Download as pdf or txt
Download as pdf or txt
You are on page 1of 38

Engineering Structures

RC U-shaped walls subjected to in-plane, diagonal, and torsional loading: new


experimental findings
--Manuscript Draft--

Manuscript Number:

Article Type: Research Paper

Keywords: C-shaped; concrete; cores; DIC; displacement-based; torque

Corresponding Author: Ryan Hoult, PhD


University of Melbourne
Lausanne, Vaud AUSTRALIA

First Author: Ryan Hoult, PhD

Order of Authors: Ryan Hoult, PhD

Katrin Beyer, PhD

Manuscript Region of Origin: Asia Pacific

Abstract: Although reinforced concrete U-shaped walls are popular in construction practice
internationally, there is a paucity of experimental research investigating the seismic
performance of such salient elements. The present paper summarizes an experimental
campaign on two slender U-shaped reinforced concrete walls detailed with a single-
layer of reinforcement. State-of-the-art instrumentation was used to capture the three-
dimensional displacement field of the wall surfaces using digital image correlation
techniques. Experimental findings are presented, including strain profiles, equivalent
plastic hinge lengths, longitudinal strains at the base, cracking distributions and widths,
and out-of-plane deformations. The longitudinal strain profiles showed a yielding region
up the boundary ends of the flanges of approximately 800 mm to 1200 mm in length,
depending on the direction of loading and at large drift levels. Approximately half of the
yielding zone length was found to be equal to the equivalent plastic hinge lengths,
which were found to decrease as a function of drift. The longitudinal strains at the base
of these walls showed some shear lag effects when subjected to in-plane or diagonal
loading. For most directions of loading, the largest crack widths were found to be
associated with flexural-shear or shear cracks. When subjected to a pure torque, the
vertical strain distribution at the base of the wall correlated with the theoretical
distribution for an open section governed by warping torsion. The out-of-plane
deformations were primarily concentrated within a small region towards the ends of the
flanges prior to the local buckling failures that were observed experimentally.

Suggested Reviewers: Carlos Arteta, PhD


Assistant Profesor, Universidad del Norte
carteta@uninorte.edu.co
A. Prof Arteta's research focuses on the seismic performance of slender RC U-shaped
walls. The submitted manuscript focuses on experimental results from such structural
specimens.

Farhad Dashti, PhD


Research Associate, University of Canterbury
farhad.dashti@canterbury.ac.nz
Dr Farhad Dashti primary research has focused on reinforced concrete walls, including
recent testing for out-of-plane failures modes.

Opposed Reviewers:

Powered by Editorial Manager® and ProduXion Manager® from Aries Systems Corporation
Cover Letter

Dr. Ryan Hoult


University of Melbourne, Parkville Campus
Infrastructure Engineering Department
Building 174, Block C, Level 3, Room C311
Parkville, VIC, 3052

22/08/2020

Dear Editor,

I wish to submit an original research article entitled “RC U-shaped walls subjected to in-plane, diagonal,
and torsional loading: new experimental findings” for consideration by the Engineering Structures
journal.

I can confirm that this work is original and has not been published elsewhere, nor is it currently under
consideration for publication elsewhere.

This research presents new experimental findings using state-of-the-art digital image correlation (DIC)
techniques to analyse the seismic behevaiour of two reinforced concrete U-shaped walls that were
recently tested. The seismic performance of such walls is largely unknown given the paucity of
experimental evidence. The tests were conducted at the Earthquake Engineering and Structural Dynamics
Laboratory (EESD Lab), École Polytechnique Fédérale de Lausanne (EPFL) in Switzerland. Some of the
highlights of the paper include: (i) comparison of equivalent plastic hinge length expressions to that
derived experimentally, (ii) strain distributions across the wall section for various loading positions,
including torsion, (iii) cracking distribution and cracking width results at large drift levels for all
directions of loading.

