to increase the flexural strength, shear strength and stiffness of a masonry element. In addition to increasing the strength of an element, prestressing forces can also close or minimize the formation of some cracks. Further, while research (refs. 14, 15) indicates that ductility and energy dissipation capacity are enhanced with prestressing, Building Code Requirements for Masonry Structures (ref. 1) conservatively does not take such performance into account.

Post-tensioned masonry can be an economical alternative to conventionally reinforced masonry. One major advantage of prestressing is that it allows a wall to be reinforced without the need for grout. Also, the number of prestressing tendons may be less than the number of reinforcing bars required for the same flexural strength.

Post-tensioning masonry is primarily applicable to walls, although it can also be used for beams, piers, and columns. Vertical post-tensioning is most effective for increasing the structural capacity of elements subjected to relatively low axial loads. Structural applications include loadbearing, nonloadbearing and shear walls of tall warehouses and gymnasiums, and commercial buildings, as well as retaining walls and sound barrier walls. Post-tensioning is also an option for strengthening existing walls.

MATERIALS

Post-tensioned wall construction uses standard materials: units, mortar, grout, and perhaps steel reinforcement. In addition, post-tensioning requires tendons, which are steel wires, bars or strands with a higher tensile strength than conventional reinforcement. Manufacturers of prestressing tendons must supply stress relaxation characteristics for their material if it is to be used as a prestressing tendon. Specifications for those materials used specifically for post-tensioning are given in Table 1. Other material specifications are covered in references 9 through 12. Construction is covered in  Post-Tensioned Concrete Masonry Wall Construction, TEK 03-14 (ref. 3).

CORROSION PROTECTION

As with conventionally reinforced masonry structures, Building Code Requirements for Masonry Structures (ref. 1) mandates that prestressing tendons for post-tensioned masonry structures be protected against corrosion. As a minimum, the prestressing tendons, anchors, couplers and end fittings in exterior walls exposed to earth or weather must be protected. All other walls exposed to a mean relative humidity exceeding 75% must also employ some method of corrosion abatement. Unbonded tendons can be protected with galvanizing, epoxy coating, sheathing or other alternative method that provides an equivalent level of protection. Bonded tendons are protected from corrosion by the corrugated duct and prestressing grout in which they are encased.

DESIGN LOADS

As for other masonry structures, minimum required design loads are included in Minimum Design Loads for Buildings and Other Structures, ASCE 7 (ref. 5), or the governing building codes. If prestressing forces are intended to resist lateral loads from earthquake, a factor of 0.9 should be applied to the strength level prestress forces (0.6 for allowable stress design) as is done with gravity loads.

STRUCTURAL DESIGN

The design of post-tensioned masonry is based on allowable stress design procedures, except for laterally restrained tendons which use a strength design philosophy. Building Code Requirements for Masonry Structures (ref. 1) prescribes allowable stresses for unreinforced masonry in compression, tension and shear, which must be checked against the stresses resulting from applied loads.

The flexural strength of post-tensioned walls is governed by either the flexural tensile stress of the masonry (the flexural stress minus the post-tensioning and dead load stress), the masonry compressive stress, the tensile stress within the tendon, the shear capacity of the masonry or the buckling capacity of the wall.

Masonry stresses must be checked at the time of peak loading (independently accounting for both short-term and long-term losses), at the transfer of post-tensioning forces, and during the jacking operation when bearing stresses may be exceeded. Immediately after transfer of the post-tensioning forces, the stresses in the steel are the largest because long-term losses have not occurred. Further, because the masonry has had little time to cure, the stresses in the masonry will be closer to their capacity. Once long-term losses have transpired, the stresses in both the masonry and the steel are reduced. The result is a coincidental reduction in the effective capacity due to the prestressing force and an increase in the stresses the fully cured masonry can resist from external loads.

Effective Prestress

Over time, the level of prestressing force decreases due to creep and shrinkage of the masonry, relaxation of the prestressing tendons and potential decreases in the ambient temperature. These prestressing losses are in addition to seating and elastic shortening losses witnessed during the prestressing operation. In addition, the prestressing force of bonded tendons will decrease along the length of the tendon due to frictional losses. Since the effective prestressing force varies over time, the controlling stresses should be checked at several stages and loading conditions over the life of the structure.

The total prestress loss in concrete masonry can be assumed to be approximately 35%. At the time of transfer of the prestressing force, typical losses include: 1% seating loss + 1% elastic shortening = 2%. Additional losses at service loads and moment strength include:

relaxation3%
temperature10%
creep8%
CMU shrinkage7%
contingency5%
total33%

Prestress losses need to be estimated accurately for a safe and economical structural design. Underestimating losses will result in having less available strength than assumed. Overestimating losses may result in overstressing the wall in compression.

Effective Width

In theory, a post-tensioning force functions similarly to a concentrated load applied to the top of a wall. Concentrated loads are distributed over an effective width as discussed in the commentary on Building Code Requirements for Masonry Structures (ref. 1). A general rule-of-thumb is to use six times the wall thickness as the effective width.

Elastic shortening during post-tensioning can reduce the stress in adjacent tendons that have already been stressed. Spacing the tendons further apart than the effective width theoretically does not reduce the compressive stress in the effective width due to the post-tensioning of subsequent tendons. The applied loads must also be consolidated into the effective width so the masonry stresses can be determined. These stresses must be checked in the design stage to avoid overstressing the masonry.

Flexure

Tensile and compressive stresses resulting from bending moments applied to a section are determined in accordance with conventional elastic beam theory. This results in a triangular stress distribution for the masonry in both tension and compression. Maximum bending stress at the extreme fibers are determined by dividing the applied moment by the section modulus based on the minimum net section.

Net Flexural Tensile Stress

Sufficient post-tensioning force needs to be provided so the net flexural tensile stress is less than the allowable values. Flexural cracking should not occur if post-tensioning forces are kept within acceptable bounds. Flexural cracking due to sustained post-tensioning forces is believed to be more severe than cracking due to transient loading. Flexural cracks due to eccentric post-tensioning forces will remain open throughout the life of the wall, and may create problems related to water penetration, freeze-thaw or corrosion. For this reason, Building Code Requirements for Masonry Structures (ref. 1) requires that the net flexural tensile stress be limited to zero at transfer of the post-tensioning force and for service loadings with gravity loads only.

Axial Compression

Compressive stresses are determined by dividing the sum of the post-tensioning and gravity forces by the net area of the section. They must be less than the code prescribed (ref. 1) allowable values of axial compressive stress.

Walls must also be checked for buckling due to gravity loads and post-tensioning forces from unrestrained tendons. Laterally restrained tendons can not cause buckling; therefore only gravity compressive forces need to be checked for buckling in walls using laterally restrained tendons. Restraining the tendons also ensures that the tendons do not move laterally in the wall when the masonry deflects. The maximum compressive force that can be applied to the wall based upon ¼ buckling is Pe, per equation 2-11 of Building Code Requirements for Masonry Structures (ref. 1).

Combined Axial and Flexural Compressive Stress

Axial compressive stresses due to post-tensioning and gravity forces combine with flexural compressive stresses at the extreme fiber to result in maximum compressive stress. Conversely, the axial compressive stresses combine with the flexural tensile stresses to reduce the absolute extreme fiber stresses. To ensure the combination of these stresses does not exceed code prescribed allowable stresses, a unity equation is checked to verify compliance. Employing this unity equation, the sum of the ratios of applied-to-allowable axial and flexural stresses must be less than one. Unless standards (ref. 5) limit its use, an additional one-third increase in allowable stresses is permitted for wind and earthquake loadings, as is customary with unreinforced and reinforced masonry. Further, for the stress condition immediately after transfer of the post-tensioning force, a 20% increase in allowable axial and bending stresses is permitted by Building Code Requirements for Masonry Structures (ref. 1).

Shear

As with all stresses, shear stresses are resisted by the net area of masonry, and the wall is sized such that the maximum shear stress is less than the allowable stress. In addition, the compressive stress due to post-tensioning can be relied on to increase allowable shear stresses in some circumstances.

Post-Tensioning Tendons

The stress in the tendons is limited (ref. 1) such that:

  1. the stress due to the jacking force does not exceed 0.94fpy, 0.80fpu, nor that recommended by the manufacturer of the tendons or anchorages,
  2. the stress immediately after transfer does not exceed 0.82fpy nor 0.74fpu, and
  3. the stress in the tendons at anchorages and couplers does not exceed 0.78fpy nor 0.70fpu.

DETERMINATION OF POST-TENSIONING FORCES

Case (a) after prestress losses and at peak loading:

Assuming that the moment, M, due to wind or earthquake loadings is large relative to the eccentric load moment, the critical location will be at the mid-height of the wall for simply-supported walls, and the following equations apply (bracketed numbers are the applicable Building Code Requirements for Masonry Structures (ref. 1) equation or section numbers):

The 1.33 factor in Equation [2-10] represents the one- third increase in allowable stress permitted for wind and earthquake loadings. If the moment, M, is a result of soil pressures (as is the case for retaining walls), the 1.33 factor in Equation [2-10] must be replaced by 1.00.

Note that if the tendons are laterally restrained, Ppf should not be included in Equation [2-11].

(under the load combination of prestressing force and dead load only)

Additional strength design requirements for laterally restrained tendons:

Equation 4-3 above applies to members with uniform width, concentric reinforcement and prestressing tendons and concentric axial load. The nominal moment strength for other conditions should be determined based on static moment equilibrium equations.

Case (b) at transfer of post-tensioning:

Assuming that vertical live loads are not present during post-tensioning, the following equations apply. The worst case is at the top of the wall where post-tensioning forces are applied.

