Download as pdf
Download as pdf
You are on page 1of 19
Presented at the Third National Concrete and Masonry Engineering Conference, San Fancisco, CA, June, 1995. EVALUATION AND RETROFIT OF UNREINFORCED MASONRY BUILDINGS Gregory R. Kingsley, PE. Senior Engineer, Atkinson-Noland & Associates, Inc. 2619 Spruce St., Boulder Colorado, 80302 U.S.A. NCMEC Session: Retrofit of Masonry Buildings INTRODUCTION ‘The recent Loma Prieta and Northridge earthquakes have provided a proving ground for extensive efforts in the last decade to decrease the seismic hazard presented by unreinforced masonry (URM) building systems in seismic zones, while underscoring the urgency of attending to the thousands of URM buildings which have not yet been retrofitted [27,29]. The task provides a considerable engineering challenge, not only due to the complex response of URM. systems to seismic loads, but also due to the many additional architectural, historical, social and economical considerations involved. This paper addresses some engineering aspects of the problem, with emphasis on the problem of evaluating URM system behavior. First, the basic mechanisms of seismic load resistance in URM building systems are outlined, concentrating primarily on simplified representations of the in-plane and out-of-plane response of URM walls and piers. Understanding the behavior and potential failure modes of URM systems will provide a basis for understanding the rationale of codes and current retrofit practice. ‘The first step of the retrofit process is then addressed: evaluation and assessment of existing URM structures, Several test techniques are presented which allow the engineer to quantify his analysis of existing structural conditions. Finally, some retrofit approaches are presented, emphasizing different solutions or approaches from around the world, some of which may be new to U.S. engineers. A detailed review of current U.S. practice, with specific design case studies, will be presented by Uzarski [32] as part of the same conference session. Building Codes for Seismic Retrofit of Unreinforced Masonry The first coordinated analytical and experimental effort to establish a rational practice for the retrofit of masonry structures was conducted by ABK - A Joint Venture, and resulted in the publication of a now well-known methodology for mitigation of seismic hazards in existing unreinforced masonry buildings [2]. With some modification to account for the use of a code ‘working stress design approach, the methodology was adopted by the City of Los Angeles Building Code as Division 88 [10]. In 1991, the Uniform Code for Building Conservation adopted provisions for the assessment and retrofit of URM structures, again based in large part on the ABK approach. This code provides the basis for current practice in most places outside of the Los Angeles area, and is the standard of reference for building retrofit codes cited in this paper. Currently, a large-scale effort has been initiated by the Federal Emergency Management Agency (FEMA), as part of the National Earthquake Hazards Reduction Program (NEHRP), to develop Guidelines for the Seismic Rehabilitation of Buildings. This project, known as ATC-33, is being conducted by the Applied Technology Council (ATC) under contract to the Building Seismic Safety Council. Completion of the final draft is expected in 1996 [3]. EXPECTED BEHAVIOR OF URM BUILDINGS Load Path ‘The interaction of components in an idealized unreinforced masonry building is illustrated in Figure 1 [23]. The energy path begins with the ground acceleration which passes first into the foundation, and then into the in-plane structural walls. The interaction of the foundation with the surrounding soils may well have a significant influence on the building response, but this subject is not treated here. The response of the in-plane structural walls to the ground acceleration depends on their height, stiffness, and contributory mass. ‘These walls, in turn, provide the input acceleration to the diaphragms, which, finally, excite the out-of-plane walls in a manner which depends on their own stiffness, damping, and dimensions. Since the diaphragms are likely to be flexible, the input acceleration to the out-of-plane walls is likely to be modified from the response of the end walls. Typically, the out-of-plane excitement and failure of URM walls presents the ‘most significant hazard in the system, and retrofit measures are often concentrated on decreasing diaphragm displacements or strengthening the attachment of the diaphragms to the walls. The expected behavior of each of the three sub-systems are addressed in tum below. In-plane Response of Walls and Piers ‘The response of URM walls and piers to in-plane forces may be idealized by considering three simplified modes of behavior independently (see Figure 2): 1. Flexural behavior characterized by rocking, 2. Shear failure due to diagonal tension, 3. Shear failure due to sliding on a single bed joint, or along a stair step sliding plane, ‘The first may be considered as the most desirable, ductile failure mode, although large in-plane rocking displacements may have a deleterious effect on the out-of-plane stifiness and strength of the pier. The second mode is characterized by brittle tensile cracking, probably through both masonry units and mortar, and a sudden loss of lateral load capacity. It is the least desirable of the three modes. In general, the shear/sliding mode is also undesirable, because it may ultimately lead to instability or significant damage; however, it is not necessarily associated with a sudden loss of vertical or lateral load capacity. For the case of squat piers under moderate vertical loads, the shear sliding response may actually be quite ductile, although significant cracking and residual displacements may result. Predicting the in-plane response of unreinforced walls and piers is not a simple matter, depending as it does on the numerous materials in a masonry system, their