I believe that this manuscript is appropriate for publication by the Engineering Structures journal because
there is currently a paucity of research that focuses on the seismic performance of reinforced concrete U-
shaped walls, and particularly for those that are unconfined with a single-layer of reinforcement. The
experimental results and corresponding dataset from this research, which is made publicly available, will
complement the limited database of non-planar and singularly reinforced walls.

The authors have no conflicts of interest to disclose.

Please address all correspondence concerning this manuscript to me at ryan.hoult@unimelb.edu.au.

Thank you for your consideration of this manuscript.

Sincerely,

Dr Ryan Hoult
Highlights

 Longitudinal strain profiles showed a yielding length of approximately 1200 mm


 The equivalent plastic hinge lengths were shown to be approximately half of the yielding
length at large drift levels
 The longitudinal strains at the base of these walls showed some shear lag effects
 The strain distribution when the wall was subjected to torque correlated with the
theoretical distribution for an open section governed by warping torsion
 Out-of-plane deformations were primarily concentrated within a small region towards the
ends of the flanges prior to the local buckling failures that were observed experimentally.
Manuscript File Click here to view linked References

1 RC U-shaped walls subjected to in-plane, diagonal, and


2 torsional loading: new experimental findings
3 Ryan Houlta), Katrin Beyerb)

4 ABSTRACT

5 Although reinforced concrete U-shaped walls are popular in construction practice internationally, there is a
6 paucity of experimental research investigating the seismic performance of such salient elements. The present
7 paper summarizes an experimental campaign on two slender U-shaped reinforced concrete walls detailed with a
8 single-layer of reinforcement. State-of-the-art instrumentation was used to capture the three-dimensional
9 displacement field of the wall surfaces using digital image correlation techniques. Experimental findings are
10 presented, including strain profiles, equivalent plastic hinge lengths, longitudinal strains at the base, cracking
11 distributions and widths, and out-of-plane deformations. The longitudinal strain profiles showed a yielding region
12 up the boundary ends of the flanges of approximately 800 mm to 1200 mm in length, depending on the direction
13 of loading and at large drift levels. Approximately half of the yielding zone length was found to be equal to the
14 equivalent plastic hinge lengths, which were found to decrease as a function of drift. The longitudinal strains at
15 the base of these walls showed some shear lag effects when subjected to in-plane or diagonal loading. For most
16 directions of loading, the largest crack widths were found to be associated with flexural-shear or shear cracks.
17 When subjected to a pure torque, the vertical strain distribution at the base of the wall correlated with the
18 theoretical distribution for an open section governed by warping torsion. The out-of-plane deformations were
19 primarily concentrated within a small region towards the ends of the flanges prior to the local buckling failures
20 that were observed experimentally.

21 1 INTRODUCTION

22 Reinforced concrete (RC) U-shaped walls are commonly used in buildings throughout the world to resist the
23 lateral loads from wind and earthquakes. The U-shaped cross-section is appealing to architects as it can enclose
24 elevators and stairs, thus increasing the net lettable area of the building. The wall with U-shaped cross-section is
25 also ideal for engineers, as it can provide stiffness in the two translational directions of the structure, whereas a
26 rectangular wall, for example, can only provide stiffness in one. Moreover, slender RC walls that take up less
27 space are believed to be more economical and have become more common in many countries (Dashti et al., 2020).
28 While these walls are popular in construction practice, there is a paucity of experimental research that has focused
29 on the seismic performance of RC U-shaped walls. Some experimental research has been conducted more recently
30 on RC U-shaped walls designed to achieve high ductilities (Behrouzi et al., 2020; Beyer et al., 2008; Constantin
31 & Beyer, 2016; Ile & Reynouard, 2005). However, to the knowledge of the author’s, there has been no previous
32 experimental testing focusing on the seismic performance of RC U-shaped walls with inferior detailing. The