For cantilevered walls, these equations must be modified to the base of the wall.

If the eccentricity of the live load, Pl, is small, neglecting the live load in Equation [2-10] may also govern.

Case (c) bearing stresses at jacking:

Bearing stresses at the prestressing anchorage should be checked at the time of jacking. The maximum allowable bearing stress at jacking is 0.50f’mi per Building Code Requirements for Masonry Structures (ref. 1) section 4.9.4.2.

DESIGN EXAMPLE

Design a simply-supported exterior wall 12 ft (3.7 m) high for a wind load of 15 psf (0.72 kPa). The wall is constructed of concrete masonry units complying with ASTM C 90 (ref. 6). The units are laid in a full bed of Type S Portland cement lime mortar complying with ASTM C 270 (ref. 7). The specified compressive strength of the masonry (f’m) is 1,500 psi (10.3 MPa). The wall will be post-tensioned with 7/16 in. (11 mm) diameter laterally restrained tendons when the wall achieves a compressive strength of 1,250 psi (8.6 MPa). Axial load and prestress are concentric.

Given:
8 in. (203 mm) CMU
tf = 1.25 in. (32 mm)
f’m = 1,500 psi (10.3 MPa)
f’mi = 1,250 psi (8.6 MPa)
Fbt = 25 psi (0.17 MPa) (Type S Portland cement/lime mortar)
fpy = 100 ksi (690 MPa) (bars)
fpu = 122 ksi (840 MPa)
Aps = 0.14 in² (92 mm²)
Es = 29 x 106 psi (200 GPa)
Em = 900 f’m = 1.35 x 106 psi (9,300 MPa)
n = Es/Em = 21.5
d = 7.625/2 in. = 3.81 in. (97 mm) (tendons placed in the center of the wall)
unit weight of CMU wall = 39 psf (190 kg/m²) (ref. 13)

Loads: M = wh²/8 = (15)(12)²/8 = 270 ft-lb (366 N-m)
Pd at mid-height = (39)(12)/2 = 234 lb/foot of wall (3,410 N/m) (Pl = 0)

Maximum tendon stresses:
Determine governing stresses based on code limits (ref. 1):

At jacking:0.94 fpy = 94.0 ksi (648 MPa)
0.80 fpu = 97.6 ksi (673 MPa)
At transfer:0.82 fpy = 82.0 ksi (565 MPa)
0.74 fpu = 90.3 ksi (623 MPa)
At service loads:0.78 fpy = 78.0 ksi (538 MPa) ⇒ governs
0.70 fpu = 85.4 ksi (589 MPa)

Because the tendon’s specified tensile strength is less than 150 ksi (1,034 MPa), fps = fse (per ref. 1 section 4.5.3.3.4).

Prestress losses: Assume 35% total loss (as described in the Effective Prestress section above).

Tendon forces:
Determine the maximum tendon force, based on the governing tendon stress determined above for each case of jacking, transfer and service. At transfer, include 2% prestress losses. At service, include the full 35% losses.
Tendon capacity at jacking = 0.94 fpyAps = 13.3 kips (59 kN)
Tendon capacity at transfer = 0.82 fpyAps A x 0.98 = 11.4 kips (51 kN) (including transfer losses)
Tendon capacity at service = 0.78 fpyAps A x 0.65 = 7.2 kips (32 kN) (including total losses)

Try tendons at 48 in. (1,219 mm) on center (note that this tendon spacing also corresponds to the maximum effective prestressing width of six times the wall thickness).

Determine prestressing force, based on tendon capacity determined above:
at transfer: Ppi = 11.4 kips/4 ft = 2,850 lb/ft (41.6 kN/m)
at service: Ppf = 7.2 kips/4 ft = 1,800 lb/ft (26.3 kN/m)

Wall section properties: (ref. 8)
8 in. (203 mm) CMU with full mortar bedding:
An = 41.5 in.²/ft (87,900 mm²/m)
I = 334 in.4/ft (456 x 106 mm4/m)
S = 87.6 in.³/ft (4.71 x 106 mm³/m)
r = 2.84 in. (72.1 mm)

At service loads:
At service, the following are checked: combined axial compression and flexure using the unity equation (equation 2-10); net tension in the wall; stability by ensuring the compressive load does not exceed one-fourth of the buckling load, Pe, and shear and moment strength.

Check combined axial compression and flexure:

Check tension for load combination of prestress force and dead load only (per ref. 1 section 4.5.1.3):

Check stability:
Because the tendons are laterally restrained, the prestressing force, Ppf, is not considered in the determination of axial load ( per ref. 1 section 4.5.3.2), and the wall is not subject to live load in this case, so equation 2-11 reduces to:

Check moment strength:
Building Code Requirements for Masonry Structures section 4.5.3.3 includes the following criteria for moment strength of walls with laterally restrained tendons:

In addition, the compression zone must fall within the masonry, so a < tf.

where 1.3 and 1.2 are load factors for wind and dead loads, respectively.

At transfer:
Check combined axial compression and flexure using the unity equation (equation 2-10) and net tension in the wall.

Check tension for load combination of prestress force and dead load only (per ref. 1 section 4.5.1.3):

Therefore, use 7/16 in. (11 mm) diameter tendons at 48 in. (1,219 mm) o.c. Note that although wall design is seldom governed by out-of-plane shear, the shear capacity should also be checked.

NOTATIONS

An     net cross-sectional area of masonry section, in.² (mm²)
Aps   threaded area of post-tensioning tendon, in.² (mm²)
As     cross-sectional area of mild reinforcement, in.² (mm²)
a       depth of an equivalent compression zone at nominal strength, in. (mm)
b        width of section, in. (mm)
d       distance from extreme compression fiber to centroid of prestressing tendon, in. (mm)
Es      modulus of elasticity of prestressing steel, psi (MPa)
Em    modulus of elasticity of masonry, psi (MPa)
ed      eccentricity of dead load, in. (mm)
el       eccentricity of live load, in. (mm)
ep      eccentricity of post-tensioning load, in. (mm)
Fa     allowable masonry axial compressive stress, psi (MPa)
Fai    allowable masonry axial compressive stress at transfer, psi (MPa)
Fb     allowable masonry flexural compressive stress, psi (MPa)
Fbi    allowable masonry flexural compressive stress at transfer, psi (MPa)
Fbt    allowable flexural tensile strength of masonry, psi (MPa)
fa      axial stress after prestress loss, psi (MPa)
fai     axial stress at transfer, psi (MPa)
fb      flexural stress after prestress loss, psi (MPa)
fbi     flexural stress at transfer, psi (MPa)
f’m    specified compressive strength of masonry, psi (MPa)
f’mi   specified compressive strength of masonry at time of transfer of prestress, psi (MPa)
fps    stress in prestressing tendon at nominal strength, psi (MPa)
fpu    specified tensile strength of prestressing tendon, ksi (MPa)
fpy    specified yield strength of prestressing tendon, ksi (MPa)
fse     effective stress in prestressing tendon after all pre-stress losses have occurred, psi (MPa)
fy     specified yield strength of steel for reinforcement and anchors, psi (MPa)
h      masonry wall height, in. (mm)
I       moment of inertia of net wall section of extreme fiber tension or compression, in.4/ft (mm4/m)
M    moment due to lateral loads, ft-lb (N⋅m)
Mn   nominal moment strength, ft-lb (N⋅m)
Mu   factored moment due to lateral loads, ft-lb (N⋅m)
n      modular ratio of prestressing steel and masonry (Es/Em)
Pd    axial dead load, lb/ft (kN/m)
Pdu  factored axial dead load, lb/ft (kN/m)
Pe    Euler buckling load, lb/ft (kN/m)
Pl     axial live load, lb/ft (kN/m)
Plu    factored axial live load, lb/ft (kN/m)
Ppi    prestress force at transfer, lb/ft (kN/m)
Ppf    prestress force including losses, lb/ft (kN/m)
r       radius of gyration for net wall section, in. (mm)
S       section modulus of net cross-sectional area of the wall, in.³ /ft (mm³/m)
tf       face shell thickness of concrete masonry, in. (mm)
w     applied wind pressure, psf (kPa)
¤      strength reduction factor = 0.8

REFERENCES

  1. Building Code Requirements for Masonry Structures, ACI 530-02/ASCE 5-02/TMS 402-02. Reported by the Masonry Standards Joint Committee, 2002.
  2. Building Code Requirements for Structural Concrete, ACI 318-99. Detroit, MI: American Concrete Institute, Revised 1999.
  3. Construction of Post-Tensioned Concrete Masonry Walls, TEK 03-14. Concrete Masonry & Hardscapes Association, 2002.
  4. International Building Code. International Code Council, 2000.
  5. Minimum Design Loads for Buildings and Other Structures, ASCE 7-98, American Society of Civil Engineers, 1998.
  6. Standard Specification for Loadbearing Concrete Masonry Units, ASTM C 90-01a. American Society for Testing and Materials, 2001.
  7. Standard Specification for Mortar for Unit Masonry, ASTM C 270-01. American Society for Testing and Materials, 2001.
  8. Weights and Section Properties of Concrete Masonry Assemblies, CMU-TEC-002-23, Concrete Masonry & Hardscapes Association, 2023.
  9. Concrete Masonry Unit Shapes, Sizes, Properties, and Specifications, CMU-TEC-001-23, Concrete Masonry & Hardscapes Association, 2023.
  10. Mortars for Concrete Masonry, TEK 09-01A. Concrete Masonry & Hardscapes Association, 2001.
  11. Grout for Concrete Masonry, TEK 09-04. Concrete Masonry & Hardscapes Association, 2005.
  12. Steel for Concrete Masonry Reinforcement, TEK 12-04D. Concrete Masonry & Hardscapes Association, 1998.
  13. Weights and Section Properties of Concrete Masonry Assemblies, CMU-TEC-002-23, Concrete Masonry & Hardscapes Association, 2023.
  14. Schultz, A.E., and M.J. Scolforo, An Overview of Prestressed Masonry, TMS Journal, Vol. 10, No. 1, August 1991, pp. 6-21.
  15. Schultz, A.E., and M.J. Scolforo, Engineering Design Provisions for Prestressed Masonry, Part 1: Masonry Stresses, Part 2: Steel Stresses and Other Considerations, TMS Journal, Vol. 10, No. 2, February 1992, pp. 29-64.
  16. Standard Specification for Steel Strand, Uncoated Seven-Wire for Prestressed Concrete, ASTM A 416-99. American Society for Testing and Materials, 1999.
  17. Standard Specification for Uncoated Stress-Relieved Steel Wire for Prestressed Concrete, ASTM A 421-98a. American Society for Testing and Materials, 1998.
  18. Standard Specification for Uncoated High-Strength Steel Bar for Prestressed Concrete, ASTM A 722-98. American Society for Testing and Materials, 1998.
  19. Standard Specification for Compressible-Washer-Type Direct Tension Indicators for Use with Structural Fasteners, ASTM F 959-01a. American Society for Testing and Materials, 2001.

