You are on page 1of 41

10

Chapter 2
Simulating Fracture in SCBF Bracing Components - Background

Several experimental and analytical investigations (Foutch et al 1987, Uriz 2005) and post-

earthquake reconnaissance reports (Tremblay et al 1995, Kelly et al 2000) have suggested that

Special Concentrically Braced Frames (SCBFs) may show unsatisfactory performance during

earthquake-type loading. Figure 2.1 illustrates a large-scale SCBF experiment by Uriz (2005) and

component tests conducted as part of the current study which both indicated local buckling-

induced fracture of the bracing members during inelastic buckling and tension yielding. In

addition to the brace fractures shown in the figure, fracture of brace-gusset plate connections

from large force and deformation demands on the connections is also a concern. Since SCBFs

rely on the buckling and yielding action of the brace components to dissipate seismic energy,

sufficient brace ductility must be provided to preclude premature fracture. Assessing the fracture

ductility of large-scale bracing components or systems is expensive, especially considering the

heavy dependence on experiment-based techniques. Moreover, testing of large-scale components

may be infeasible due to the size and strength of structural members. Considering these issues,

reliable fracture simulation and prediction techniques for SCBF systems are highly attractive and

allow for improved practicality in modeling the effects of diverse material properties and loading

conditions at the component, as well as system, level.


11

In this context, this chapter presents a broad overview of various experimental and state-of-the-art

simulation-based investigations that aim to characterize the performance (specifically the fracture

performance) of SCBF systems. Seismic performance assessment of structural systems typically

involves two distinct methodological components – (1) Evaluating deformation or strain demands

at the story, component or continuum level (2) Determining the corresponding capacities through

a combination of experimental and analytical methods. A comparision between the two results in

characterization of structural performance. Although “demands” and “capacities” are abstract

concepts that interactively affect each other in reality, they form a convenient framwork for

characterizing structural performance within the limitations of current simulation techniques.

Thus, the background and literature review presented in this chapter is broadly divided into two

main parts – one focusing on demand characterization, and the other focusing on capacities.

Throughout the chapter the literature is reviewed and synthesized within the context of a

Performance Based Earthquake Engineering (PBEE) methodology.

The first section of the chapter is devoted to a literature review on the performance assessment of

SCBF systems, either through experimental or analytical investigations. These provide

background for subsequent chapters. Next, various modeling techniques are examined to describe

the state-of-the art in structural analysis and fatigue-fracture prediction. With respect to the latter,

these fall into various broad categories including (1) empirical equations to predict fatigue life as

a function of the geometrical parameters of bracing elements (2) fiber based approaches that

employ a critical strain as an indicator of fracture and (3) and micromechanics-based models that

operate at the continuum level and simulate the processes of void growth and coalescense that are

responsible for ductile fracture initiation.

While the major objective of this study is to examine these continuum-based fracture models for

large scale specimens, the chapter also presents work that has advanced the state-of-the art in
12

analyis techniques which have been used on braced-frame components. These are broadly

classified as (1) phenomenological or rule-based models (2) lumped plasticity analysis techniques

(3) fiber-based elements and (4) continuum-based or finite element analyses.

The chapter concludes by summarizing the current state of research, and identifying areas where

advanced simulation and fracture models may benefit structural and earthquake research and

practice.

2.1 PERFORMANCE EVALUATION OF SCBF SYSTEMS

To provide context for the subsequent literature review, this section discusses braced-frame

systems by first describing the desired response and possible fracture events during earthquake

loading, and then reviewing modern design provisions for SCBFs. Afterwards, the literature

review is presented by first focusing on braced-frame demand evaluation and then on the ductility

capacity of these systems.

2.1.1 SCBF SYSTEM BEHAVIOR

A schematic illustration of a one story chevron (inverted-V) braced-frame is shown in Figure 2.2a

where the hollow circles indicate locations of assumed pinned connections. Shear-tabs are used in

the beam-column connections and gusset plates (flexible out of plane) connect bracing members

to beams and columns. While these are more properly classified as partially fixed connections,

they are typically considered to be pinned in design and analysis procedures, although recent

research (e.g. Uriz, 2005) has illustrated that these may be substantially more rigid than often

presumed. As illustrated in Figure 2.2b, frame deformation is accommodated by the rotation of

these connections and axial deformation of the bracing members. Thus, unlike moment frames

where the beam-column connection is fixed and inelastic behavior is concentrated in beam plastic

hinges, braced frames rely on cyclic inelastic buckling and tension yielding of the bracing
13

members to accommodate the inelastic deformations and dissipate seismic energy. The figure

shows a schematic of a chevron braced frame deformed laterally, such that one brace is buckled

in compression and the other brace is in tension. Figure 2.2c shows the frame under larger

deformations, such that the compression brace has now formed a plastic hinge in the middle of

the brace. After several loading cycles – the number depending on cycle amplitude as well as

brace geometry and material type – a local buckle forms at the middle plastic hinge on a

compressive excursion. This localizes the plastic strain accumulation to the region of local

buckling. Fracture initiation closely follows local buckling, with complete cross-section rupture

and strength loss occurring soon thereafter.

Besides local buckling-induced fracture at the middle plastic hinge, other possible fracture events

could occur at the slotted-end brace-gusset connections, beam-column shear tab, or the gusset to

beam-column connections. These are illustrated in Figure 2.3a-b and are described below –

• Slotted-end brace-gusset (net section) connection fracture is shown schematically in

Figure 2.3a where the slot extending beyond the gusset plate creates a reduced section

susceptible to fracture during severe tensile loading cycles. As the brace is loaded in

tension, inelastic strains accumulate across a short gage length at the reduced area of the

connection, initiating fracture at the net section. Fabrication flaws could promote this

type of failure by introducing surface roughness or imperfections at the slotted section.

Furthermore, the net section is adjacent to the weld between the gusset-plate and brace

such that Heat Affected Zone (HAZ) defects could decrease the ductility of the brace

base metal as a result of material phase transitions or considerable grain growth.

• Beam-column shear tab fracture is illustrated in Figure 2.3b and was observed during a

large-scale test by Uriz (2005). It was assumed that fracture initiated at fabrication cracks

below the fillet weld of the shear tab to column connection. At large story drifts,
14

considerable rotation demands on the beam-column connection can pry the shear tabs

from the column flange and create localized stress intensities on the weld, and

surrounding base, material. These increased demands, combined with the variability of

base metal properties near welds could cause column fracture at large drift levels.

• Gusset-plate to beam or column connection fracture shown in Figure 2.3b is also

promoted by large rotation demands at the beam-column connection and the prying

action of the gusset plate on the flanges of the beam or column. Similar to the previous

two fracture events, fracture would also initiate close to the weld in the HAZ-affected

base metal.

In addition to these fracture limit states, inverted-V chevron configurations are susceptible to

concentrated inelastic behavior, and possible plastic hinge formation, at the upper-beam to brace

connection from a large force imbalance between the compression brace (small force) and the

tension brace (large force). The formation of a hinge at the mid-point of the beam could lead to a

story mechanism as illustrated in Figure 2.4. However, braced-frame system geometries with

alternating inverted-V and V bracing members (by story) or a zipper-frame configuration with

vertical axial columns at the beam mid-points (Bruneau et al, 2005) can prevent this type of

behavior by carrying the unbalance force.

Considering these possible failure mechanisms, current American Institute of Steel Construction

(AISC, 2005) Seismic Provisions aim to reduce inelastic effects in beams and columns and ensure

high ductility of the bracing members. While significant brace buckling and yielding is expected

in SCBFs during moderate to large earthquake events, detailing requirements guard against brace

and connection fracture. The design provisions most applicable to SCBF behavior are discussed

in the following section.


15

2.1.2 SEISMIC DESIGN PROVISIONS FOR SCBF SYSTEMS

The aim of modern steel seismic codes, such as the AISC Seismic Provisions (2005), is to ensure

acceptable system behavior through ductile performance of members and connections. Thus, in

the context of SCBFs, ductile system behavior is achieved through proper connection and brace

detailing to account for large rotations and repeated inelastic buckling and tension yield

excursions. The latter (ductile brace behavior) is attained through limits on geometric features

that control buckling mechanisms. While larger brace global slenderness has been shown to

increase brace ductility (Liu and Goel, 1988 and Tremblay, 2002), current seismic provisions

ensure ductile brace performance only through limits on the cross-section width-thickness, or

compactness, ratio (i.e., b/t for HSS). Referring to Figure 2.5, a more compact cross-section will

have a smaller width-thickness ratio and is less susceptible to local buckling during cyclic or

compressive loading. For HSS cross-sections, the Seismic Provisions (AISC, 2005) limit b/t

ratios as follows –

b / t < 0.64 E / Fy = 16 (for Fy = 46ksi ) (2.1.1)

Where b/t is used generically to represent the brace width-thickness ratio – as tabulated in AISC

design charts, such that b/t = B/t – 3 (i.e. (B-3t)/t), where B and t are the overall length of the

buckling face and the design thickness, respectively, as shown in Figure 2.5. Similar to the b/t

limit placed on square and rectangular bracing members, the current AISC Seismic Provisions

(2005) restrict D/t for Pipe cross-sections, such that –

D / t < 0.044 E / Fy = 36.5 (for Fy = 35ksi ) (2.1.2)

Where D/t represents the brace width-thickness ratio shown in Figure 2.5 – as tabulated in AISC

design charts, where D is the nominal outside diameter of the pipe and t is the design thickness.