a) Corresponding Author, Postdoctoral Researcher, Department of Infrastructure Engineering, University of Melbourne, Parkville
Campus, Victoria, 3052, Australia. ryan.hoult@unimelb.edu.au
b) Associate Professor, Earthquake Engineering and Structural Dynamics Laboratory, School of Architecture, Civil, and
Environmental Engineering, École Polytechnique Fédérale de Lausanne, EPFL ENAC IIC EESD, GC B2 495, Station 18,
Lausanne, CH-1015, Switzerland
1
33 definition of ‘inferior detailing’ here corresponds to a wall designed with a light amount of vertical reinforcement
34 (e.g., a single layer) and no confinement (e.g., absence of traverse hoop bars) in the critical boundary regions of
35 the wall. Inferior detailing in RC walls is common in low-to-moderate seismic regions, such as Australia (Wibowo
36 et al., 2013). This type of detailing also corresponds to the current construction practice in some regions of
37 moderate-to-high seismicity, such as Colombia, due to some loopholes in the current material standards in
38 Colombia, resulting in a proliferation of slender (80 – 100 mm), unconfined RC walls with a single-layer of
39 vertical reinforcement (Blandón & Bonett, 2020; Blandón et al., 2018; Rosso et al., 2018).

40 To investigate the seismic performance of slender RC U-shaped walls with inferior detailing, an experimental
41 program was recently conducted at the Earthquake Engineering and Structural Dynamics Laboratory (EESD
42 Lab), École Polytechnique Fédérale de Lausanne (EPFL), Switzerland, testing two full-scale RC U-shaped wall
43 specimens. The work is part of a larger research program in collaboration with three Colombian universities:
44 Universidad EIA, Universidad del Norte and Universidad de Medellín. The full experimental test setup, loading
45 procedures, material properties, and initial test results for the two specimens is described in some detail in Hoult
46 et al. (2020). For sake of brevity, only a brief overview of the experimental test setup will be provided here, where
47 full details can be found in Hoult et al. (2020). The purpose of this research paper is to report new experimental
48 findings from the tests using the results from state-of-the-art instrumentation.

49 2 SUMMARY OF EXPERIMENTAL TESTS OF SPECIMENS TUE AND TUF

50 The authors investigated the seismic performance and inelastic behaviour of two slender RC U-shaped walls with
51 a single-layer of vertical reinforcement (Hoult et al., 2020). A summary of the test specimens, the loading
52 protocol, the instrumentation used, and the failure modes observed is given in the following sections. The original
53 report by the authors provides initial experimental results for both TUE and TUF using conventional instruments
54 as well as optical triangulation measurements using a grid of infrared light-emitting diodes (LEDs) (Hoult et al.,
55 2020). Instead, for this research paper, the experimental results will largely focus on that derived from digital
56 image correlation measurements.

57 2.1 GEOMETRY AND DETAILING

58 The cross section of the two wall specimens, denoted “TUE” and “TUF”, is shown in Figure 1a. Both TUE and
59 TUF had the same geometry and reinforcement layout but were subjected to different loading protocols. The walls
60 were representative of full-scale models of the lower three storeys of an idealized core that could enclose a small
61 elevator. The effective height (He) of the wall specimens differed depending on the applied loading direction,
62 with a He of 4.65 m and 4.25 m for the wall bending about the major and minor axes respectively.

2
63 Figure 1 Test units TUE and TUF: (a) cross-section and reinforcement layout and (b) elevation view (not to scale; dimensions in mm)

64 2.2 LOADING PROTOCOL

65 TUE was subjected to loading along the principal axes (North-South and East-West) while TUF was subjected to
66 loading in the diagonal directions. Rotation of the collar at the top of the wall was restrained by applying equal
67 displacements with the North-South West (NS-W) actuator and the North-South East (NS-E) actuator. The
68 loading protocol of TUF also contained small cycles in which a twist was applied at the collar in order to
69 investigate the degradation of the torsional stiffness. Neither wall was subjected to an axial load. The loading
70 directions that caused the flange boundary ends to have a large compression zone were limited to approximate
71 yield in this experimental study. The loading positions for both TUE and TUF are shown in Figure 2b.