ASD Tables for Reinforced Concrete Masonry Walls Based on the 2012 IBC & 2011 MSJC

INTRODUCTION

The combination of concrete masonry and steel reinforcement provides a strong structural system capable of resisting large compressive and flexural loads. Reinforced masonry structures have significantly higher flexural strength and ductility than similarly configured unreinforced structures and provide greater reliability in terms of expected load carrying capacity at failure.

Concrete masonry elements can be designed using several methods in accordance with the International Building Code (IBC, ref. 1) and, by reference, Building Code Requirements for Masonry Structures (MSJC Code, ref. 2): allowable stress design, strength design, direct design, empirical design, or prestressed masonry. The design tables in this TEK are based on allowable stress design provisions.

The content presented in this edition of TEK 14-19B is based on the requirements of the 2012 IBC (ref. 1a), which in turn references the 2011 edition of the MSJC Code (ref. 2a). For designs based on the 2006 or 2009 IBC (refs. 1b, 1c), which reference the 2005 and 2008 MSJC (refs. 2b, 3c), respectively, the reader is referred to TEK 14-19B (ref. 3).

Significant changes were made to the allowable stress design (ASD) method between the 2009 and 2012 editions of the IBC. These are described in detail in TEK 14-07C, ASD of Concrete Masonry (2012 IBC & 2011 MSJC) (ref. 4), along with a detailed presentation of all of the allowable stress design provisions of the 2012 IBC.

LOAD TABLES

Tables 1 and 2 list the maximum bending moments and shears, respectively, imposed on walls simply supported at the top and bottom and subjected to uniform lateral loads with no applied axial loads.

WALL CAPACITY TABLES

Tables 3, 4, 5 and 6 contain the maximum bending moments and shear loads that can be sustained by 8-, 10-, 12-, and 16-in. (203-, 254-, 305-, 406 mm) walls, respectively, without exceeding the allowable stresses defined in the 2012 IBC and 2011 MSJC (refs. 1a, 2a). These wall strengths can be compared to the loads in Tables 1 and 2 to ensure the wall under consideration has sufficient design capacity to resist the applied load.

The values in Tables 3 through 6 are based on the following criteria:

  • Maximum allowable stresses:
  • f’m = 1500 psi (10.3 MPa)
  • Em  = 900f’m or 1,350,000 psi (9,310 MPa)
  • Es = 29,000,000 psi (200,000 MPa)
  • Type M or S mortar
  • running bond or bond beams at 48 in. (1,219 mm) max o.c.
  • reinforcement spacing does not exceed the wall height
  • only cores containing reinforcement are grouted.

Reinforcing Steel Location

Two sets of tables are presented for each wall thickness. Tables 3a, 4a, 5a and 6a list resisting moment and resisting shear values for walls with the reinforcing steel located in the center of the wall. Centered reinforcing bars are effective for providing tensile resistance for walls which may be loaded from either side, such as an above grade exterior wall which is likely to experience both wind pressure and suction.

Tables 3b, 4b, 5b and 6b list resisting moment and resisting shear values for walls with the reinforcing steel offset from the center.

Placing the reinforcement farther from the compression face of the masonry provides a larger effective depth of reinforcement, d, and correspondingly larger capacities. A single layer of off-center reinforcement can be used in situations where the wall is loaded from one side only, such as a basement wall with the reinforcement located towards the interior. For walls where loads can be in both directions (i.e. pressure or suction), two layers of reinforcement are used: one towards the wall exterior and one towards the interior to provide increased capacity under both loading conditions. In Tables 3b, 4b, 5b and 6b, the effective depth of reinforcement, d, is a practical value which takes into account construction tolerances and the reinforcing bar diameter.

Figure 1 illustrates the two steel location cases.

  
  

  
  

  
  

DESIGN EXAMPLE

A warehouse wall will span 34 ft (10.4 m) between the floor slab and roof diaphragm. The walls will be constructed using 12 in. (305 mm) concrete masonry units. What is the required reinforcing steel size and spacing to support a wind load of 20 psf (0.96 kPa)?

From interpolation of Tables 1 and 2, respectively, the wall must be able to resist:
M = 34,800 lb-in./ft (12.9 kN-m/m)
V = 340 lb/ft (4.96 kN/m)

Assuming the use of offset reinforcement, from Table 5b, No. 6 bars at 40 in. on center (M#19 at 1,016 mm) or No. 7 bars at 48 in. (M#22 at 1,219 mm) on center provides sufficient strength: for No. 6 bars at 40 in. o.c. (M#19 at 1,016 mm):
Mr = 35,686 lb-in./ft (13.3 kN-m/m) > M             OK
Vr = 2,299 lb/ft (33.5 kN/m) > V                           OK

for No. 7 bars at 48 in. (M#22 at 1,219 mm) :
Mr = 40,192 lb-in./ft (14.9 kN-m/m) > M              OK
Vr = 2,133 lb/ft (31.1 kN/m) > V                            OK

As discussed above, since wind loads can act in either direction, two bars must be provided in each cell when using off-center reinforcement—one close to each faceshell.

Alternatively, No. 6 bars at 24 in (M#19 at 610 mm) or No. 8 at 40 in (M#25 at 1,016 mm) could have been used in the center of the wall.

NOTATION

As   = area of nonprestressed longitudinal reinforcement, in.² (mm²)
b     = effective compressive width per bar, in. (mm)
d     = distance from extreme compression fiber to centroid of tension reinforcement, in. (mm)
Em  = modulus of elasticity of masonry in compression, psi (MPa)
Es   = modulus of elasticity of steel, psi (MPa)
Fb   = allowable compressive stress available to resist flexure only, psi (MPa)
Fs    = allowable tensile or compressive stress in reinforcement, psi (MPa)
Fv    = allowable shear stress, psi (MPa)
f’m   = specified compressive strength of masonry, psi (MPa)
M    = maximum calculated bending moment at section under consideration, in.-lb, (N-mm)
Mr   = flexural strength (resisting moment), in.-lb (N-mm)
V     = shear force, lb (N)
Vr    = shear capacity (resisting shear) of masonry, lb (N)

REFERENCES

  1. International Building Code. International Code Council.
    • 2012 Edition
    • 2009 Edition
    • 2006 Edition
  2. Building Code Requirements for Masonry Structures. Reported by the Masonry Standards Joint Committee.
    • 2011 Edition: TMS 402-11/ACI 530-11/ASCE 5-11
    • 2008 Edition: TMS 402-08 /ACI 530-08/ASCE 5-08
    • 2005 Edition: ACI 530-05/ASCE 5-05/TMS 402-05
  3. Allowable Stress Design Tables for Reinforced Concrete Masonry Walls, TEK 14-19B. Concrete Masonry & Hardscapes Association, 2009.
  4. Allowable Stress Design of Concrete Masonry Based on the 2012 IBC & 2011 MSJC, TEK 14-07C. Concrete Masonry & Hardscapes Association, 2011.

 

Strength Design of Concrete Masonry Walls for Axial Load & Bending

INTRODUCTION

Building structural design requires a variety of structural loads to be accounted for: dead and live loads, those from wind, earthquake, lateral soil pressure, lateral fluid pressure as well as forces induced by temperature changes, creep, shrinkage and differential movements. Because any load can act simultaneously with another, the designer must consider how these various loads interact on the wall. For example, an axial load can offset tension due to lateral load, thereby increasing flexural capacity, and, if acting eccentrically, can also increase the moment on the wall. Building codes dictate which load combinations must be considered, and require that the structure be designed to resist the most severe load combination.

The design aids in this TEK cover combined axial compression or axial tension and flexure, as determined using the strength design provisions of Building Code Requirements for Masonry Structures (ref. 3). For concrete masonry walls, these design provisions are outlined in Strength Design of Concrete Masonry (ref. 1). Axial load-bending moment interaction diagrams account for the interaction between moment and axial load on the design capacity of a wall. This TEK shows the portion of the interaction diagram that applies to the majority of wall designs. Although negative moments are not shown, the figures may be used for these conditions, since with reinforcement in the center of the wall wall strength will be the same under either a positive or negative moment of the same magnitude. Conditions outside of this area may be determined using Concrete Masonry Wall Design Software or Concrete Masonry Design Tables (refs. 4, 5). The reader is referred to Loadbearing Concrete Masonry Wall Design (ref. 2) for a full discussion of interaction diagrams.