Current AISC Seismic Provisions (2005) ensure ductile behavior by prescribing limits on Wide-

Flange width-thickness ratios –

b f / 2t f < 0.30 E / Fy = 7.2 (for Fy = 50ksi ) (2.1.3)


16

Where bf and tf are the flange width and thickness dimensions, respectively. Note that Equations

2.1.1-2.1.3 are independent of other parameters, such as the brace slenderness, which is also

presumed to have an impact on brace ductility (Tang and Goel, 1989 and Tremblay, 2002).

Figure 2.5 suggests that a more slender brace will also provide an increase in ductility through the

decreased curvature, and therefore strain, demand at the middle plastic hinge for a longer brace as

compared to a shorter brace for the same axial deformation. However, as the brace becomes

increasingly slender the force imbalance between the tensile yield and compressive buckling

loads will generally increase, negatively affecting the energy dissipation capabilities while

increasing the system overstrength factor and the force demand on the beam in a chevron-type

braced-frame configuration. For these reasons, the current AISC Seismic Provisions (2005) limit

the slenderness of bracing members in SCBF design by –

KL / r < 4 E Fy (2.1.4)

Thus, KL/r limits are prescribed as 100 (Fy = 46 ksi), 115 (Fy = 35 ksi) and 96 (Fy = 450 ksi) for

HSS A500 Grade B, Pipe A53 Grade B and A992 Wide-Flange bracing members, respectively.

AISC (2005) has recently incorporated an exception to this limit by allowing braces with KL/r <

200 and greater than Equation 2.1.4 if adequate compressive capacity is supplied by the adjoining

columns. This seeks to incorporate the positive affect of large slenderness ratios on brace

ductility.

As mentioned in the previous section, connections in SCBF systems (between the brace, beam

and column) are susceptible to several brittle modes of failure, including weld failure and net-

section fracture at the end of the slotted brace. To prevent these types of failure, the Seismic

Provisions (AISC, 2005) require the design of these adjoining connections to be based on the

maximum force that the system can transfer to the connection. Although the Seismic Provisions
17

allow calculation of this force based on pushover or nonlinear time history analyses, they

recognize that (quoting from Section C13.3) – “In most cases, providing the connection with a

capacity large enough to yield the member is needed because of the large inelastic demands

placed on a structure by a major earthquake.” Consequently, these connections are typically

designed for forces corresponding to the expected yield force of the brace as described in

Equation 2.1.5 –

Py = Ry Fy Ag (2.1.5)

Where Py is the expected yield force of the brace, Fy is the minimum specified yield stress of the

material and Ag is the gross cross-sectional area of the bracing member. The ratio Ry between the

expected yield stress and the minimum specified yield stress of the member material recognizes

that the expected strength will typically be larger than the minimum specified strength. Typical Ry

values are in the range of 1.1 to 1.6 (Liu et al, 2007).

2.1.3 STATE-OF-THE-ART IN SCBF PERFORMANCE ASSESSMENT

To describe the current state-of-the art in SCBF system and component evaluation, the following

sections present experimental and analytical results, as well as fracture predictive approaches

from various investigations focused on the performance of SCBF systems. In general, an

emphasis is placed on brace component behavior, where buckling induced fracture is a concern at

the center of the brace. The literature review is presented by first considering results from SCBF

demand assessment studies, followed by investigations which quantify SBCF ductility capacity.

Given this format, it is useful to first reflect on the motivation, to separate demands from

capacities for structural/earthquake engineering simulations and the resulting dependence on post-

processing techniques for performance evaluation.


18

In the overall context of the theme of “demands” and “capacities”, it is interesting to discuss that

an ideal simulation of structural response should include the direct simulation of each physical

phenomenon (local buckling, fracture initiation, fracture propagation, etc.), leading up to the

complete failure of the system. In this ideal scenario, the simulation would describe the state of

the structure after the loading event and little post-processing (i.e., such as checking the demands

against capacities) effort would be required. The necessity for a post-analysis comparison

between demands and capacities (or limit states) arises from the lack of sophistication of

modeling techniques and their inability to simulate complex phenomena. This lack of

sophistication is an important issue in earthquake engineering where structures undergo large

inelastic deformations, accompanied by various forms of buckling and fracture. While some of

these aspects are routinely incorporated in structural analysis programs, events such as fracture

are typically evaluated by comparing the analysis results (cumulative deformation, maximum

drift, maximum strain, etc.) with experiment-based empirical relationships or other models that

seek to describe the facture ductility of the material or component. This approach is advantageous

in that fracture does not need to be explicitly modeled as part of the analysis, thereby significantly

reducing the computational expense of the analysis model while providing reasonable predictions

of fracture. However, the accuracy of this approach relies on (1) Accurate simulations of

structural behavior leading up to fracture, including effects such as local buckling, etc, that drive

the fracture strains (2) General fracture models that can accurately predict fracture based on these

localized strains and (3) The accurate characterization of local material properties.

Considering the difficulty of simulating complex phenomena leading to fracture events in large-

scale systems, SCBF performance has been largely characterized through a comparison between

earthquake demands from dynamic analysis techniques and ductility capacities obtained primarily

through experimental techniques. Based on these simulations and experiments, several empirical

and semi-empirical approaches have been suggested to predict the fracture response of braces in
19

these systems. These models represent important advances in brace fracture predictions and they

are reviewed in detail in this section. However, these approaches may be difficult to generalize to

situations different than the experiments used to calibrate them. Braced-frame shake-table tests

are also reviewed in the context of assessing earthquake demands and ductility capacities of the

system, components and connections. Note that shake-table tests experimentally combine demand

and capacity analyses as an assessment of structural behavior and performance is simply the

investigation of the state of the structure following the applied ground motion. Along these same

lines, there has been recent work on incorporating fracture predictions during simulations, thereby

illustrating the “on-the-fly” effect of fracture on structural demands and behavior. However,

suitable analysis and fatigue-fracture prediction models which consider the mechanisms leading

to brace failure are critical elements of such simulation methods.

2.1.3.1 STUDIES EXAMINING SEISMIC DEMANDS IN SCBF SYSTEMS THROUGH NONLINEAR DYNAMIC

ANALYSIS

Several analytical studies have investigated earthquake-imposed demands on SCBF systems

through nonlinear time history dynamic analyses. In this section, three analysis studies on 3 and

6-story SCBF systems, designed according to American design codes (FEMA 1997, AISC 1997),

are reviewed along with three analysis studies on Canadian code (NRCC 1990 and 2005, CSA

1989 and 2005) compliant braced-frames. The analysis results of the US designed frames are

presented first followed by the Canadian designed frames. After the analysis results are presented,

experimental results from a braced-frame shake-table test are presented in the context of

earthquake demands. Finally, a synthesis of these results is provided to compare and contrast the

different studies.

Interestingly, the studies by Sabelli, Uriz and McCormick et al use matching frames and ground

motions with different modeling techniques for the braces. The results from these three analysis
20

programs provide some background for the later discussion on brace modeling approaches

(section 2.2).

2.1.3.1.1 SABELLI (2001) ANALYSES

The design and analysis of the braced-frames for this study is based on a site-specific design for

downtown Los Angeles following the NEHRP Recommends (FEMA, 1997) and the AISC

Seismic Provisions (AISC, 1997). The reader is directed to Sabelli (2001) for the detailed design

of the structures. A suite of 20 horizontal ground motions are used in the time history analysis of

a 3-story and 6-story inverted-V SCBF. The ground motions correspond to the design level

earthquake hazard intensity of 10% chance of exceedance in 50 years, determined according to

the elastic spectral acceleration at the first period of the structure. The reader is referred to

Somerville et al (1997) for more information on the ground motions.

The analyses were performed using the SNAP-2D structural analysis program (Rai et al, 1996),

where the braces were modeled with axial truss members. The axial struts were assigned a

phenomenological (rule-based) uniaxial hysteretic response based on experimental work by Black

et al (1980). The uniaxial force-deformation model depends primarily on the brace slenderness

and other cross-section properties. Fracture was modeled through an empirical cycle-counting

scheme. Upon fracture prediction as per this scheme, the member was removed from the

simulation model.