3
72 Figure 2 (a) cardinal points, sign convention for forces and displacements, wall parts and (b) loading positions

73 The loading history for TUE is described below:

74  0.0% - 0.4% drift: position O-D-C-O-B-A-O, one cycle (increments of 0.1% drift)
75  0.5% - 1.2% drift: position O-D-C*-O-B-A-O, two cycles (increments of 0.1% drift)
76  1.2% - 2.0% drift: position O-D-C*-O-B-A-O, two cycles (increments of 0.2% drift)
*
77 Position C is limited to 0.4% drift

78 The loading history for TUF subjected the wall to diagonal displacements. The drift levels indicated are the square
79 root of the sum of squares (SRSS) of the north-south and east-west drifts. The loading history for TUF is described
80 below:
81  0.0% - 0.4% SRSS drift: position O-G-H-O-F-E-O, one cycle (increments of 0.2% drift)
82  0.4% - 1.0% SRSS drift: position O-G-H*-O-F-E*-O, one cycle (increments of 0.2% drift)
83  1.0% - 3.5% SRSS drift: position O-D-Ca-O-B-A-O, two cycles (increments of 0.5% drift)
84  0.0% SRSS drift: position O, apply a twist to determine the elastic torsional stiffness (applied torque of
85 39 kNm)
86  0.4% and 1.0% SRSS drift: position O, G and H, apply a twist (applied torque of 39 kNm)
87  1.5 and 2.0% SRSS drift: position O, G and H, apply a twist (applied torque of 26 kNm)
88  2.5% SRSS drift: position O and G, apply a twist (applied torque of 13 kNm)
*
89 Position E and H are limited to 0.4% SRSS drift

90 2.3 TESTING INSTRUMENTATION

91 Throughout testing the wall behaviour was monitored with conventional instrumentation (i.e., linear variable
92 differential transducers, or LVDTs, and load cells). The layout of the measurement systems is shown in Figure 3.
93 The LVDTs were used to measure the global horizontal (in-plane) displacements at the collar (top of the wall).
4
94 Furthermore, the three-dimensional displacement field of the surface ends of the flanges (i.e., boundary ends to
95 the South) of each wall was measured by a grid of infrared light-emitting diodes (LEDs). The position of each
96 LED was tracked by a camera consisting of three digital optical sensors. More information regarding the optical
97 triangulation measurements and corresponding results can be found in Hoult et al. (2020).

98
99 Figure 3 (a) location of conventional and optical measurement devices (dimensions in mm)

100 A speckle pattern for digital image correlation (DIC) measurements was applied on three outside surfaces of the
101 wall (East flange, web and West flange) covering a height from the base of approximately 2 m (Figure 3). The
102 speckle pattern was applied by a stencil with a computer-generated random pattern to produce black dots with an
103 approximate diameter of 5 mm. As each wall was either subjected to bidirectional or diagonal loading, 3-
104 dimensional DIC systems were used. Thus, for each unit, 3 sets of two three-dimensional Manta camera systems
105 were used as depicted in Figure 4, each set capturing a different perimeter section of the wall. The Manta cameras

5
106 recorded black and white photographs to decrease file size and simultaneously have an increased capacity for
107 sensor dimensions, resulting in high resolution images. The DIC systems recorded images of the wall segments
108 at a frequency of 0.05 Hz (i.e., every 20 seconds). The images were processed using VIC-3D (Correlated
109 Solutions, 2019) and the displacement field of the perimeter of the wall (i.e., west flange, web, and east flange)
110 was exported in a grid of 20 mm x 20 mm datapoints.

111
112 Figure 4 Local coordinate systems and camera locations for DIC measurements with corresponding camera pair names

113 2.4 FAILURE MODES

114 2.4.1 SPECIMEN TUE

115 The onset of failure of specimen TUE occurred after splitting cracks formed in the East flange end. Following
116 this, the reinforcing bars in the flange end and one of the two concrete vertical layers buckled out-of-plane (Figure
117 5a). This local failure occurred during loading in the A-B direction (loading parallel to the web) when the East
118 flange was in compression at position B. The local out-of-plane buckling occurred during the first stage of the
119 second cycle with a drift of 2.0%. A similar buckling of the vertical reinforcement at the West flange end was
120 observed when pushing towards C after having reached for the first time 2.5% drift for position D (Figure 5b).
121 As testing continued after observing the localised out-of-plane buckling failures, the concrete crushed in the
122 corners of the web-flange regions during the second cycles of 2.5% drift to positions A, B and D, resulting in
123 shear-sliding along a primary crack in the web and ultimately the loss of concrete within this region of the wall.
124 Correspondingly, a drastic drop in shear strength (≈72%) was observed, instant at which the testing of TUE was
125 stopped.