Figures 1 through 8 apply to fully or partially grouted reinforced concrete masonry walls with a specified compressive strength f’m of 1,500 psi (10.34 MPa), and a maximum wall height of 20 ft (6.10 m), Grade 60 (414 MPa) vertical reinforcement, with reinforcing bars positioned in the center of the wall and reinforcing bar spacing s from 8 in. to 120 in. ( 203 to 3,048 mm).  Figures 1 through 8 apply to fully or partially grouted reinforced concrete masonry walls with a specified compressive strength, f’m, of 1500 psi (10.34 MPa), and a maximum wall height of 20 ft (6.09 m), Grade 60 vertical reinforcement, with reinforcing bars positioned in the center of the wall and reinforcing bar spacing, s, from 8 in. to 120 in. ( 203 to 3,048 mm). Each figure applies to one specific wall thickness and one reinforcing bar size. For walls less than 20 ft (6.1 m) high, figures 1 through 8 will be slightly conservative due to PΔ effects.

DESIGN EXAMPLE

An 8-in. (203-mm) thick, 20 ft (6.10 m) high reinforced simply supported concrete masonry wall (115 pcf (1,842 kg/m³)) is to be designed to resist wind load as well as eccentrically applied axial live and dead loads as depicted in Figure 9. The designer must determine the reinforcement size spaced at 24 in. (610 mm) required to resist the applied loads, listed below.

D = 520 lb/ft (7.6 kN/m), at e = 0.75 in. (19 mm)
L = 250 lb/ft (3.6 kN/m), at e = 0.75 in. (19 mm)
W = 20 psf (1.0 kPa)

The maximum moment due to the wind load is determined as follows.

The axial load used for design is the axial load at the location of maximum moment. This combination may not necessarily be the most critical section for combined axial load and flexure, but should be close to the critical location. The wall weight is estimated to be halfway between fully grouted and hollow (82 and 38.7 psf (400 and 189 kg/m2), respectively, for 115 pcf (1842 kg/m3) unit concrete density).

The eccentricity of the axial loads also induces bending in the wall and should be included in the applied moment. The magnitude of the moment due to the eccentric axial load must be found at the same location as the maximum moment.

During design, all load combinations should be checked, with the controlling load case used for design. For brevity, only the two combinations above will be evaluated here, since the axial load actually increases the flexural capacity for the first combination by offsetting tension in the wall due to the lateral load.

Figure 2 shows that No. 4 bars at 24 in. (M #13 at 610 mm) on center are adequate. If a larger bar spacing is desired, No. 5 at 32 in. (M #16 at 813 mm) or No. 6 at 48 in. (M #19 at 1219 mm) will also meet the design requirements. Although wall design is seldom governed by out-of-plane shear, the shear capacity should be checked. In addition, the axial load should be recalculated based on the actual wall weight (based on grout spacing chosen), then the resulting required capacity should be recalculated and plotted on the interaction diagram to check adequacy.

NOMENCLATURE

D        dead load, lb/ft (kN/m)
e        eccentricity of axial load – measured from centroid of wall, in. (mm)
f’m      specified masonry compressive strength, psi (MPa)
h         height of wall, in. (mm)

L         live load, lb/ft (kN/m)
Lr        roof live load, lb/ft (kN/m)

Mu     factored moment, in.-lb/ft or ft-lb/ft (kN⋅m/m)
Pu      factored axial load, lb/ft (kN/m)

s        spacing of vertical reinforcement, in. (mm)
W      wind load, psf (kN/m²)
y distance measured from top of wall, ft (m)

REFERENCES

  1. Strength Design of Concrete Masonry, TEK 14-04B. Concrete Masonry & Hardscapes Association, 2002.
  2. Loadbearing Concrete Masonry Wall Design, TEK 14-05A. Concrete Masonry & Hardscapes Association, 2006.
  3. Building Code Requirements for Masonry Structures, ACI 530-02/ASCE 5-02/TMS 402-02. Reported by the Masonry Structures Joint Committee, 2002.
  4. Concrete Masonry Wall Design Software, CMS-10. Concrete Masonry & Hardscapes Association, 2002.
  5. Concrete Masonry Design Tables, TR 121A. Concrete Masonry & Hardscapes Association, 2000.
  6. Minimum Design Loads for Buildings and Other Structures, ASCE 7-02. American Society of Civil Engineers, 2002.
   
   
   
   
   
   
   
   

Strength Design Provisions for Concrete Masonry

INTRODUCTION

Concrete masonry elements can be designed using one of several methods in accordance with Building Code Requirements for Masonry Structures (ref. 1): empirical design, strength design or allowable stress design. This TEK provides a basic overview of design criteria and requirements for concrete masonry structures designed using the strength design provisions contained in Chapter 3 of the 2002 edition of Building Code Requirements for Masonry Structures (also referred to as the MSJC Code) (ref. 1) as referenced and modified in Section 2108 of the 2003 International Building Code (IBC) (ref. 2). In addition, changes to the strength design method incorporated into the 2005 edition of the MSJC Code (ref. 3) through Section 2108 of the 2006 International Building Code (ref. 4) are also reviewed, as are modifications included in the 2008 MSJC Code (ref. 5).

For empirical and allowable stress design requirements, the user is referred to TEK 14-08B, Empirical Design of Concrete Masonry Walls (ref. 6), and TEK 14-07C, ASD of Concrete Masonry (2012 IBC & 2011 MSJC) (ref. 7), respectively. Tables, charts, and additional design aids specific to the design of various concrete masonry elements can be found in other related TEK.

Strength design is based on the following design assumptions in conjunction with basic principles of engineering mechanics (refs. 1, 3, 5), as shown in Figure 1 for a reinforced element:

  • Plane sections before bending remain plane after bending. Therefore, strain in the masonry and in reinforcement, if present, is directly proportional to the distance from the neutral axis.
  • For unreinforced masonry, the flexural stresses in the masonry are assumed to be directly proportional to strain. For reinforced masonry, the tensile strength of the masonry is neglected when calculating flexural strength, but considered when calculating deflection.
  • The units, mortar, grout and reinforcement for reinforced masonry act compositely to resist applied loads.
  • The nominal strength of masonry cross-sections for combined flexure and axial load is based on applicable conditions of equilibrium.
  • The maximum masonry compressive stress is 0.80f’m for both reinforced and unreinforced masonry.
  • The maximum usable strain, εmu, at the extreme compression fiber of concrete masonry is 0.0025.
  • For reinforced masonry, compression and tension stresses in the reinforcement below the specified yield strength, fy, are taken equal to the modulus of elasticity of the reinforcement, Es, times the steel strain εs. For strains greater than the yield strain corresponding to fy, stress in the reinforcement is taken equal to fy.
  • For reinforced masonry, the compressive stress is rectangular and uniformly distributed over an equivalent compression zone, bounded by the compression face of the masonry with a depth of a = 0.80c.

Based on the prescribed design model outlined above, the internal distribution of stresses and strains is illustrated in Figure 1 for a reinforced masonry element. A more comprehensive review of the design model is provided in Masonry Structures, Behavior and Design (ref. 8).

2003 IBC STRENGTH DESIGN MODIFICATIONS

The 2003 IBC adopts the 2002 MSJC Code with two modifications specific to the strength design procedure in IBC Section 2108. The two modifications are as follows.

  • Section 2108.2 introduces a maximum effective compression width for out-of-plane bending of six times the nominal wall thickness, not to exceed the reinforcement spacing. This is similar to limits historically used by the allowable stress design provisions in the MSJC Code as well as those adopted into the 2005 MSJC Code for strength design, as reviewed below.
  • Welded and mechanical splices incorporated into masonry elements designed by the strength design method must also comply with Section 2108.3 of the 2003 IBC. For welded splices, the reinforcement to be welded must comply with ASTM A 706 (ref. 9). Splicing by mechanical connectors is classified as either Type 1 or Type 2 mechanical splices in accordance with ACI 318, Building Code Requirements for Structural Concrete (ref. 10). Type 1 mechanical splices are only required to develop 125 percent of the nominal yield strength of the reinforcement being spliced. Type 2 mechanical splices, conversely, must develop the full specified tensile strength of the reinforcement. Welded and Type 1 mechanical splices are not permitted to be used in the plastic hinge region of intermediate or special reinforced masonry shear walls.

2002 MSJC CODE STRENGTH DESIGN CRITERIA

Using strength design, the design strength of a masonry element is compared to the required (or factored) strength (indicated by the subscript u), which includes load factors to account for the uncertainty in predicting design loads and the probability of more than one design load acting simultaneously. The required strength is based on the strength design load combinations as required by Section 1605 of the IBC. At the option of the designer, or when the MSJC Code is used in conjunction with another building code that does not contain load combinations, masonry structures are designed to resist the load combination specified in ASCE 7, Minimum Design Loads for Buildings and Other Structures (ref. 11). For strength design, these load combinations are effectively the same.

The design strength of masonry is the nominal strength (indicated by the subscript n) multiplied by an appropriate strength reduction factor, Φ. The design is acceptable when the design strength equals or exceeds the factored strength (i.e., when ΦMnMu) for all prescribed load combinations. The following sections cover the general strength design requirements applicable to both unreinforced and reinforced masonry assemblies, with the exception of design requirements for anchor bolts and lap splices. For these topics, the user is referred to TEK 12-03C, Design of Anchor Bolts Embedded in Concrete Masonry (ref. 12) and TEK 12-06A, Splices, Development and Standard Hooks for CM Based on the 2009 & 2012 IBC (ref. 13), respectively.