Referring to Table 2.1 and the 10/50 hazard level, Sabelli (2001) reported mean maximum story

drift values of 3.9 and 1.8% (standard deviations of 3.1 and 0.8) for the 3 and 6-story SCBF,

respectively.
21

2.1.3.1.2 URIZ (2005) ANALYSES

Uriz re-analyzed the previous 3-story and 6-story inverted-V SCBF structures designed by Sabelli

(2001) with the Open System for Earthquake Engineering Simulation (OpenSEES) (McKenna

and Fenves, 2004) analysis platform. The analyses of Uriz are more sophisticated as compared to

those of Sabelli. Nonlinear fiber-based elements were used to model the bracing members where

an initial camber (or member sweep) was assigned to the elements to promote global buckling. In

addition to the design level, 10% chance of exceedance in 50 years, ground motion suite, the

study employed two additional suites of ground motions corresponding to earthquake hazard

intensities of 50 and 2% chance of exceedance in 50 years. While the 10/50 hazard level is a

typical design level event, the 2/50 hazard is the Maximum Considered Earthquake (MCE) level

(IBC, 2006). Table 2.1 reports the results of the nonlinear dynamic analyses which used the sets

of scaled ground motions (10/50 and 2/50 hazard levels) developed for the SAC study (see

Somerville et al, 2007). The Table also distinguishes between analyses which incorporated a

fatigue model to track brace fracture events. The model was applied at the fiber level and once the

fracture limit state was reached, the fiber was removed, effectively diminishing the cross section.

Referring to Table 2.1 and the 10/50 ground motion analyses without the brace fatigue model,

Uriz (2005) reported median maximum story drifts of 1.5 and 1.4% (standard deviations of 0.9

and 0.8) for the 3 and 6-story SCBF, respectively. By modeling brace fracture during the analysis,

the maximum story drift recordings remain approximately the same at 1.6 and 1.1% (0.9 and 0.6)

for the 3 and 6-story SCBF, respectively. At the MCE level (2/50), the median maximum drifts

increase substantially to 5.7 and 5.1% (3.0 and 3.4) without the fatigue model and 5.7 and 4.4%

(2.4 and 2.2) with the fatigue model for the 3 and 6-story SCBF, respectively. Note that several

analyses (3 for the 3-story and 6 for the 6-story) using the 2/50 ground motions revealed brace
22

fracture, as predicted by the built-in fatigue model, and eventual structural collapse. These

analyses are not included in the calculation of the median drifts reported in Table 2.1.

In addition to the maximum story drift data from nonlinear time history analyses, it is also

interesting to note the unsymmetrical response in assessing the demands on SCBF systems and

components. Unlike moment frames, where the response is more symmetrical (under far-field

ground motions), braced frames tend to show a more unsymmetrical response, presumably

because of the unsymmetrical strength and stiffness properties once the compressive brace

buckles. To illustrate this, Figure 2.6 shows a typical story drift time history from the 3-story

SCBF system during a 2/50 ground motion. The important observation is that the response is

highly unsymmetrical, such that the θmax is more than two times θmin, i.e. θmin = 0.4θmax for this

particular story and ground motion. Figure 2.7 illustrate this point further by plotting the

minimum story drift, θmin, versus maximum story drift, θmax, for each ground motion and story of

the 3-story and 6-story SCBF, respectively (Uriz, 2005). On average, the minimum story drift is

shown on the figures as approximately 40% of the maximum story drift for both frames and the

sixty LA-based ground motions. This behavior is notable because braced frame components (such

as braces) are typically subjected to symmetric loading protocols, based on adaptations of

protocols designed to reflect demands in moment frames (Fell et al, 2006, Han et al, 2007,

Shaback and Brown, 2003 and Archambault et al, 1995). Consequently, when the maximum

equivalent drift is reported as a capacity measure, it is calculated as half the range of equivalent

drift applied to the component (e.g. Uriz, 2005). With reference to Figure 2.7, this may

underestimate the capacity of the brace which, in general, will not be subjected to symmetric

cycles under seismic excitation. This issue will be discussed in more detail in Chapter 6 which

looks at the capacity characterization of bracing components.


23

2.1.3.1.3 MCCORMICK ET AL (2007) ANALYSES

McCormick et al (2007) used the same 3 and 6-story SCBF systems as Sabelli (2001) and Uriz

(2005). Similar to Uriz (2005), the analyses were performed using the OpenSEES platform.

Nonlinear fiber-based elements are used for the beams and columns while the braces are assigned

a phenomenological model to describe the hysteretic response. This is similar to the approach by

Sabelli (2001) where the hysteretic model depends on the brace slenderness and other cross-

section and material properties. Note that a fatigue-fracture model for the bracing members was

not used in this study.

Referring to Table 2.1, results from nonlinear analyses using the 10/50 and 2/50 ground motions

by Somerville et al (1997) are reported by McCormick et al. For the design-level, 10/50, event

mean maximum story drifts are listed as 3.57 and 1.97 (standard deviation of 1.61 and 0.68) for

the 3 and 6-story, respectively. For the MCE-level, 2/50, event, the corresponding maximum

drifts are 8.13 and 4.67 (standard deviation of 3.03 and 2.70).

2.1.3.1.4 REDWOOD ET AL (1991) ANALYSES

Redwood et al analyzed several 8-story braced-frames designed as per CAN/CSA-S16.1-M89

(CSA, 1989) and the 1990 edition of the National Building Code of Canada (NRCC, 1990). At

the time of this study, three categories of braced-frame systems existed in the Canadian code

corresponding to the expected ductility; these were (i) Ductile Braced Frames (DBF) (ii) Nominal

Ductility Braced Frames (NDBF) and (iii) Braced frames with no special ductility provisions

(SBF). In the context of our discussion on SCBFs, only the DBF results are summarized here.

While the detailed design is provided in Redwood and Channagiri (1991), it should be noted that

a considerably smaller R-factor (or strength reduction factor) is used in the design of the

Canadian DBF as compared to the American designed SCBF from above.


24

The 8-story DBF was analyzed using DRAIN-2D (Kannan and Powell, 1973) where the bracing

elements were modeled with axial truss elements and a phenomenological model (Jain and Goel,

1978) to describe the nonlinear hysteretic behavior. A set of ten earthquake ground motions were

chosen, primarily from recordings in the western United States, and scaled by the Peak Ground

Velocity (PGV) to correspond to a design level event for Vancouver, British Columbia. In terms

of Peak Ground Acceleration (PGA), the scaled earthquake records varied from 0.2g to 0.6g.

Note that from UBC (1997), the design PGA for zone IV, soil profile type SD, is approximately

1.1g. Thus, considering the smaller R factor used for the design and the less intense ground

motions, as compared to the previous 3 studies, these results are not directly comparable to SCBF

behavior but are included here for completeness.

From the nonlinear dynamic analyses, the largest story drift values were recorded on the seventh

story of the 8-story DBF. At this story, the mean maximum story drift across all ten ground

motions was listed as 0.7% (standard deviation of 0.88) by Redwood et al, with a maximum drift

equal to 1.1%. The drifts of the other stories ranged from approximately 0.3 to 0.6%.

2.1.3.1.5 TREMBLAY AND PONCET (2005) ANALYSES

Tremblay and Poncet (2005) analyzed several multi-story concentrically braced frames designed

as per the 2005 edition of the National Building Code of Canada (NRCC, 2005). The primary aim

of the study was to evaluate mass and geometric irregularity affects on behavior. For comparison

with the previous investigations, only the control structure with regular mass and geometric

distributions is considered here. This structure is an 8-story braced-frame with X bracing

members in each story. Similar to the Redwood et al (1991) study, DRAIN-2D (Kannan and

Powell, 1973) is used to analyze the structures where the bracing elements are modeled with axial

truss elements and a phenomenological model (Jain and Goel, 1978).


25

A suite of ten earthquake ground motions, scaled to a design-level event for Vancouver, British

Columbia, was used in the analyses. After scaling, the PGAs ranged from 0.3 to 0.6g across the

ten records. In comparison, PGA values for the Zone IV as per the 1997 UBC design spectra

(corresponding to approximately a 10/50 hazard) disregarding near fault effects are on the order

of 1.1g. The largest drifts were consistently recorded at the 6th story of the structure. The mean

maximum story drift at this story was approximately 1.7% (standard deviation of 0.5). The

maximum drift across all ground motions was listed as 2.7% and also occurred at the 6th story.

2.1.3.1.6 IZVERNARI ET AL (2007) ANALYSES

Izvernari et al (2007) presents results from analyses on several braced-frames designed according

to the 2005 edition of the National Building Code of Canada (NRCC, 2005) as per the Limited

Ductility (LD) and Moderate Ductility (MD) categories. The LD structures are not discussed

here. A total of 5 frames with 2, 4, 6, 8 and 12-stories were analyzed with the OpenSEES

platform where, similar to Uriz (2005), the braces were modeled with fiber-based elements. A

suite of twenty ground motions (ten historical and ten simulated) were used in the nonlinear

analysis of the frames. Referring to Table 2.2, the median drift recordings range from 1.1 to 1.8%

for the five frames, where the shorter 2 and 4-story frames had smaller drifts than the 8 and 12-

story frames.