6
126 Figure 5 Local out-of-plane failure of TUE in (a) the east flange during loading to position B at 2.0% drift and (b) the west flange during
127 loading to position C after reaching 2.5% drift at position D

128
129 2.4.2 SPECIMEN TUF

130 Specimen TUF experienced local out-of-plane buckling of the East flange during loading in the G-H direction for
131 the first cycle of 2.5% SRSS drift. The out-of-plane buckling occurred after extensive horizontal cracks opened
132 up in the boundary ends of the East flange when the wall was pushed to position G; on returning to position O,
133 the plastic zone of the East flange noticeably buckled inwards (towards the West) before the boundary end split
134 open, revealing the lumped longitudinal reinforcement, at approximately zero displacement (i.e. when the wall
135 was centred). As no significant reduction in strength had been observed, the test continued, and a similar local
136 out-of-plane buckling of the boundary end of the West flange was observed during the load cycle towards position
137 E, after having reached 2.5% SRSS drift at position F. In this last case, the rebars buckled outwards (i.e., towards
138 the exterior of the unit to the West). Global failure of Specimen TUF ultimately occurred due to compression
139 failure of the wall at the web-flange intersection while pushing the wall to position G at 3.5% SRSS drift.

7
140 Figure 6 Local out-of-plane failure of TUF in (a) the east flange during loading to position H after reaching SRSS 2.5% drift at position G
141 and (b) the west flange during loading to position E after reaching SRSS 2.5% drift at position F

142 3 STRAIN DISTRIBUTIONS

143 This section presents strain profile and distribution results using the DIC data for both specimens TUE and TUF.
144 More specifically, Section 4.1 presents strain profiles up the height of the wall at the boundary ends of the flanges,
145 which are compared to the data derived from the infrared light emitting diodes (LEDs). Section 4.2 uses the strain
146 profiles derived from the DIC data to determine equivalent plastic hinge lengths, which are compared to some
147 expressions available in the literature. Section 4.3 presents the longitudinal strain distributions at the base of the
148 walls for the different in-plane and diagonal loading positions. Section 4.4 provides the longitudinal strain
149 distributions at the base of wall TUF subjected to torsion.

150 3.1 STRAIN PROFILES AT THE FLANGE BOUNDARY ENDS

151 The DIC data was used to determine the vertical strain profile up the height of the wall using a base length of 100
152 mm. A similar base length of 75 mm was used in Constantin and Beyer (2016) to derive the vertical strains at the
153 base of two different RC U-shaped walls which had the same cross-sectional dimensions as those discussed here.
154 Figure 7 presents the vertical strain results from the DIC data for TUE when the flanges ends were in compression
155 at position C for drift levels of 0.1%, 0.2%, 0.3%, and 0.4%, noting that position C was limited to a drift of 0.4%.
156 Figure 8 presents the vertical strain results from the DIC data for TUE when the flange ends were in tension at
157 position D for drift levels of 0.5%, 1.0%, 1.4%, and 2.0%, where EF and WF correspond to the East flange and