Strength Reduction Factors

To account for uncertainties in construction, material properties, calculated versus actual strengths and anticipated failure modes, the nominal strength of a masonry element is multiplied by an appropriate strength reduction factor, Φ. Strength reduction factors are used in conjunction with the load factors applied to the design loads. The values of the strength reduction factors for various types of loading conditions are:

  • for reinforced masonry elements subjected to flexure or axial loads; Φ = 0.90;
  • for unreinforced masonry elements subjected to flexure or axial loads; Φ = 0.60;
  • for masonry elements subjected to shear loads; Φ = 0.80;
  • for bearing on masonry elements; Φ = 0.60.

Drift and Deflection

When designing for earthquakes, the story drift (the relative displacement of adjacent stories) must be checked against the IBC prescribed allowable story drifts. When the MSJC Code is used in conjunction with a building code that does not contain allowable story drifts, the provisions of ASCE 7 are used. For masonry buildings with cantilevered shear walls, the IBC limits the story drift to 0.01hsx, where hsx is the height of the story below the level for which the drift is being calculated. For other types of masonry shear wall buildings, except masonry frames, the allowable story drift is limited to 0.007hsx. While the IBC includes story drift limits for masonry frame wall buildings, such structural configurations are rarely used. When calculating story drift, the calculated elastic deflection is multiplied by the deflection amplification factor, Cd, as prescribed in the IBC for the type of structural system being designed. The deflection amplification factor approximates the additional deflection due to inelastic response (if applicable) of the system during an earthquake. Due to the inherent in-plane stiffness of masonry assemblies, in-plane deflection and story drift are rarely a controlling limit unless a relatively large number of openings is incorporated that reduces the strength and stiffness along a line of lateral resistance.

Unlike allowable stress design, which permits deflections to be calculated assuming uncracked sections for both reinforced and unreinforced masonry, strength design requires that deflections of reinforced masonry elements be based on cracked section properties, which are limited to one-half of the gross section properties unless a rigorous cracked section analysis is performed. The deflection of unreinforced masonry elements, which are required to remain uncracked, use uncracked section properties.

Because unreinforced masonry elements must be designed to remain uncracked, deflection is rarely a controlling design limit for these systems. Reinforced masonry elements, however, particularly tall, slender walls bending in the out-of-plane direction, may exhibit excessive deflection even at relatively low applied loads. As such, the MSJC Code limits the mid-height deflection, δs, of reinforced masonry elements bending in the out-of-plane direction due to service level lateral and axial loads to 0.007h. Second order effects due to P-delta contributions must also be taken into account, which is usually accomplished through iteration until convergence is achieved.

When the applied moment, Mser, is less than the moment required to cause cracking, Mcr, (Mser Mcr) then the mid-height deflection of a masonry element subjected to a uniform out-of-plane load can be determined using Equation 1.

Conversely, when the applied moment, Mser, is greater than the moment required to cause cracking, Mcr, but less than the nominal moment strength of the assembly (Mcr < Mser < Mn) the mid-height deflection of a masonry element subjected to a uniform out-of-plane load can be determined using Equation 2.

The MSJC does not prescribe a method of determining the cracked moment of inertia, Icr. As such, any rational method of determining cracked section properties is permitted. CMUTEC-002-23, Weights and Section Properties of Concrete Masonry Assemblies (ref. 14), provides typical section properties for various uncracked wall sections. For use in Equations 1 and 2, the cracking moment can be taken as:

Where the modulus of rupture, fr, is obtained from Table 1 for the type of mortar and construction under consideration.

Material Properties

Due to the lack of available research data substantiating its use, the specified compressive strength of concrete masonry, f’m, designed by the strength design method is required to be at least 1,500 psi (10.34 MPa), but not larger than 4,000 psi (27.58 MPa). In addition, when used in a structural role, the specified compressive strength of grout is required to be at least equal to the specified compressive strength of concrete masonry, but not greater than 5,000 psi (34.47 MPa). For each of these upper limits on masonry assembly or grout compressive strength, the actual tested strength is permitted to exceed these values: the restriction applies only to specified strengths upon which the design is based. Note that these provisions are included in the 2005 MSJC Code as well.

Strength design of reinforced masonry is based on the specified yield strength of reinforcement, fy, which is limited to 60,000 psi (413.7 MPa). The actual yield strength of the reinforcement is limited to 1.3 times the specified yield strength. The combination of these requirements effectively precludes the use of bed joint reinforcement to be used as primary structural steel in masonry designed by the strength design method, because the nominal yield strength of bed joint reinforcement exceeds these limits. The compressive resistance of steel reinforcement is not permitted to be used unless lateral reinforcement is provided in compliance with Chapter 2 of the MSJC Code, except as permitted when checking the maximum reinforcement limits as described later.

Unreinforced Masonry

For unreinforced masonry, the masonry assembly (units, mortar and grout, if used) is designed to carry all applied stresses. The additional capacity from the inclusion of reinforcing steel, if present (such as reinforcement added to control shrinkage cracking or prescriptively required by the code), is neglected when designing unreinforced masonry elements. Because the masonry resists both tension and compression stresses resulting from applied loads, the masonry must be designed to remain uncracked.

Unreinforced Nominal Flexural Strength

The nominal flexural tensile strength of unreinforced concrete masonry is given by the modulus of rupture as prescribed in the MSJC Code, which varies with the direction of span, mortar type, bond pattern and percentage of grouting as shown in Table 1. These values apply to masonry subject to out-of-plane bending. For walls spanning horizontally between supports, the code conservatively assumes that stack bond masonry has no flexural bond strength across the mortared head joints, thus only the grout area (for horizontally grouted sections) is used. For this case, the modulus of rupture of the grout is taken equal to 250 psi (1720 kPa). Likewise, for masonry subjected to in-plane bending, the modulus of rupture normal and parallel to the bed joints is taken as 250 psi (1720 kPa).

For masonry elements subjected to a factored bending moment, Mu, and a compressive axial force, Pu, the resulting flexural bending stress is determined using Equation 4.

If the resulting value of Fu is positive, then the masonry section is controlled by tension and the modulus of rupture values of Table 1, reduced by the appropriate strength reduction factor (Φ = 0.60), must be satisfied. Conversely, if Fu as given by Equation 4 is negative, the masonry section is in compression and the design compressive stress of 0.80f’m applies. When using axial load to offset flexural bending stresses as described above, only dead loads or other permanent loads should be included in Pu.

Unreinforced Nominal Axial Strength

When unreinforced masonry walls are subjected to compressive axial loads only, the nominal axial compressive strength, Pn, is determined using equation 5 or 6, as appropriate. Unreinforced masonry is not permitted to carry net axial tension forces.

For elements with h/r not greater than 99:

For elements with h/r greater than 99:

Unreinforced Nominal Shear Strength

Shear stresses on unreinforced masonry elements are calculated using the net cross-sectional properties of the masonry in the direction of the applied shear force using:

Equation 7 is applicable to determining both in-plane and out-of-plane shear stresses. Because unreinforced masonry is designed to remain uncracked, it is not necessary to perform a cracked section analysis to determine the net section properties. In turn, the applied shear stresses (factored accordingly for the appropriate load combination) are compared to the nominal shear strength, Vn, of an unreinforced masonry section, which is the least of:

Reinforced Masonry

The design of reinforced masonry in accordance with the MSJC Code neglects the tensile resistance provided by the masonry units, mortar and grout in determining the strength of the masonry assemblage. (The tensile strength of the units, mortar, and grout is considered, however, in determining the stiffness and deflection of a reinforced masonry element.) Thus, for design purposes, the portion of masonry subject to net tensile stress is assumed to have cracked, transferring all tensile forces to the reinforcement.

Using strength design, reinforcing bars used in masonry may not be larger than No. 9 (M #29) and bars may not be bundled. Further, the nominal bar diameter is not permitted to exceed one-eighth of the nominal member thickness or one-quarter of the least clear dimension of the cell, course or collar joint in which it is placed. The total area of reinforcing bars placed in a single cell or in a course of hollow unit construction may not exceed 4% of the cell area. Note that this limit does not apply at sections where lap splices occur. At lap splices, the maximum reinforcing bar area is increased to 8%, in both the 2002 and 2005 editions of the MSJC Code.

Maximum Flexural Reinforcement Ratio

To provide for a prescribed level of reinforced masonry ductility in the event of failure, the maximum reinforcement ratio, ρmax, is limited in accordance with Equation 8 or 9, as appropriate. Equation 8 applies to masonry cross sections that are fully grouted or where the neutral axis falls within the face shell of the masonry units in partially grouted construction. When the neutral axis falls within the cores of partially grouted construction, Equation 9 is used.

The tension reinforcement yield strain factor, α, varies with the seismic response modifi cation factor, R, masonry element, and type of loading as follows:

  1. α = 1.3 for walls subjected to out-of-plane forces and designed using an R value greater than 1.5,
  2. α = 5.0 for walls subjected to in-plane forces, for columns and for beams designed using an R > 1.5,
  3. α = 2.0 for masonry structures designed using an R ≤ 1.5.

In the above set of requirements, R is larger for out-of-plane loads when R is less than or equal to 1.5, which is contrary to the underlying intent of providing increased ductility for systems and elements whose ductility demand may be relatively high. Several updates and revisions to the maximum have been incorporated into subsequent editions to the 2002 MSJC Code as reviewed below.

Reinforced Nominal Axial Strength

The nominal axial strength, Pu, of masonry walls, piers and columns, modified to account for the effects of slenderness, is determined using equation 10 or 11, as appropriate. The MSJC Code also limits the factored axial stress to 0.20f’m.