2.1.3.2 SHAKE TABLE TESTING

Shake-table experiments can provide valuable earthquake demand data for structural systems,

while at the same time, evaluating the capacity of components and connections. However, it

should be noted that shake-table experiments only provide a single performance data point due to

the single ground motion used to damage the structure. While the experiments are visually

appealing and quite exciting, the results should be viewed objectively as the performance of a

single system during subjected to a single ground motion.


26

Tang (1987), Foutch et al (1987) and Roeder (1989) report findings from a 6-story braced-frame

shake table test conducted in Tsukuba, Japan. The design of the frame was based on both US

(UBC, 1979) and Japanese (Watabe and Ishiyama, 1980) building codes and lacked the detailing

of current design codes (AISC, 2005). For example, gusset plates were not provided at the ends of

the inverted-V bracing members as the braces were welded to the flanges of the beams. Table 2.3

lists maximum story drift measurements from a shake-table test using the Miyagi-Ken-Oki

earthquake scaled to 0.51g, approximately two times that of the “moderate” intensity level

experiment. A corresponding hazard level of the scaled ground motion was not discussed (Tang,

1987). A maximum drift prior to any fracture event was 1.6% in the third story of the frame,

which is within the range of the analytical results of Uriz (2005) and Izvernari et al (2007)

discussed previously with 10/50 simulation-based mean/median maximum drifts between 1.1 and

1.7%. However, while evaluating this comparison it should be noted that U.S. (and Japanese)

design codes have changed considerably between the time of the tests and the nonlinear time

history analyses. After fracture in stories 2-5, the maximum drifts increase significantly due to the

softening affect of fracture. This is most noticeable in story 3 where a maximum drift of 2.5% is

recorded after complete brace fracture and strength loss.

2.1.3.3 SUMMARY OF DEMAND ASSESSMENT FOR SCBF SYSTEMS

Current SCBF design requirements state that “braces could undergo post-buckling axial

deformations 10 to 20 times their yield deformation” (AISC, 2005). Given a system yield level

drift of approximately 0.3% to 0.5% (corresponding to initial brace buckling), the Seismic

Provisions may be interpreted as desiring a deformation capacity of approximately 3% to 5% for

SCBF systems. In this context, the story drift demands reported by Sabelli (2001), Uriz (2005)

and McCormick et al (2007) for the 10/50 (design) event in Table 2.1 range from 1.5 to 3.9% for

the 3-story frame and 1.1 to 2.0% for the 6-story frame. At the MCE, or 2/50, level the drift

demands increased substantially to 5.7 to 8.1% and 4.4 to 5.1% for the 3 and 6-story frame,
27

respectively. Thus, for design-level events, the 10/50 analyses seem to corroborate the expected

demands of the current AISC Provisions. On the other hand, at the MCE level, these results

suggest that SCBF earthquake demands could exceed 5% drift.

The discrepancy between the results of Table 2.1 should be noted, especially considering the

same structures and ground motions were used. This highlights the significance of robust

structural analysis techniques that can accurately characterize response. The different brace

modeling techniques, discussed more in section 2.2, between the three investigations may be

responsible for these differences.

The next section of the literature review discusses experimental studies focused on the cyclic

fracture/fatigue resilience of bracing members.

2.1.3.4 STUDIES INVESTIGATING THE CAPACITY OF BRACING MEMBERS

Where the previous discussion focused on braced-frame demand assessment, this section presents

experimental results which characterize the capacity of SCBF systems and bracing members. Of

primary interest is brace component behavior, where buckling-induced fracture at the middle

hinge could lead to eventual failure. The ductility of bracing components, often expressed in

terms of a maximum, or cumulative axial deformation can be compared to story drift demands

(discussed above) through simple kinematic relationships. While one such relationship is

introduced in Chapter 3, the primary aim of this section is to review experimental work which has

characterized the ductility of inelastic buckling braces. Many of the experimental studies

presented here have sought to generalize the experimental results with fatigue-fracture models of

varying degrees of complexity. However, these will be the focus of a later section. Furthermore,

while experimental investigations on small-scale (for example, 1x1 inch cross-sections) bracing

members have been a popular means to identify important trends in ductility and strength
28

requirements (Khan and Hanson 1976, Jain and Goel 1978, Jain et al 1980), they are not covered

in depth here. Given the highly nonlinear behavior of cyclic brace response, such as local

buckling-induced fracture (where the strains are sensitive to issues such as the length of the local

buckle), similitude might not exist between large and small scale experiments. Furthermore,

material properties and residual stresses within the walls of the cross-section may vary between

small and large-scale sections. Uriz (2005) provides an excellent summary of experimental work

with smaller bracing members.

As discussed previously and illustrated in Figure 2.2, inelastic buckling and yielding of bracing

elements serve as the primary seismic energy dissipation mechanisms in SCBF systems. To

provide background for the current discussion, it is important to consider the events leading to

bracing component failure during earthquake-type cyclic loading. Referring to Figure 2.8, the

first major limit state is brace buckling at point A, which is evident by large lateral deformations

and accompanied by flaking of the whitewash paint due to large strains associated with folding at

the end gusset plates and plastic-hinging at mid-length of the brace. As seen in region A-B,

buckling is followed by a sudden drop in load. However, as the compression in the brace reduces,

the response is mainly driven by bending of the buckled brace, resulting in a more gradual drop in

load after point B. During subsequent cycles, the localized yielding in the gusset plates and mid-

point hinge becomes more severe as the amplitude of compressive loading increases. Upon

reversed loading at point C, the stiffness gradually increases (point D) as the out-of-plane

deformations decrease. As the strut straightens, the out-of-plane deformations reduce, and the

brace yields in tension at point E. The subsequent compressive excursion results in a smaller

buckling load (see point F) as compared to the first buckling event due to the Baushinger effect,

increased brace length (from tension yielding), and residual out-of-plane deformations. After

repeated compression-tension loading cycles, a local buckle forms at the middle hinge, similar to

Figure 2.8c, which amplifies the plastic strain and triggers ductile fracture soon thereafter (Figure
29

2.8d). While these events are not evident on the load-deformation curve, the photos of Figure 2.8

are fairly representative of the local buckling and fracture initiation observed in most tests. Upon

further cycling, the rupture propagates in a ductile manner across the section, i.e., for square HSS,

the buckled face ruptures first as shown in Figure 2.8e. Finally, at some point during a subsequent

tensile excursion, the entire cross-section fractures suddenly, severing the brace. This progression

is typical in large-scale braced-frames and has been detailed across numerous studies.

Experimental investigations often compare the cyclic fracture capacity of bracing members to the

brace slenderness ratio, the width-thickness ratio, or a combination of the two. Thus, a general

trend of the following sections will be to discuss the capacity of bracing members with respect to

these two geometric properties. Generally, it is difficult to observe their independent affects as

cross-section compactness is linked with the slenderness ratio through the radius of gyration. This

difficulty provides context for subsequent chapters as 1) there is a lack of general models to

explain brace ductility through geometry properties and 2) continuum-based parametric studies

can be used to expand the experimental date presented in this and future studies.

The current discussion on brace capacity begins with a review of component tests on HSS bracing

members, the most commonly used section in SCBF design. While square and rectangular HSS

have been the most wide-spread shape used in concentrically braced frames, their ability to

provide adequate performance has recently been questioned due to their perceived poor

performance during cyclic loading. Thus, results from Pipe and Wide-Flange experiments are

presented and reviewed as they have gained increased popularity over HSS shapes as bracing

members in SCBF systems. Finally, frame tests – either static or dynamic – are presented to gain

an understanding of system level capacity and performance.


30

2.1.3.4.1 SQUARE AND RECTANGULAR HOLLOW STEEL STRUCTURAL (HSS) BRACING

COMPONENTS

The various experimental programs on square and rectangular HSS bracing members are

summarized in Table 2.4. While a more detailed synthesis of these experiments is the topic of

Chapter 6 (combined with the full-scale tests from this investigation – presented in Chapter 3 and

4), this section provides a general review of experimental research on HSS bracing members.

Given that brace ductility is a function of several parameters, it is difficult to discuss specific

conclusions across all investigations without a common basis that can be used to interpret data

from various test programs, which may have different brace lengths and cross-sectional

dimensions. Thus, the following discussion introduces experimental research on HSS bracing

members by highlighting several important trends that have been reported in the context of HSS

brace ductility.

In general, the findings from all programs listed in Table 2.4 concur that width-thickness and

slenderness ratios have the most significant effect on brace ductility, such that higher width-

thickness ratios and lower slenderness are detrimental to brace performance. For example, Han et

al (2007) found that for approximately the same slenderness ratio (KL/r ≈ 80), increasing the

width-thickness ratio by 107% (from 13.7 to 28.3) resulted in a decrease in fracture ductility by

75%. In regard to slenderness influence, Tremblay et al (2003) reported that increasing the

slenderness by 60%, while holding the width-thickness ratio constant at 25.7, increased the axial

deformation ductility by 80%.