interaction, and the magnitude of vertical loads. The subject has been the focus of increasing attention of analysts who have adopted many different approaches, with various degrees of success, ranging from nonlinear finite element analysis (FEM) with discrete crack models [18], FEM models with homogenized brick and mortar properties [15], simple component models [12], or hybrids of component and FEM ‘models [11]. For purposes of design, however, some rather broad simplifying assumptions can be made to allow a simplified analysis. In the following, equations defining the simplified response of a wall or pier of height H and width D are presented. The pier is assumed to be under a vertical dead load of Pp . Recommended values given in the UCBC are also presented. Pier Stiffness. The combined flexural and shear stiffness, K, of an uncracked URM pier of height H may be approximated based on the simple linear elastic theory: oeatae oe K eects] () where ¥y is a factor which takes a value of 2 when the pier is constrained against rotation at the top and the bottom, or 1 when the pier acts as a cantilever free to rotate atthe top. The elastic constants Eq and Gq can be based on in-situ tests (see below), and the area A,, and moment of inertia J, are based on the gross cross section. The UCBC allows the first, flexural term to be neglected in the calculation of pier stiffness. ‘The stiffness following cracking will depend on which mode of response occurs in the pier. Flexural Response (Rocking). Assuming that the interface between mortar and masonry unit may have an initial tensile capacity, the shear in a rocking pier at first cracking is given by (20) WD(Pp , foDt Vee OH ( 6° 6 ® where the dead load term neglects the self-weight of the pier itself for simplicity. (See the appendix of notation for definition of terms). Ultimate strength of a rocking pier will occur at significantly larger deflections, and, depending on the mortar joint tensile strength and axial restraint of the pier, at significantly larger loads. Assuming that the stress distribution in the compression zone at ultimate may be described by a standard stress block with an ultimate stress Of fm the shear atthe ultimate rocking shear strength will be: z) : ‘The shear at the ultimate flexural failure will be greater than the shear at first cracking when the joint tensile strength satisfies the criteria: fa 32p. fe eo ) Po 085fz . Since the tensile strength at the brick/mortar interface is generally quite low, and under seismic actions, the join is ikely to have cracked during low level cycles, Section A108.4 of the UCBC specifies that unreinforced masonry shall be assumed to have no tensile capacity. ‘The pier rocking shear capacity given in Section A112.2.2 of the UCBC assumes a simplified form of equation (3): D sy © For walls without openings, Section A112.2.3 allows a portion panel self weight, Py ,to be included in the axial load: V, = (05P, +1 oasp,}? © Shear Response. For a masonry pier with a characteristic diagonal tension strength of fu, the maximum panel shear may be approximated by [20]: LuDt |, Po > tu? Although diagonal tension failure modes have been observed in laboratory tests and in actual earthquakes, the term fis difficult to characterize, especially for existing structures, and it is therefore difficult to confidently predict the diagonal tension shear strength using equation (7). The UCBC states that unreinforced masonry wall piers need not be analyzed for tension stress, and this failure mode is neglected by the code. Sliding Response. When the response of a pier is characterized by sliding along the joimts, the pier shear capacity may be approximated by: V,=1,Dt+ Py ® Considering the repeated load reversals expected under seismic loading, it is advisable to take the initial cohesion to be 1,=0 [23], thus: V,=nP, o The sliding coefficient of friction value of 0.7 has been suggested as a reasonable lower bound for a wide range of masonry types used in the U.S.[5]. Va, o=4. ossis 7) =>° ¢ ‘The UCBC specifies the shear capacity of a pier to be (10) where the allowable shear stress, v, , is dependent on the axial load and the results of in-place shear tests as described below. Identifying failure modes. The failure mode which will govern in a given pier can be determined by calculating the shear strength for each mode and comparing them. Altematively, some simple rules of thumb may be useful for quick analysis. For example, the sliding shear strength will be greater than the diagonal tension shear strength when the joint coefficient of friction satisfies the relationship [20]: f, ’p fa fy, 20 epee yee ay Similarly, a flexural rocking mode will occur before a sliding shear mode if the pier aspect ratio satisfies [23]: He ee D its #] a2) Defining Pier Height. Experimental results [19] have shown that, even when a pier behaves in a rocking mode, failure planes may not necessarily occur at clear horizontal planes at the top and bottom of piers. Further analytical studies have verified that the definition of pier “height” is neither straightforward nor unique [19], particularly in a wall containing doors and windows of different sizes. Dolce [12] recommends the use of an “equivalent height” in stiffness and strength calculations. Referring to Figure 3, the effective height is calculated as el nensi{(eapsn, (3) where hi, is the interstory height, d is the pier width, and his defined as the distance between the midpoints of the segments that join the top comers and the lower comers of the adjacent ‘openings. A maximum inclination of + 30° on the horizontal is allowed for these segments. Diaphragms Very little experimental work has been conducted to understand the behavior of the types of flexible diaphragms typically found in URM buildings. The primary resource for this information is a study conducted by ABK [1] in support of their methodology. ‘They found that the diaphragm response to input motions may be strongly nonlinear for