For elements with h/r not greater than 99:

For elements with h/r greater than 99:

Note that the reinforcing steel area, As, is included in the nominal axial strength calculation only if it is laterally confined in accordance with Chapter 2 of the MSJC Code.

Reinforced Nominal Flexural Strength

The nominal flexural strength, Mn, of a reinforced masonry element is determined as outlined below. In addition, the nominal flexural strength at any section along a member must be at least one-fourth of the maximum nominal flexural strength at the critical section.

When there are no axial loads, or when axial loads are conservatively neglected as may be appropriate in some cases, there are several circumstances to consider when determining the nominal flexural strength of reinforced masonry walls. For a fully grouted element, the internal moment arm between the resulting compressive and tensile forces is resolved to determine the resisting capacity of the section. Partially grouted walls are analyzed in the same way, but with the additional consideration of the possible influence of the ungrouted cores on the strength of the section. For partially grouted masonry bending out-of-plane, there are two types of behavior to consider.

  1. In the first case, when the neutral axis (the location of zero stress) lies within the compression face shell, the wall is analyzed and designed using the procedures for a fully grouted wall.
  2. In the second case, when the neutral axis lies within the core area, the portion of the ungrouted cells must be deducted from the area of masonry capable of carrying compression stresses.

The neutral axis location depends on the reinforcement spacing as well as the reinforcement ratio, ρ, and the distance between the reinforcement and the extreme compression fiber, d.

When analyzing partially grouted walls, it is typically initially assumed that the neutral axis lies within the compression face shell, as the analysis is more straightforward. The value of c is then calculated based on this assumption. If it is determined that the neutral axis lies outside the compression face shell, the more rigorous tee beam analysis is performed. Otherwise, the rectangular beam analysis is carried out.

Rectangular Beam Analysis

For fully grouted masonry elements and for partially grouted masonry walls with the neutral axis in the compression face shell, the nominal flexural strength, Mn, is calculated using equations 12 and 13 as follows:

Tee Beam Analysis

For partially grouted masonry walls where the neutral axis is located within the cores, the nominal flexural strength, Mn, is calculated using equations 14, 15, and 16 as follows:

To account for deflection resulting from out-of-plane loads and the additional bending moment due to eccentrically applied axial loads, the factored bending moment at the mid-height of a simply supported wall under uniform loading is required to be determined by Equation 17.

Multiple iterations may be necessary to determine the converging value of the resulting deflection.

Reinforced Nominal Shear Strength

Shear acting on reinforced masonry members is resisted by the masonry and shear reinforcement, if provided, in accordance with the following:

Where Vn may not exceed the values given by Equations 20 or 21, as appropriate.

Where M/Vdv is less than or equal to 0.25:

Where M/Vdv is greater than 1.00:

For values of M/Vdv between 0.25 and 1.00, the maximum value of Vn may be linearly interpolated.

The nominal shear strength provided by the masonry is determined in accordance with the following.

Where the value of M/Vdv need not be taken greater than 1.0.

When shear reinforcement is incorporated into reinforced masonry construction, the shear strength provided by the reinforcement is calculated in accordance with the following.

2006 IBC REQUIREMENTS

The following is a brief summary of the changes and modifications adopted into the 2005 MSJC Code as referenced in the 2006 IBC. The majority of the basic design provisions remain unchanged, with several notable exceptions as reviewed below.

2006 IBC Strength Design Modifications

The same modifications for welded and mechanical splices previously discussed in the 2003 IBC Strength Design Modifications section remain in the 2006 IBC. In addition, Section 2108 of the 2006 IBC has incorporated a limit of 72db on the maximum required length of lap splices used in masonry designed by the strength design method as determined by the 2005 MSJC lap splice provisions. While this limit is typically not triggered for relatively small bar diameters, it can reduce the required lap length of large diameter reinforcing bars or in cases where small cover distances are used. Refer to TEK 12-06A (ref. 13) for detailed splice length requirements.

Other changes to Section 2108 of the 2006 IBC reflect updates and modifications to the 2005 MSJC Code to remove redundant or conflicting requirements.

2005 MSJC Code Strength Design Criteria Bearing Strength

Due primarily to oversight, the 2002 MSJC Code did not include explicit provisions for determining the bearing strength of masonry subjected to concentrated loads. The bearing strength requirements adopted into the 2005 MSJC Code are similar to those used for allowable stress design, modified as necessary for use in strength design. The nominal bearing strength of masonry is taken as the greater of Equations 24 and 25:

The value of 0.6 in Equations 24 and 25 is a design coefficient, not the strength reduction factor, Φ, which also happens to be equal to 0.6 for determining the design bearing strength. For use in Equation 25, the following expression must also be satisfied:

Where A1 is the actual bearing area directly under the bearing plate and A2 is the effective bearing area, defined as a right o pyramid having A1 as a top plane and sloping at 45 . For walls laid in stack bond, the area defined by A2 is terminated at head joints unless a bond beam or other means of distributing stresses is used. The concentrated stresses are distributed over a length equal to the width of the bearing area plus four times the thickness of the supporting element, but not to exceed the center-to-center distance between the concentrated loads.

Modulus of Rupture

The 2005 MSJC Code incorporated a few modifications to the modulus of rupture values presented in Table 1 for flexural tension stresses normal to the bed joints of hollow, fully grouted assemblies, as shown in Table 2. These modifications recognize that the type of mortar has less of an influence on the modulus of rupture when the element is fully grouted.

In addition, instead of prescribing a single value of 250 psi (1.72 MPa) for masonry subjected to in-plane bending as in the 2002 MSJC Code, the 2005 MSJC Code requires the use of Table 1, as modified by Table 2, for all cases of in-plane and out-of-plane bending.

Maximum Flexural Reinforcement Limits

The maximum flexural reinforcement requirements in the 2005 MSJC employ the same strain gradient approach as reviewed above, with several notable revisions.

  • Reinforcement is permitted to carry compression stresses, with or without lateral confining reinforcement. Further, the 1.25 factor on the nominal yield stress of the reinforcement as been removed. The resulting equations for the maximum flexural reinforcement limits are given by Equation 26 for fully grouted members or members subjected to in-plane loads, and Equation 27 for partially grouted masonry assemblies subjected to out-of-plane loads where the neutral axis falls in an open cell.
  • The tension reinforcement yield strain factor, α, is taken equal to the following values when Mu/Vudv ≥ 1.0 (indicating performance is dominated by flexure):
    • for intermediate reinforced masonry shear walls subjected to in-plane loads: α = 3.0,
    • for special reinforced masonry shear walls subjected to in-plane loads: α = 4.0,
    • for other masonry elements and loading conditions: α= 1.5.
  • For cases where Mu/Vudv ≤ 1.0 (indicating shear dominated performance) and R > 1.5: α = 1.5.
  • For masonry members where Mu/Vudv ≤ 1.0 and R ≤ 1.5, there is no limit on the amount of reinforcement permitted.
  • The axial load, P, for use in Equations 26 and 27, is the gravity axial load component determined from the following load combination: D + 0.75L + 0.525QE.

Effective Compression Width per Reinforcing Bar

Section 2108 of the 2003 IBC included limits on the effective compression width per reinforcing bar, which were subsequently removed from the 2006 IBC, as similar provisions were incorporated into the 2005 MSJC Code. For masonry construction in running bond, and for masonry in stack bond construction with bond beams space no further than 48 in. (1,219 mm) on center, the maximum width of the compression area used is limited to the least of:

  • the center-to-center spacing of the reinforcement;
  • six times the nominal thickness of the element; or
  • 72 in. (1,829 mm).

Boundary Elements

New to the 2005 MSJC Code are design provisions for boundary elements incorporated into the end zones of masonry shear walls. Because the MSJC does not include minimum prescriptive reinforcement detailing requirements for boundary elements, it requires that when used, their performance be verified by testing. In addition, when meeting the requirements for geometrical symmetry, axial load and flexural performance, the MSJC Code permits the maximum reinforcement limits as reviewed above to be waived. This exception may or may not require the incorporation of special boundary elements, depending on the design circumstances.

2008 MSJC STRENGTH DESIGN REQUIREMENTS

At the time of publication, the 2008 edition of the MSJC Code had been finalized, but the 2009 edition of the IBC had not. In anticipation of the 2009 IBC adopting the 2008 MSJC by reference, following is a brief overview of changes to the strength design provisions of the 2008 MSJC Code. In addition to some reorganization, substantive revisions to the strength design method include:

  • The shear strength check to help preclude brittle failure of a masonry element during an earthquake was revised to apply only to special reinforced masonry shear walls. Previously, this check applied to both reinforced and unreinforced masonry elements regardless of seismic considerations.
  • The requirement to provide at least 80 percent of the lateral stiffness through shear walls was revised to apply only to Seismic Design Categories C and higher.
  • Instead of requiring that the deflection of all reinforced masonry elements be calculated on cracked section properties, the 2008 MSJC Code simply requires that the design “consider the effects of cracking and reinforcement on member stiffness,” thereby permitting more flexibility in design application.
  • A moment magnifier approach was incorporated for unreinforced masonry design in determining the influence of P-delta effects. The magnified moment, Mc, is determined either through a second-order analysis, or by the following:

For members with h/r ≤ 45, it is permitted to take δ = 1.0. For members with 45 < h/r ≤ 60, it is permitted to take δ = 1.0 provided that the maximum factored axial stress on the element does not exceed 0.72f’m.