The investigations by Gugerli, Liu, and Lee with Goel (1982, 1988 and 1988, respectively) and

later Zhao et al (2002) were focused on the relative performance of concrete-filled HSS to

unfilled HSS bracing members. The concrete acts to prevent the inward local buckling observed
31

in most hollow sections, producing a less severe local buckling shape as the cross-section walls

buckle outward. In general, it was concluded that concrete fill has a beneficial effect on ductility

by delaying cross-section local buckling. For example, Lee and Goel (1988) reported a maximum

axial deformation during cyclic loading of 1.7 inches for an unfilled HSS cross-section, whereas

the same concrete-filled brace survived a maximum deformation of 2.6 inches, fracturing

approximately 8 cycles after the unfilled brace.

Archambault et al (1995) and Tremblay et al (2003) used the effective brace length to inform

design provisions on properties such as compressive strength degradation, energy dissipation and

brace fracture resistance. A kinematic relationship was also proposed to predict out-of-plane

buckling deformations. The investigation found that using the brace effective length (KL) resulted

in an accurate prediction of the compressive buckling load and, further, that as the slenderness

increases the energy dissipation capacity of the brace decreases linearly. Lastly, an empirical

fracture resistance equation was proposed, which is proportional to the brace global slenderness

and inversely proportional to the cross-section compactness ratio. Thus, the equation is

descriptive of the behavior that one would expect from the qualitative depiction shown in Figure

2.5, and employs data from all of the studies prior to 1995 (in Table 2.4) to calibrate various

constants of the model. The relationship and its accuracy will be discussed in more detail in

subsequent sections. Shaback and Brown (2003) present similar data and conclusions on HSS

bracing members as those of Tremblay et al (2003) concerning out-of-plane deformations, energy

dissipation and compressive buckling strengths, while employing a different empirical fracture

model that had been developed previously by Lee and Goel (1988) and Tang and Goel (1987).

Similar to the approach of Tremblay et al (2003) the fracture relationship is also a function of

slenderness and cross-section compactness while using a cumulative deformation measure to

assess capacity. This model will also be reviewed in section 2.2.2. While the experimental

investigation on bracing components by Yang and Mahin (2005) provides valuable data on cyclic
32

brace performance, the series of tests were conducted primarily to study the performance of the

slotted brace, gusset-plate end connection.

2.1.3.4.2 ROUND STEEL PIPE BRACING COMPONENTS

As mentioned earlier, there has been a recent shift from using square and rectangular HSS cross-

section shapes for bracing members to Pipe and Wide-Flange shapes. In part, the restrictions on

square HSS width-thickness ratios (i.e., Equation 2.1.1) have limited the number of shapes that

are available during design, making Pipe and Wide-Flange shapes a more attractive option.

Moreover as mentioned in the previous section, the cold working of square HSS sections during

fabrication was thought to significantly reduce the fracture toughness of the corner material,

driving fracture initiation at the corners first. Thus, the more gradual bend radii of Pipes and the

rolling process of Wide-Flange sections was thought to provide tougher material as compared to

square or rectangular HSS. This section, and the next, will briefly discuss several experimental

investigations on these alternative shapes.

At the time of the writing of this dissertation, an exhaustive study on steel Pipe bracing members

was being conducted by Tremblay et al at École Polytechnique, Montréal. The study is expected

to carry significant importance as, compared to square HSS members, there are relatively few

studies that have investigated the performance of round Pipes. However, there have been several

other studies that have looked at the cyclic behavior of large-scale Pipe bracing members. One of

the earliest studies on the cyclic behavior of steel Pipe sections was by Popov and Black (1981).

Although the reported experimental results focused more on compressive strength degradation

rather than on fracture events, an important finding was that the more compact Pipe4STD (width-

thickness ratio = 19, approximately 0.52 times the current maximum ratio recommended by

AISC, 2005) and Pipe4X-Strong (width-thickness ratio of 13.4, 0.37 times the current maximum

limit) braces locally buckled at very large axial deformations.


33

More recently, Elchalakani et al (2003) conducted a series of twenty steel pipe braces under

earthquake-type cyclic loading. The testing matrix for the braces tested to fracture included a

variety of different width-thickness ratios – ranging from a D/t ratio from 9 to 21 – with

slenderness ratios of approximately 24 or 35. Surprisingly, the results suggested that, when

grouped by slenderness, the less slender braces (those with KL/r ratios of approximately 24)

fractured later in the loading history and at larger axial deformations than the braces with a larger

slenderness ratio. Furthermore, the investigation concluded that brace fracture ductility is “more

sensitive” to the cross-section width-thickness ratio as compared to the slenderness ratios. While

the results consistently confirm the less slender braces are more ductile, it should be noted that

the study queried a much larger range of D/t ratios (9 to 21) as compared to brace slenderness

ratios (24 to 35). Nonetheless, the study provides valuable fracture capacity data that can be used

to assess the performance of steel Pipe bracing members.

2.1.3.4.3 WIDE-FLANGED BRACING COMPONENTS

Similar to steel Pipe cross-sections, Wide-Flanged braces have becoming increasingly popular in

recent years due to their perceived superior ductility over HSS members during earthquake

loading. However, there are several challenges with using Wide-Flanged sections in braced-frame

construction. First, novel gusset plate end-connection details, not discussed as a part of this

chapter (see Chapter 3 for an example connection detail), must be devised to ensure out-of-plane

buckling about the member weak axis. These connection details can be quite complex as, unlike

HSS or Pipe sections, traditional slotted-end connections are not feasible for Wide-Flange shapes.

Moreover, the number of Wide-Flanged shapes that can be used as bracing members is somewhat

limited to shapes with small to moderate tensile yield to buckling load ratios. If not selected

correctly, the section properties can create a large force imbalance on the beam (in the case of

chevron-type configurations) and a reduced energy dissipation capacity. While this is accounted
34

for through Equation 2.1.4 (AISC, 2005) it does have the effect of limiting the shapes that can be

used as bracing components.

Popov and Black (1981) and Gugerli and Goel (1982) conducted a series of experiments on

Wide-Flanged bracing members. Popov and Black tested Wide-Flanged members with width-

thickness ratios ranging from 5 to 8.25 with slenderness ratios from 40 to 120 yet provide little

detail on the relative performance of the members as related to fracture ductility. Gugerli and

Goel present brace fracture capacities for very slender (KL/r from 95 to 175) Wide-Flanged

bracing members with width-thickness ratios between 7.4 and 11.5. As expected, these results

show that ductility improves with increasing slenderness and decreasing compactness.

2.1.3.4.4 GENERAL DUCTILITY TRENDS OF HSS, PIPE, AND WIDE-FLANGED STEEL BRACES

Figure 2.9 compares experimental results from three separate investigations by Shaback and

Brown (2003) on square and rectangular HSS, Elchalakani et al (2003) on round Pipe, and

Gugerli and Goel (1982) on Wide-Flanged cross-sections. The data is presented here for example

purposes to identify issues and questions with respect to assigning brace ductilities based on

slenderness and/or compactness. Furthermore, as each study applies the same loading history to

each brace, the results within each cross-section type can be discussed without considering the

influence of loading history. The shake-table and quasi-static test results from Tang (1987) and

Uriz (2005), respectively, are also shown but will be discussed in the following section. Figures

2.9b and 2.9c illustrate the influence of width-thickness and slenderness ratio, respectively, on the

maximum deformation range in tension and compression (shown in Figure 2.9a) of the brace

prior to failure, Δrange, normalized by the yield deformation, Δy. The brace width-thickness and

slenderness ratios are normalized by the AISC (2005) limits described previously. It is important

to note that the brace component loading protocols varied between the three investigations where

a symmetric history in tension and compression was used for the Pipe tests, a slightly biased
35

compressive loading history for the HSS, and a compressive dominated history for the Wide-

Flanged tests. While a cumulative deformation quantity, such as energy dissipation or cumulative

plastic deformation, may be a more descriptive measure in comparing relative brace capacities,

the maximum range sustained by the brace prior to failure (Figure 2.9a) is used here as this can be

most readily transferred to a simplified maximum story drift. As illustrated in the demand section

(see Table 2.1-2.3), this quantity is most often used when discussing structural system demands

and will set the context for later chapters.

Referring to Figure 2.9b, as the width-thickness ratio (normalized by the respective AISC, 2005

limits) increases within a group of braces with similar slenderness ratios, the maximum

deformation capacity tends to decrease. For example, the Pipe braces with normalized slenderness

ratios of 0.2 (solid circles) show a definite downward trend for increasing width-thickness ratio.

This trend is also observed for HSS and Wide-Flange members when grouped according to

relative slenderness ratios. In general, the experiments selected for this example confirms the

current design methodology of placing upper-bounds on section width-thickness ratios (AISC,

2005).

Figure 2.9c illustrates the influence of slenderness ratio on maximum deformation capacity using

the same data points. The braces are grouped according to relative width-thickness ratios,

facilitating a comparison between ductility capacities as a function of slenderness. The figure

shows an apparent trend between increasing axial deformation ductility and slenderness ratio.