large ground accelerations, and that response is most strongly influenced by (1) the initial diaphragm stiffness, (2) the yield capacity of the diaphragm, (3) the available inertial mass coupled with the diaphragm, (4) the materials and construction, and (5) the span to depth ratio. They pointed out that realistic analysis of diaphragms should include nonlinear effects. Rather than complicate the design methodology, considering the degree of confidence likely in ‘measuring the relevant parameters in an existing building, the authors made use of the strong influence of crosswalls (generally sheathed wood walls, parallel to shear walls and attached to the diaphragms) on the damping and displacements of the diaphragm, and defined a set of relatively simple criteria for determining diaphragm response based on the materials and construction, dimensions, and number and spacing of crosswalls, see [32]. Out-of-plane Response of Walls and Piers ‘The out-of-plane response of walls and piers may be somewhat more difficult to evaluate than the in-plane response. The load input to the wall panel depends very much on the response of the diaphragms and the in-plane walls, and the response mechanism is dependent upon the boundary conditions around the panel and the eccentricities of the applied vertical loads. A rigorous analysis would require knowledge of: accelerations input at the top and bottom of a wall by the diaphragms ‘end restraint provided by slabs, beams, or spandrels above and below the walls wall compressive strength joint tensile strength initial wall stiffness axial loads, including eccentricity of application ‘As in the analysis in-plane response, however, some simplifying assumptions can be made, if only to gain some insight into the probable response [25]. Consider, for example, an unreinforced panel ideally fixed at top and bottom, under vertical load P, and a uniform pressure p (inertial loading) acting perpendicular to the face. The lateral load resistance derives from the development of an arching mechanism between the floor slabs or spandrel beams. It can be shown that first cracking will occur when the uniform pressure attains a value of Pe = +E) a4) If the joint tensile strength is taken as zero, this simplifies to Fo ft. narts(t) 0 Atultimate strength, the same panel will sustain a lateral uniform pressure of Py Fea) (16) where a is the depth of the compression zone. Since the compression zone depth will be nearly zero, the ultimate strength will have an upper bound of (4) 17 P= Or a7 ‘Comparison of equations (15) and (17) reveals that, for this particular case, the ultimate strength may be as much as 3 times the strength at first cracking. ‘This helps to account for the fact that many URM panels survive wind and seismic pressures when elastic analysis indicates that they should have failed. Because of the simplifying assumptions required and the complexity of the dynamic loading situation, direct use of equations (14)-(17) is probably not justified. However, itis instructive to note that, for both the elastic and ultimate load cases, the load capacity of the panel is directly proportional to the axial load, and inversely proportional to the wall slendemess (hit) ratio. This corresponds to the well known fact that out-of-plane failures of URM walls often occur at the upper stories of structures, where the overburden load is a minimum, and the amplification of ground motions by the end walls and diaphragms is likely to be a maximum. It is also clear that, since the axial load level is not easily controlled as a design parameter, decreasing the wall height- to-thickness ratio is the most direct means of improving the seismic resistance of an out-of-plane wall. This may be accomplished either by bracing the wall to decrease the effective wall height, or strengthening the bond between multiple wythes to ensure that the largest possible wall thickness is engaged. These measures are discussed further below. To avoid undue complexity in the analysis, the UCBC takes an empirical approach, providing simple limits to wall h/t ratios as specified in UCBC Table A-1-B, reproduced below. UCBC TABLE A-1-B - ALLOWABLE VALUE OF HEIGHT-TO-THICKNESS RATIO OF UNREINFORCED MASONRY WALLS WALLTYPES | SEISMICZONE 7B | SESWIGZONES | SESMICZONE4 | SEISWICZONES BUILDINGS BuLDINGs | BULDWes wih | ~ ALLOTHER cRosswauis’ | BUILDINGS Walls of one-story 16 r B buildings cu e Fist-tory wall of 20 18 16 15 listory bulking Wallin top story of “4 14 14s 9 rutistory building Allother walls 20 16 16 13 “Applies to the special procedures of Section A111 only. See Section AIL1,7 for other restrictions ‘This value of height-1o-thickness ratio may be used only where mortar shear tess establish a tested mortar shear strength, ¥, ‘of not less than 100 psi (690 kPa). This value may also be used where the tated mortar shear strength i not less than 60 psi (414 KPa) and a visual examination ofthe collar joint indicates not less than 50 percent mortar coverage, ‘Where visual examination of the collar joint indicates no less than 50 percent mortar coverage, and the tested mortar shear strength, v, is greater than 30 psi (207 kPa) but less than 60 psi (#14 kPa), the allowable heightto-hickness ratio may be f\\_ I Teo DIAL GAGE —J TA IE — 2 ae JOINT “REMOVED Figure 4. The in-situ sher test measures shear resistance of a mortar bed joint [7]. Figure 5. A modified procedure for measuring bed joint shear strength uses flatjacks, above and below the test unit to control vertical stresses on the test joints [7]. Removable Flatjack Dial Gage Gage Point: 6 Goge xisting : Stress Hydraulic Hand Pump Figure 6. The in-situ stress test is used for determination of the magnitude of compressive stresss acting normal to the bed joints [7]. ‘Surface—Mount Dial Gage Figure 7. The in-situ deformability test provides a measure of the masonry strength and deformability using two flatjacks to subject the intervening masony to compressive stress 7].

You might also like