NOTATIONS

An     = net cross-sectional area of masonry, in.² (mm²)
As      = effective cross-sectional area of reinforcement, in.² (mm²)
Av     = cross-sectional area of shear reinforcement, in.² (mm²)
A1     = bearing area under bearing plate, in.² (mm²)
A2     = effective bearing area, in.² (mm²)
a       = depth of an equivalent compression zone at nominal strength, in. (mm)
b       = width of section, in. (mm)
bw    = for partially grouted walls, width of grouted cell plus each web thickness within the compression zone, in. (mm)
C      = resultant compressive force, lb (N)
c       = distance from the fiber of maximum compressive strain to the neutral axis, in. (mm)
Cd    = seismic deflection amplification factor
Cn    = nominal bearing strength, lb (N)
D     = dead load, lb (N)
d      = distance from the extreme compression fiber to centroid of tension reinforcement, in. (mm)
db     = nominal diameter of reinforcement, in. (mm)
dv     = actual depth of masonry in direction of shear considered, in. (mm)
d’      = distance from the extreme tension fiber to centroid of compression reinforcement, in. (mm)
Em    = modulus of elasticity of masonry, psi (MPa)
Es     = modulus of elasticity of reinforcement, psi (MPa)
eu     = eccentricity of Puf in. (mm)
Fu     = net flexural bending stress due to factored loads, psi (MPa)
Fvu   = shear stress on unreinforced masonry elements, psi (MPa)
f’m    = specified compressive strength of masonry, psi (MPa)
fr      = modulus of rupture, psi (MPa)
fy      = specified yield strength of reinforcement, psi (MPa)
h       = effective height of masonry element, in. (mm)
hsx    = effective height of masonry element below level x, in. (mm)
Icr     = moment of inertia of cracked cross-sectional area of a member, in.4 (mm4)
Ig      = moment of inertia of gross cross-sectional area of a member, in.4 (mm4)
In      = moment of inertia of net cross-sectional area of a member, in.4 (mm4)
L       = live load, lb (N)
M      = unfactored maximum calculated bending moment at the section under consideration, in.-lb (N-mm)
Mc     = factored moment magnified for the effects of member curvature, in.-lb (N-mm)
Mcr    = cracking bending moment, in.-lb (N-mm)
Mn     = nominal moment strength, in.-lb (N-mm)
Mser  = applied bending moment, in.-lb (N-mm)
Mu    = factored moment, in.-lb (N-mm)
Nv     = compressive force acting normal to the shear surface, lb (N)
P       = unfactored axial load, lb (N)
Pn     = nominal axial strength, lb (N)
Pu     = factored axial load, lb (N)
Puf    = factored load from tributary floor or roof areas, lb (N)
Puw   = factored weight of wall area tributary to wall section under consideration, lb (N)
QE     = the effect of horizontal seismic forces, lb (N)
Qn      = first moment about the neutral axis of a section of that portion of the net cross section lying between the neutral axis and extreme fiber, in.³ (mm³)
R        = seismic response modification factor
r         = radius of gyration, in. (mm)
Sn       = section modulus of cross-section, in.³ (mm³)
s         = spacing of shear reinforcement, in. (mm)
T        = tension in reinforcement, lb (N)
t         = specified thickness of masonry element, in. (mm)
tfs       = concrete masonry face shell thickness, in. (mm)
V        = unfactored shear force, lb (N)
Vm     = shear strength provided by masonry, lb (N)
Vn      = nominal shear strength, lb (N)
Vs       = shear strength provided by shear reinforcement, lb (N)
Vu        = factored shear, lb (N)
wu      = out-of-plane factored uniformly distributed load, lb/in. (N/mm)
X        = for partially grouted masonry, distance from extreme compression fiber to centroid of the compression resultant, in. (mm)
α        = tension reinforcement yield strain factor
δ         = moment magnification factor
δs        = deflection due to service loads, in. (mm)
δu        = deflection due to factored loads, in. (mm)
εmu      = maximum usable compressive strain of masonry
εs         = steel strain
εy         = yield strain of reinforcement
ρ          = reinforcement ratio
ρmax    = maximum reinforcement ratio
Φ         = strength reduction factor

REFERENCES

  1. Building Code Requirements for Masonry Structures, ACI 530-02/ASCE 5-02/TMS 402-02. Reported by the Masonry Standards Joint Committee, 2002.
  2. 2003 International Building Code. International Code Council, 2003.
  3. Building Code Requirements for Masonry Structures, ACI 530-05/ASCE 5-05/TMS 402-05. Reported by the Masonry Standards Joint Committee, 2005.
  4. 2006 International Building Code. International Code Council, 2006.
  5. Building Code Requirements for Masonry Structures, ACI 530-08/ASCE 5-08/TMS 402-08. Reported by the Masonry Standards Joint Committee, 2008.
  6. Empirical Design of Concrete Masonry Walls, TEK 1408B, Concrete Masonry & Hardscapes Association, 2008.
  7. ASD of Concrete Masonry (2012 IBC & 2011 MSJC), TEK 14-07C, Concrete Masonry & Hardscapes Association, 2004.
  8. Drysdale, R. G., Hamid, A. A. and Baker L. R., Masonry Structures, Behavior and Design, Second Edition. The Masonry Society, 1999.
  9. Standard Specification for Low-Alloy Steel Deformed and Plain Bars for Concrete Reinforcement, ASTM A 706/A 706M-01. ASTM International, 2001.
  10. Building Code Requirements for Structural Concrete, ACI 318-02. American Concrete Institute, 2002.
  11. Minimum Design Loads for Buildings and Other Structures, ASCE 7-02. Structural Engineering Institute of the American Society of Civil Engineers, 2002.
  12. Design of Anchor Bolts Embedded in Concrete Masonry, TEK 12-03C. Concrete Masonry & Hardscapes Association, 2013.
  13. Splices, Development and Standard Hooks for CM Based on the 2009 & 2012 IBC, TEK 12-06A, Concrete Masonry & Hardscapes Association, 2013.
  14. Weights and Section Properties of Concrete Masonry Assemblies, CMU-TEC-002-23, Concrete Masonry & Hardscapes Association, 2023.

Joint Reinforcement for Concrete Masonry

INTRODUCTION

Standard joint reinforcement for concrete masonry is a factory fabricated welded wire assembly consisting of two or more longitudinal wires connected with cross wires forming a truss or ladder configuration. It was initially conceived primarily to control wall cracking associated with thermal or moisture shrinkage or expansion and as an alternative to masonry headers when tying masonry wythes together. Note that horizontal steel requirements for crack control can be met using joint reinforcement or reinforcing bars. See Crack Control Strategies for Concrete Masonry Construction, CMU TEC-009-23 (ref. 6).

Joint reinforcement also increases a wall’s resistance to horizontal bending, but is not widely recognized by the model building codes for structural purposes. In some instances, it may be used in design for flexural resistance or to meet prescriptive seismic requirements.

This TEK discusses the code and specification requirements for joint reinforcement and presents a general discussion of the function of joint reinforcement in concrete masonry walls. Detailed information on additional uses for joint reinforcement can be found in other TEK as referenced throughout this publication.

MATERIALS

Reinforcement types used in masonry principally are reinforcing bars and cold-drawn wire products. Joint reinforcement is governed by Standard Specification for Masonry Joint Reinforcement, ASTM A 951 (ref. 1), or Standard Specification for Stainless Steel Wire, ASTM A 580/580M Type 304 or Type 316 (ref. 2), if the joint reinforcement is stainless steel according to the Specification for Masonry Structures (ref. 3). Cold-drawn wire for joint reinforcement varies from W1.1 to W4.9 (11 gage to 1/4 in. diameter; MW7 to MW32), the most popular size being W1.7 (9 gage, MW11). Wire for masonry is plain, except side wires for joint reinforcement are deformed by means of knurling wheels.

Because Building Code Requirements for Masonry Structures (ref. 4) limits the size of joint reinforcement to one half the joint thickness, the practical limit for wire diameter is W2.8, (3/16 in., MW17) for a 3/8 in. (9.5 mm) bed joint. Joint reinforcement of this thickness may be difficult to install however, if a uniform mortar joint thickness of 3/8 in. (9.5 mm) is to be maintained.

Types of Joint Reinforcement

Reflecting its multiple purposes in masonry walls, joint reinforcement comes in several configurations. One longitudinal wire is generally required for each bed joint (i.e., two wires for a typical single wythe wall), but code or specification requirements may dictate otherwise. Typical joint reinforcement spacing is 16 in. (406 mm) on center. Adjustable ties, tabs, third wires and seismic clips are also available in combination with joint reinforcement for multi-wythe and veneer walls.

  • Ladder-type joint reinforcement (Figure 1) consists of longitudinal wires flush welded with perpendicular cross wires, creating the appearance of a ladder. It is less rigid than truss type joint reinforcement and is recommended for multi-wythe walls with cavity spaces or unfilled collar joints. This permits the two wythes to move independently, yet still transfers outof-plane loads from the exterior masonry to the interior masonry wall. Cross wires 16 in. (406 mm) on center should be used for reinforced concrete masonry construction, to keep cross wires out of the core spaces, thus preventing them from interfering with the placement of vertical reinforcement and grout.
  • Truss-type joint reinforcement (Figure 2) consists of longitudinal wires connected with diagonal cross wires. This shape is stiffer in the plane of the wall than ladder-type joint reinforcement and if used to connect multiple wythes restricts differential movement between the wythes. For this reason, it should be used only when differential movement is not a concern, as in single wythe concrete masonry walls. Because the diagonal cross wires may interfere with the placement of vertical reinforcing steel and grout, truss type joint reinforcement should not be used in reinforced or grouted walls.
  • Tabs, ties, anchors, third wires and seismic clips of assorted configurations are often used with the joint reinforcement to produce a system that works to: control cracking; bond masonry wythes together; anchor masonry; and, in some cases, resist structural loads. Tie and anchor spacing and other requirements are included in Anchors and Ties for Masonry, TEK 12-01B (ref.5).