Interestingly, slenderness also appears to mitigate unfavorable width-thickness ratios, as the

compactness of the experiments presented here seems to decrease, on average, as slenderness

increases. However, referring back to Figure 2.9b, the influence of brace slenderness within the

Pipe and HSS sections is questionable. For example, a relative slenderness ratio decrease from

0.62 (solid squares) to 0.53 (hollow squares) for HSS members does not seem to influence
36

ductility, rather it is mostly controlled by the width-thickness ratio. Furthermore, the Pipe cross-

sections show that increasing slenderness (from 0.2 to 0.3) acts to decrease ductility. However, it

should be mentioned that the HSS and Pipe specimens sample a smaller range of slenderness

ratios as compared to the Wide-Flange braces which have a much larger range of slenderness

ratios (1.0 to 1.8). Thus, the investigations within the HSS and Pipe experiments are somewhat

limited with respect to slenderness affects. Considering the trends presented in Figure 2.9, several

key concerns and knowledge gaps can be raised pertaining to the current state-of-the-art brace

capacity assessment procedures –

1. Given that brace slenderness tends to influence the axial deformation capacity of a brace

during inelastic cyclic loading, should design codes restrict the use of stocky members in

braced-frame construction? For example, in the case of the steel Pipe specimens shown in

Figure 2.9, the axial deformation capacity is severely reduced as compared to the more

slender HSS and Wide-Flanged braces. However, according to AISC Seismic Provisions,

the Pipe members have lower relative width-thickness ratios as compared to the HSS and

Wide-Flanged members. Thus, while the experiments suggest otherwise, the code deems

Pipe members as “more acceptable” compared to the HSS and Wide-Flanged braces

presented here.

Furthermore, should design codes restrict all braces to meet the same compactness

requirement (within cross-section type)? According to the data of Figure 2.9, doing so

may unnecessarily penalize cases where slender braces would perform well, even with a

large width-thickness ratio. For example, the large slenderness ratios (or possibly the

nature of the local buckling deformation) of the Wide-Flanged members presented in

Figure 2.9 seems to mitigate the competing influence of a large width-thickness ratio.
37

2. Does the maximum axial deformation capacity of individual bracing components, used in

Figure 2.9, directly translate to a system level ductility measure? Experimental brace

capacities are often expressed in terms of axial deformation capacity, but few have

translated this into a story drift capacity. This would facilitate a comparison between

earthquake system level demands and component ductility levels.

3. With respect to the first point, can the interactive influence of brace slenderness and

width-thickness ratio be rationally evaluated without relying on a fully empirical,

experiment-based approach? While several relationships have been developed (presented

in the following sections) to describe the ductility capacities for HSS shapes, they are

largely empirical and are not derived from a fundamental mechanics-based approach.

Thus, it could be beneficial to develop a simulation-based methodology that could

investigate the relationship between brace ductility and the governing geometric (or

material) brace properties. Although not explicitly mentioned previously, the brace

length, for example, is often constrained in most experimental due to out-of-plane testing

restraints. Furthermore, actuator force capacity and tests setup, not to mention economics,

can limit the scale and breadth of large-scale brace investigations.

4. Is brace capacity influenced by loading history? With reference to Figure 2.9, the Wide-

Flanged members survived the largest deformation capacity prior to failure, yet were

subjected to a compression dominated loading history (refer Gugerli and Goel, 1982)

with relatively small tensile excursions. Furthermore, the least ductile HSS and Pipe

braces were tested with a slightly biased compressive and standard symmetric loading

history with (i.e., equal compressive and tensile excursions as illustrated in a typical

symmetric hysteretic response in Figure 2.9a), respectively. It could be envisioned that a

symmetric history could be more severe as the tensile action during local buckle

straightening could severely increase the cumulative plastic strain in the middle plastic

hinge region.
38

These issues are specifically addressed in subsequent chapters.

2.1.3.4.5 FULL-SYSTEM BRACED-FRAME TESTS

This section presents results from two experimental investigations on large-scale CBF systems.

The first is the shake-table test on a 6-story CBF that was discussed previously in the demand

section and reported in Tang (1987), Foutch et al (1987), and Roeder (1989). The second is a

static test on a 2-story SCBF system tested as part of the work by Uriz (2005).

Listed in Table 2.3 and illustrated in Figure 2.9 are the approximate brace ductility capacities

from the 6-story shake-table test reported in Tang (1987). Interestingly, the axial deformation

capacities (Δrange/ΔY) prior to brace fracture (stories 2 through 5) are similar to the capacities

discussed in the previous section from brace component tests. This is a valuable comparison as it

provides some justification to apply static, component results to inform large-scale system

behavior. The story drifts recorded prior to the onset of brace fracture range between 1.1 and

1.6%. However, as mentioned previously, this frame lacked modern detailing requirements, such

as gusset plate connections and featured braces with b/t ratios that exceeded current width-

thickness limits.

Also shown in Figure 2.9 are the brace ductility capacities (in terms of Δrange/ΔY) for two braces in

the lower story of a 2-story SCBF by Uriz (2005). The second story braces did not fracture.

Similar to the shake-table results, these results are in line with the HSS component tests by

Shaback and Brown (2003), thereby supporting the use of component tests to inform system

behavior. One should also note the ability of the Δrange/ΔY to accurately describe the brace

ductility. This will be important in Chapters 6 where the maximum deformation range prior to

brace fracture is used to approximate ductility. The maximum story drift for the first story prior to

any brace fracture for the quasi-static test was approximately 1%. The results illustrated in Figure
39

2.9 lend validity to the numerous component tests described previously. Furthermore, and as

discussed in the following section, if brace component behavior can be simulated correctly, the

extrapolation to large-scale behavior can be relatively straight-forward considering these results.

2.2 MODELING TECHNIQUES FOR SCBF SYSTEMS

Simulating complex physical events has long been a focus of the academic community.

Simulation models reduce the necessity of exhaustive experimentation and extend experimentally

observed phenomena to cases which may be difficult to test. Thus, through the application of

theoretical and analytical models, routine design and construction need not rely on the testing of a

replicate structure. Doing so would be an impractical, and in many cases, an impossible feat. In

this light, the current section presents structural modeling techniques in the context of the braced-

frame system. Similar to the previous section, the discussion will be separated by first considering

demand modeling, and second, capacity modeling. The former contains models which aim to

characterize quantities such as deformations, strains and stresses from an applied load, ground

motion, etc. Capacity modeling will be treated as post-processing techniques which are applied

after the analysis and aim to describe the fracture ductility of the steel bracing members.

As discussed in the introduction to this chapter, performance assessment generally takes the form

of, first, developing an analysis model that captures well-known geometric or material behavior

and, second, post-processing the results to gain insights into a variety of different phenomena

which are not included as part of the analysis model. Phenomena investigated in the second stage

are typically more difficult to incorporate into the primary model due to computational expense, a

lack of scientific understanding, or a combination of both. For example, fracture propagation

under certain conditions can be computationally prohibitive if the model is the scale of an actual

building while the elastic modulus of any structural material can be accurately described and is

included in any analysis model. Thus, while the relationship between the elastic strain and stress
40

is contained in the primary model, most fracture assessments are made following the analysis by

comparing, for example, material strains (demands) to critical toughness parameters (capacity).

In this context, current state-of-the-art modeling approaches are presented by first discussing

simulation methods that aim to characterize earthquake demands on structural systems for steel

braced-frames. Second, models that describe the capacity (usually in terms of a fracture event) of

a braced-frame system or component are presented. Generally, the capacity models are somewhat

empirical in nature and serve the purpose of expanding, or providing an explanation of, trends

from a particular experimental data set. Other models utilize data across a wide-range of

experiments, but could still be classified as empirical in nature and tend not to address the actual

physical mechanisms that control brace failures. The work by Uriz (2005) is highlighted in the

following sections as an example where performance assessment is completed during the analysis

procedure by comparing strain demands to a critical strain measure, thus eliminating a post-

processing-comparison approach for the fracture limit state. While only applied to one material

type (and one brace cross-section type), the work provides context to discuss the advantages of an

“on-the-fly” performance-based framework and areas where the community could benefit from

advanced simulation capabilities.

2.2.1 DEMAND CHARACTERIZATION

This section presents modeling techniques to assess demand quantities (such as deformations,

strains, etc.) for bracing components subjected to earthquake-type loading histories. For purposes

of this discussion, it is assumed that earthquake demand assessment first begins with a site-

specific earthquake hazard analysis, followed by a ground motion scaling procedure (e.g. as

outlined in Haselton, 2006). Once the ground motion(s) are determined, they serve as an input

base excitation to the structure, with the response of the superstructure being determined through

the fundamental equations of dynamic equilibrium. While copious literature is available on the
41

subjects of hazard, risk and dynamic analysis procedures, this section is specifically focused on

determining the response a bracing member once the input deformations (or histories) are known.

These could either be ground motion-induced deformations or some form of a standard loading

history.

2.2.1.1 TECHNIQUES FOR BRACE SIMULATION

Based on the previous discussion describing braced frame response, and Figure 2.8, this section

describes various techniques that have been used to simulate braced-frame response (specifically

the response of the bracing elements themselves). Some of the techniques discussed are aimed

only at providing good simulation of brace load-deformation response, whereas other, more

sophisticated techniques can simulate complex aspects of response including local buckling and

fracture.