Recommendations for the use of some of the different types of joint reinforcement are listed in Table 1.

CORROSION PROTECTION

Grout, mortar and masonry units usually provide adequate protection for embedded reinforcement, provided that minimum cover and clearance requirements are met.

Coating Requirements

The carbon steel in joint reinforcement can be protected from corrosion by coating with zinc (galvanizing). The zinc protects steel in two ways. First, it provides a barrier between the steel and oxygen and water. Second, during the corrosion process, the zinc provides a sacrificial coating. The protective value of the zinc coating increases with increased coating thickness; therefore the required amount of galvanizing increases with the severity of exposure, as listed below (refs. 3, 4):

  • Interior walls exposed to a mean relative humidity less than or equal to 75%:
    Mill galvanized, ASTM A 641 (0.1oz/ft2)
    (0.031 kg/m2)
    Hot-dip galvanized, ASTM A 153 (1.5 oz/ft2)
    (458 g/m2)
    Stainless steel AISI Type 304 or Type 316
    conforming to ASTM A 580
  • Exterior walls or interior walls exposed to a
    mean relative humidity > 75%:
    Hot-dip galvanized, ASTM A 153 (1.5 oz/ft2 (0.46 kg/m2)
    Epoxy coated, ASTM A 884 Class A Type 1, >
    7 mils (175 mm)
    Stainless steel AISI Type 304 or Type 316
    conforming to ASTM A 580

Cover Requirements

Specification for Masonry Structures also lists minimum cover requirements for joint reinforcement as a further means of corrosion protection. It must be placed so that longitudinal wires are embedded in mortar with a minimum cover of:

  • 1/2 in. (13 mm) when not exposed to weather or earth,
  • 5/8 in. (16 mm) when exposed to weather or earth.

PRESCRIPTIVE CODE REQUIREMENTS

Building Code Requirements for Masonry Structures includes prescriptive requirements for joint reinforcement. There are multiple uses for joint reinforcement in masonry structures. Joint reinforcement can be used to provide crack control, horizontal reinforcement, and bond for multiple wythes, corners and intersections. The following list highlights only those requirements specific to joint reinforcement. Crack control topics are covered in CMU-TEC-009-23 (ref. 6). For information on anchors and ties, see Anchors and Ties for Masonry, TEK 12-01B (ref. 5). There is also a useful discussion on joint reinforcement as structural reinforcing in Steel Reinforcement for Concrete Masonry, TEK 12-04D (ref. 7).

General Requirements for Joint Reinforcement

  • For masonry in other than running bond: Horizontal reinforcement shall be 0.00028 times the gross vertical cross-sectional area of the wall. This requirement can be met with joint reinforcement placed in the horizontal bed joints. For 8in. (203-mm) masonry walls, this amounts to W1.7 (9 gage, MW11) joint reinforcement every other course. There are additional criteria for stack bond masonry in Seismic Design Categories D, E and F.
  • Seismic Requirements: In Seismic Design Category C and higher (for concrete masonry other than veneer), horizontal joint reinforcement spaced not more than 16 in. (406 mm) on center vertically with at least two wires of W1.7 (MW11) is required. Horizontal reinforcement also must be provided at the bottom and top of all wall openings and must extend at least 24 in. (610 mm) past the opening. Additional details on seismic requirements, including shear walls, are covered in Seismic Design and Detailing Requirements for Masonry Structures, CMHA TEK 14-18B (ref. 8).

Allowable Stress Design Requirements

  • In addition to the requirements above, concrete masonry walls designed by the allowable stress method and bonded by wall ties must have a maximum tie spacing of 36 in. (914 mm) horizontally and 24 in. (610 mm) vertically. Joint reinforcement cross wires can be used in place of wall ties to meet this requirement.
  • When the walls are designed for noncomposite action, truss-type joint reinforcing is not to be used for tying the wythes.
  • Combination joint reinforcement with tabs or adjustable ties are popular options for bonding multiwythe walls and are governed by additional code requirements.

Empirical Design Requirements

  • When two wythes of masonry are bonded with joint reinforcement, at least one cross wire must serve as a tie for each 22/3 ft2 (0.25 m2) of wall area. The vertical spacing of the joint reinforcement can not exceed 24 in. (610 mm), and the cross wires must be W1.7 (9 gage, MW11) minimum, without drips, and embedded in mortar.
  • Intersecting walls, when depending on each other for lateral support, can be anchored by several prescriptive methods including the use of joint reinforcement spaced no more than 8 in. (203 mm) on center vertically. The longitudinal wires must extend at least 30 in. (762 mm) in each direction at the intersection and be at least W1.7 (9 gage, MW11).
  • Interior nonloadbearing wall intersections may be anchored by several prescriptive methods, including joint reinforcement at a maximum spacing of 16 in. (406 mm) o.c. vertically.

Requirements for Use in Veneer

  • Prescriptive requirements for joint reinforcement in masonry veneer are included in Building Code Requirements for Masonry Structures, Chapter 6. These provisions are limited to areas where the basic wind speed does not exceed 110 mph (177 km/hr) as listed in ASCE 7-02 (ref. 9). Additional limitations are covered in the Code. The information below is for joint reinforcement or the joint reinforcement portion of a tie/anchor system. For information on anchor and tie requirements see Concrete Masonry Veneers, TEK 03-06C (ref. 10).
  • Ladder-type or tab-type joint reinforcement is permitted in veneer construction with the cross wires used to anchor the masonry veneer. Minimum longitudinal and cross wire size is W1.7 (9 gage, MW11), and maximum spacing is 16 in. (406 mm) on center vertically.
  • Adjustable anchors combined with joint reinforcement may be used as anchorage with the longitudinal wire of the joint reinforcement being W1.7 (9 gage, MW11) minimum.
  • Joint reinforcement may also be used to anchor masonry veneer to masonry provided the maximum distance between the inside face of the veneer and the outside face of the concrete masonry backup wythe is 4 1/2 in. (114 mm).
  • In Seismic Design Categories E and F, the 2005 edition of Building Code Requirements for Masonry Structures requires continuous single wire joint reinforcement, W1.7 (9 gage, MW11) minimum, in the veneer wythe at a maximum spacing of 18 in. (457 mm) on center vertically. Clips or hooks must attach the wire to the joint reinforcement. The International Building Code 2003 (ref. 11) also mandates this requirement for Seismic Design Category D.
  • Anchor spacings, and, as a result, possibly joint reinforcement spacing, are reduced for Seismic Design Categories D, E and F and in high wind areas.

Requirements for Use in Glass Unit Masonry

  • Horizontal joint reinforcement is to be spaced no more than 16 in. (406 mm) on center, located in the mortar bed joint, and must not span across movement joints.
  • Minimum splice length is 6 in. (152 mm).
  • Joint reinforcement must be placed immediately above and below openings in the panel.
  • Joint reinforcement must have at least 2 parallel, longitudinal wires of size W1.7 (9 gage, MW11) and have welded cross wires of W1.7 (9 gage, MW11) minimum.

INSTALLATION

Joint reinforcement installation is a routine task for masons. The joint reinforcement is placed on the face shells and mortar is placed over it. Cover requirements must be maintained. Installing the correct type of joint reinforcement with the specified corrosion resistant coating is important, as is making sure it is installed at the proper spacings and locations. Quality assurance provisions related to joint reinforcement generally include:

Submittals

Material Certificate indicating compliance should include:

  • material meets specified ASTM standard,
  • corrosion protection specified has been supplied,
  • configuration specified has been supplied, and
  • other criteria as required or specified.

Inspection

Oil, dirt and other materials detrimental to bond should be
removed. Light rust and mill scale are permissible.

  • Cover requirements are met.
  • Splices are a minimum of 6 in. (152 mm) (see Figure 3) to properly transfer tensile stresses. Tying is not necessary. Construction documents may specify longer splices, especially if the joint reinforcement is being used as part of the structural horizontal reinforcing steel.
  • Verify that joint reinforcement utilized for crack control does not continue through movement joints.
  • If ties or anchors are part of the joint reinforcement, check that embedment in the adjoining wythe, alignment and spacing are within specified values.

REFERENCES

  1. Standard Specification for Masonry Joint Reinforcement, ASTM A 951-02. ASTM International, 2002.
  2. Standard Specification for Stainless Steel Wire, ASTM A 580/580M-98(2004). ASTM International, 2004.
  3. Specification for Masonry Structures, ACI 530.1-05/ASCE 6 05/TMS 602-05. Reported by the Masonry Standards Joint Committee, 2005.
  4. Building Code Requirements for Masonry Structures, ACI 530 05/ASCE 5-05/TMS 402-05. Reported by the Masonry Standards Joint Committee, 2005.
  5. Anchors and Ties for Masonry, TEK 12-01B, Concrete Masonry & Hardscapes Association 2011.
  6. Crack Control Strategies for Concrete Masonry Construction, CMU-TEC-009-23, 2023.
  7. Steel Reinforcement for Concrete Masonry, TEK 12-04D, Concrete Masonry & Hardscapes Association, 2023.
  8. Seismic Design and Detailing Requirements for Masonry Structures, TEK 14-18B, Concrete Masonry & Hardscapes Association, 2003.
  9. Minimum Design Loads for Buildings and Other Structures, ASCE 7-02, American Society of Civil Engineers, 2002.
  10. Concrete Masonry Veneers, TEK 03-06C, Concrete Masonry & Hardscapes Association, 2012.
  11. International Building Code 2003. International Code Council, 2003.