2.2.1.1.1 RULE-BASED (OR PHENOMENOLOGICAL) BRACE MODELS

Uniaxial spring elements with rule-based phenomenological, constitutive response are often used

to model bracing elements (Zayas et al 1980, Khatib et al 1988, Sabelli 2001, and McCormick et

al 2007). As illustrated in Figure 2.10a, hysteretic rules are applied to a single degree of freedom

(axial deformation in the case of bracing elements) with the resulting behavior described by

various linear segments, transitioning between different slopes according to empirical calibration

techniques. While convenient, this approach is limited because it does not directly model

buckling events, and thus, cannot provide insight into localized brace behavior. Furthermore,

calibration is often based on empiricism and thus, these models may be more difficult to

generalize to different materials, cross-sectional shapes or loading histories. On the other hand,

phenomenological models are advantageous if the analysis is primarily aimed at assessing

“global” demands, for example, maximum story drift or axial deformations.


42

2.2.1.1.2 CONCENTRATED HINGE BRACE MODELS

Elastic beam-column elements with lumped plasticity end-nodes are often used in structural

analysis programs where the inelastic behavior is known to concentrate at the ends of structural

members (for example, beam elements in moment frame systems). The inelastic behavior at the

end-nodes is simulated through stress-resultant plasticity formulations (Powell and Chen, 1986;

El-Tawil and Deierlein, 1998 and Hajjar and Gourley, 1977) which operate at the cross-sectional

level, i.e., through a P-M interaction curve as shown in Figure 2.10b. For example, Higginbotham

and Hassan (1976), Ikeda and Mahin (1986), and Hassan and Goel (1991) used variations of this

approach to generalize the behavior of buckling braces. It was shown that for general cyclic

loading, the behavior of the bracing elements at the global force deformation level is captured

fairly accurately. This technique is attractive because the key aspects of response (i.e. buckling

and the post-buckling geometric nonlinearities) are explicitly modeled, and calibration is less

subjective.

While these models are generally more robust than the phenomenological models discussed

previously, they are still limited by some degree of empiricism in calibration. Furthermore,

lumped plasticity elements may not accurately capture yielding, such as may occur over the entire

length of the brace during tensile excursions. Refinements to this approach include Jin and El-

Tawil (2003), based on the prior work of El-Tawil and Deierlein (2001). This research

incorporates the distribution of plasticity along the length and across the cross-section of the

brace.

2.2.1.1.3 FIBER-ELEMENT BASED BRACE MODELS

Recently, the more powerful and versatile fiber-based element has been used to simulate brace

response. In general, a fiber-based model is formulated with integration points along the length of
43

the element where the member cross-section is discretized by a fiber mesh as shown in Figure

2.10c. A constitutive model is defined at the fiber level and allows for strain and stress gradients

through the cross-section.

The fiber-based element formulation can directly model the spread of plasticity along the length

of the member as well as within the depth of the cross-section, whereas a lumped plasticity model

accounts for these effects only indirectly (see Jin and El-Tawil, 2003). Unfortunately, as the fiber

discretization exists at the cross-section level (at integration points along the element), the

element can not explicitly model localized affects along the length of the brace, such as the strain

amplification during local buckling. Thus, in the context of modeling bracing members, which

can develop significant local buckles during cyclic loading, the fiber-element is somewhat

inadequate for capturing localized stress and strain demands. However, Uriz (2005) and others

(Hanbin et al, 2007; Krishnan 2008-in review) have implemented schemes to track a critical

strain measure at the cross-section level of a fiber-element to predict fracture. The approach relies

on empirical calibration to account for the affects of local buckling, but has been shown to

provide reasonably accurate fracture predictions in bracing element, as discussed later.

2.2.1.1.4 CONTINUUM-BASED BRACE MODELS

Continuum finite element models, such as shown in Figure 2.10d and 2.11 constructed through

commercial or research finite element software (e.g., ABAQUS, 2004) can be applied to model

the brace response during cyclic loading. These simulations incorporate continuum (shell or

brick) elements, large displacement and deformation formulations, and continuum cyclic

constitutive response. In addition, by simulating local as well as global imperfections, these

simulations can directly model several complex phenomena (such as local buckling) that lead to

fracture. Although computationally expensive, the resulting simulations provide a high resolution

of the stresses and strains at the local buckle (e.g. Figure 2.11), from which ductile fracture
44

initiation is assessed using micromechanics-based models, such as ones proposed by Kanvinde

and Deierlein (2004 and 2007). Furthermore, the simulations provide interesting insights into

various damage mechanisms. For example, the simulation illustrated in Figure 2.11 indicates that

while the strains in the cross-section due to global buckling and bending alone are on the order of

0.02, the strains induced through the HSS wall thickness by local buckling are more than an order

of magnitude larger (≈ 0.6), thereby underscoring the importance of simulating local buckling

during brace cyclic response.

Continuum-based formulations of bracing elements are advantageous as compared to integration

point fiber-based modeling as the continuum model directly evaluates inelastic brace response at

each point along the length of the member as well as through the depth of the cross-section. Thus,

contrasted with fiber-based elements, a continuum approach allows localized deformation effects,

such as local buckling, to develop along the length the cross-section of the bracing member.

While solid finite element analyses present notable advantages over lumped-plasticity and fiber-

based models, there is a heavy computational expense that accompanies these models, especially

considering the complex events of cyclic inelastic compressive buckling and tension yielding of

bracing members during earthquake-type loading. Moreover, currently, fracture initiation and

propagation events are not typically modeled as part of the analysis routine in standard finite

element software packages. In large part, this can be attributed to a lack of fundamental models

that seek to capture the complex micromechanical events that trigger fracture initiation and

propagation. Furthermore, innovative re-meshing schemes often need to be employed at an

advancing crack tip as the fracture propagates through the continuum body (Khoei et al, 2008).

While mechanics-based methodologies have been developed to address these issues (Rao et al,

2007), they are typically applied to relatively simple geometries and loading conditions and, in

general, would add a significant computational expense to large-scale inelastic cyclic brace

analyses.
45

With the advent of general, physics-based initiation and propagation models, continuum analyses

provide the framework to eliminate the separation between demand and capacity evaluation.

However, these fracture models often describe complex micromechanical processes and need to

be rigorously evaluated prior to implementation into FEM analyses. One such initiation model,

which is presented through a post-processing demand versus capacity approach, will be presented

in the following section.

2.2.2 CAPACITY CHARACTERIZATION

The previous section illustrated several modeling techniques that can be used to simulate the

various aspects of inelastic brace behavior during cyclic loading, and the corresponding

deformation or strain demands in the braces. Given these demands, this section provides a brief

review of the various approaches that have been developed to evaluate the fracture ductility of

bracing members, relative to these demands.

2.2.2.1 BRACE GEOMETRY-BASED FRACTURE CAPACITY RELATIONSHIPS

Lee and Goel (1988) described the cyclic ductility capacity of HSS bracing members with (KL/r)

> (KL/r)critical in terms of a normalized (by ΔY) cumulative axial deformation, δf,pred, as per the

Equations below –

2
⎡ ⎛ F ⎞ 1 ⎛ b ⎞ a2 ⎤ ⎛ B
a a a
⎞ 3 ⎛ KL ⎞ 4
δ f , pred = ⎢C1 ⎜ Y
⎟⎟ ⎜ ⎟ ⎥ ⎜ + C2 ⎟ ⎜ ⎟ (2.2.1)
⎢ ⎝⎜ FY , meas ⎠ ⎝ t ⎠ ⎦⎥ ⎝ H ⎠ ⎝ r ⎠

and,

2
⎡ ⎛ F ⎞ 1 ⎛ b ⎞ a2 ⎤ ⎛ B
a a3 a4
⎞ ⎛ KL ⎞
δ f , pred = ⎢C1 ⎜ Y
⎟⎟ ⎜ ⎟ ⎥ ⎜ + C2 ⎟ ⎜ ⎟ (2.2.2)
⎢ ⎜⎝ FY , meas ⎠ ⎝ t ⎠ ⎥⎦ ⎝ H ⎠ ⎝ r ⎠critical

if (KL/r) ≤ (KL/r)critical, where C1, C2, and a1 – a4 are determined by an empirical fit of the

available HSS data, and a4 = 0 (i.e., slenderness was not considered) in the original formulation
46

by Lee and Goel (1988). The measured and specified yield stresses are included as FY,meas and FY,

respectively. In Equations 2.2.1 and 2.2.2, b/t is used generically to represent the brace width-

thickness ratio, such that b/t = B/t – 2, where B and t are the overall length of the buckling face

and the design thickness, respectively, as shown in Figure 2.51. The cross-section dimensions B,

H and t are the nominal dimensions of the square or rectangular brace and are also illustrated in

Figure 2.5. The axial displacement fracture prediction, Δf,pred = δf,pred.ΔY, is a cumulative measure

of brace ductility and is determined by summing the compressive and tensile normalized axial

deformations (defined in Figure 2.7a) up to failure of the brace –

Δ f,exp = ΔY ∑ ( 0.1δ compression + δ tension ) (2.2.3)

Several other experimental investigations on HSS bracing members have used the empirical

relationship proposed by Lee and Goel (1988) to calibrate the constants in Equations 2.2.1 and

2.2.2. Table 2.5 lists the results of the various calibration studies conducted on HSS bracing

members tested by Tang and Goel (1989), Archambault et al. (1995), and Shaback and Brown

(2003). It is important to note that Tang and Goel expressed ductility in terms of a standardized

number of cycles to fracture (Nf) instead of the normalized axial deformation, δf,pred, used by the

others (see Tang and Goel, 1989, for more details). Also, Shaback and Brown used both

Σ(0.1δcompression + δtension) and Σ(δcompression + δtension) to calibrate Equations 2.2.1 and 2.2.2 as they

argued that the “penalty” of 0.1 on the normalized cumulative compressive excursions is not

representative of the full deformation demands on the brace, especially considering the largely

compressive dominant loading histories applied by the Goel et al investigations. Due to these

discrepancies, Table 2.5 does not intend to suggest average calibration values for Equations 2.2.1

and 2.2.2; rather, the constants are supplied to provide a synthesis of the fracture predictions for

HSS bracing member which take the form of the above model. The accuracy of each model

compared to three separate testing programs will be discussed later.

1
Note: Since the original formulation of Equations 2.2.1 and 2.2.2, b/t is now generally listed as B/t-3.
47

Tremblay (2002) and Tremblay et al (2003) developed similar empirical relationships for the

cyclic fracture life of HSS bracing components. The first (Tremblay, 2002) provides a rigorous

synthesis of past experimental work on the cyclic performance of HSS bracing members. Thus,

the fracture model is calibrated across six separate experimental studies with a variety of loading

histories and section properties. Interestingly, the total fracture ductility, measured in terms of the

maximum compressive and tensile ductility, μcompression and μtension, respectively, was found to be

function only of the global brace slenderness ratio, λ, defined as –

KL FY ,meas
λ= (2.2.4)
r π 2E

Assuming a linear relationship, the expected ductility of an HSS brace component was found to

be –

μ f , pred = μcompression + μtension = 2.4 + 8.3λ (2.2.5)

or –

FY , meas
Δ f , pred = ΔY ,meas ( 2.4 + 8.3λ ) = L ( 2.4 + 8.3λ ) (2.2.6)
E

Note that, Δf,pred in Equation 2.2.6 differs from the previous empirical models as it simply

describes the sum of the maximum compressive and tensile deformations rather than the

cumulative compressive and tensile deformations of each cycle. A later publication by Tremblay

et al (2003) suggested a relationship for the maximum rotation prior to brace failure that is more

similar in form to the original model by Lee and Goel (1988) –

−0.1 0.3
⎛b d ⎞ ⎛ KL ⎞
Θ f , pred = 0.091⎜ ⎟ ⎜ ⎟ (2.2.7)
⎝t t ⎠ ⎝ r ⎠

Where b/t and d/t are used generically to represent the brace width-thickness ratio, such that b/t =

B/t – 4 (or d/t = D/t – 4). Note these ratios differ from the B/t – 2 and B/t – 3 used in the previous

empirical relationships and current AISC standards (2005), respectively. Considering the
48

kinematic relationship between the rotation and axial deformation of a buckling brace with plastic

hinges at both ends and at the midpoint, Equation 2.2.7 can also be expressed as –

⎧F 0.3 2 ⎫
⎪ Y ,meas ⎡ ⎛ b d ⎞ ⎛ KL ⎞ ⎤ ⎪
−0.1
Δ f , pred = 2L ⎨ + ⎢0.046 ⎜ ⎟ ⎜ ⎟ ⎥ ⎬ (2.2.8)
⎪⎩ 2 E ⎢⎣ ⎝ t t ⎠ ⎝ r ⎠ ⎥⎦ ⎪

Where Δf,pred is the sum of the maximum tensile and compressive axial deformations.

To show the accuracy of these empirical fracture predictions, Figure 2.12 illustrates the ratio of

experimental fracture measurements, Δf,exp, for three separate testing programs to predicted

fracture capacities, Δf,pred, versus the model number used to calculate the capacity according to the

following key –

1. Lee and Goel (1988) model described by the general Equation 2.2.1 (see Table 2.5 for

calibration constants), where the cumulative experimental deformation, Δf,exp, is

calculated with Equation 2.2.3. Note this model does not consider global slenderness

affects (a4 = 0 in Equation 2.2.1).

2. Archambault et al (1995) model; similar in form to Lee and Goel (1988) model, but

considering affects of brace slenderness (i.e., a4 ≠ 0).

3. Shaback and Brown (2003); Archambault et al (1995) model with different calibration

coefficients (see Table 2.5).

4. Tremblay (2002) model; linear relationship between the summed maximum compressive

and tensile deformations before fracture (Δf) and the global slenderness parameter

(Equation 2.2.6).

5. Tremblay et al (2003) model described by Equation 2.2.8; empirical relationship similar

in form to models 1–3, but with Δf, taken as the sum of the maximum compressive and

tensile deformations before fracture.


49

Referring to Table 2.5 and Figure 2.12, the fracture predictions of model 1 are the farthest from

the experimental fracture deformations (average Δf,exp to Δf,pred ratio of 2.6 with a COV of 0.42),

most likely from the omission of a slenderness dependent quantity. Models 2 and 3 include a

global slenderness term (KL/r), resulting in more accurate fracture capacity predictions (average

Δf,exp/Δf,pred = 1.3 and 1.2, respectively) with model 3 showing less scatter (COV = 0.30) than

model 2 (0.41). Interestingly, model 4 is one of the more accurate models (Δf,exp/Δf,pred = 0.8, COV

= 0.41) despite the absence of a b/t term in the formulation (refer Figure 2.5). The last model (5),

is the most accurate empirical-based model (Δf,exp/Δf,pred = 0.9, COV = 0.37). Interestingly,

whereas the first three models used a cumulative deformation measure to assess fracture capacity,

model 5 relies only on the summed maximum compressive and tensile deformations (or

deformation range). This is an important basis for the work presented in Chapter 6 which relies

on the assumption that brace capacity can be accurately characterized through the maximum

deformation range prior to fracture.

While empirical-based relationships are useful, they tend to be quite sensitive to variations in

loading history (Shaback and Brown, 2003) and brace material property. Moreover, the results in

Table 2.5 and Figure 2.12 are calibrated for only HSS bracing sections, under a limited set of

loading histories. To extend the methodology to other brace cross-sections, more experimental

data would be needed to calibrate accurate relationships. While the form of additional equations

would be similar to the relationships presented above, the confidence of empirical-based models

relies solely on a comprehensive data set with tests encompassing a wide range of b/t and KL/r

ratios and loading histories. Thus, one could argue that there is a need to develop general models

that implicitly reflect the relationships described above while not relying on numerous large-scale

tests data sets to calibrate brace ductility.


50

2.2.2.2 FIBER ELEMENT-BASED CRITICAL STRAIN MODELS

Uriz (2005) recently implemented an “on-the-fly” rain-flow counting scheme to monitor the

accumulated cyclic damage of an HSS6x6x3/8 bracing member for several different loading

histories. The model was used together with a fiber-based brace model (described previously) and

was shown to predict fracture quite accurately. The damage is expressed in terms of an equivalent

cycle, ni, for a cycle with a strain amplitude of, εi, at any point and is compared to a critical strain

measure according to –

ε crit < ε i nim (2.2.9)

Where m and εcrit are calibrated according to standardized fatigue testing procedures. For the case

of monotonic loading, ni = 1 and Equation 2.2.9 simplifies to a comparison between the strain at

any point, εi, and the critical monotonic fracture strain, εcrit. During cyclic loading (i.e., ni > 1) the

strain demand needed to induce fracture at a point is reduced by the assumption that the material

is undergoing a fatigue capacity reduction. Once the critical strain is exceeded, the material is

assumed to have fractured at that point. Referring to the previous discussion on the fiber-based

modeling approach, and considering that stress and strain is defined at a fiber level, a fiber can be

“removed” (strength and stiffness set equal to zero) once the strain reaches the critical strain

according to Equation 2.2.9. This allows the model to mimic the behavior of fracture initiation

through to propagation using a simple fiber-based model. In the context of performance-based

earthquake engineering, this approach is highly attractive as fiber-based models are

computationally cheap compared to more detailed continuum models, yet are more fundamental

in nature as compared to other simplified models (i.e., phenomenological models).

While the work by Uriz (2005) suggests a critical strain of εcrit = 0.095 with m = 0.5 for an

HSS6x6x3/8 bracing member, one may argue that these strains, being monitored at the fiber

level, are not the “true” strains responsible for fracture, and thus may be configuration dependent.

You might also like