8
158 the West flange. It should be noted that the local coordinates (x,y,z) of the wall relative to the two pairs of Manta
159 cameras used for the DIC technique and corresponding to the location of these results are illustrated in Figure 4.
160 Superimposed in these Figure 7 and Figure 8 are the strain results using the infrared LEDs that were glued on the
161 ends of the flanges up the height of the wall. It should be noted that the LEDs were spaced vertically at 100 mm,
162 corresponding to the same vertical base length used to determine the strains from the DIC data. Furthermore, only
163 the outermost vertical array of LEDs (i.e., columns A and F on the surface of West and East flange ends,
164 respectively) were used to drive the strain values here to be comparable to the strains calculated using the DIC
165 data from the surface of the face of flanges. More information on the LED setup can be found in Hoult et al.
166 (2020). The experimental (DIC and LEDs) compression strains for specimen TUE are similar for the four different
167 drift levels in Figure 7. Similarly, when the flanges ends are in tension (Figure 8), the DIC and LED strains are
168 also comparable. The slight variances in obtained strain values from the DIC and LED data are likely to be a
169 result of (i) the two different techniques to obtain the displacement field, and (ii) the two different vertical areas
170 of the wall. It should be noted that the strains corresponding to strain penetration into the foundation at the base
171 of the wall have been neglected in this study.

172 Figure 7 LED and DIC strain measurements from TUE at position C for West flange (blue) and the East flange (magenta) (a) δ = 0.1%, (b)
173 δ = 0.2%, (c) δ = 0.3%, and (d) δ = 0.4%

9
174 Figure 8 LED and DIC strain measurements from TUE at position D for the West flange (blue) and the East flange (magenta) (a) δ = 0.5%
175 (b) δ = 1.0% (c) δ = 1.4% (d) δ = 2.0%

176 Figure 9 and Figure 10 present the strain profiles for the flange ends of specimen TUF using the sources from the
177 DIC technique and LEDs. When the wall is in compression (i.e., diagonal positions E and H), the strain results
178 from the DIC appear to be reasonable in comparison to the LED results. When the flange ends of TUF are in
179 tension (Figure 10), the strain results from the DIC and LEDs show very similar results. A yielding zone of
180 approximately 1200 mm is clearly presented in Figure 10d for specimen TUF at position G and F.

181 Figure 9 LED and DIC strain measurements from TUF for the West Flange at position E (a) δ = 0.2%, (b) δ = 0.4%, and the East Flange at
182 position H (c) δ = 0.2%, (d) δ = 0.4%

10
183 Figure 10 LED and DIC strain measurements from TUF at position F for the West Flange (blue) and position G for the East Flange
184 (magenta) (a) SRSS δ = 0.5% (b) SRSS δ = 1.0% (c) SRSS δ = 1.4% (d) SRSS δ = 2.0%

185 3.2 EQUIVALENT PLASTIC HINGE LENGTHS

186 One of the most widely used methods for calculating the force-displacement capacity of RC members is the plastic
187 hinge analysis (Almeida et al., 2016). Although the inelastic strains (and curvatures) will vary considerably up
188 the height of a cantilevered wall, plastic hinge analysis assumes that the inelastic curvature is uniform over the
189 plastic hinge length (e.g. Fenwick & Dhakal, 2007). The many empirical expressions that are available in the
190 literature for calculating the equivalent plastic hinge length (Lp) have primarily been derived from experimental
191 or numerical investigations on beams, columns, highly confined walls, or bridge piers with a large amount of
192 longitudinal reinforcement (Bohl & Adebar, 2011; Kazaz, 2013; Moehle, 1992; Paulay, 1986; Priestley et al.,
193 2007; Priestley & Park, 1984, 1987; Thomsen & Wallace, 2004; Wallace & Moehle, 1992). For example,
194 Equation 1 is an expression from Priestley et al. (2007) (denoted PCK07 in the corresponding figures)
195 recommended for assessment purposes. It should be noted that for design purposes, the 0.2Lw in Equation 1 is
196 reduced to 0.1Lw. Furthermore, the contribution of plasticity due to the strain penetration, which is typically
197 included in the expression proposed by Priestley et al. (2007), is excluded here in Equation 1.

𝐿𝑝 = 𝑘𝐻𝑒 + 0.2𝐿𝑤 1

198 where k is a constant reflecting the distribution of plasticity dependent on the ratio of the ultimate strength to yield
199 strength of the reinforcing steel (k = 0.2(fu / fy-1) ≤ 0.08) and Lw is the wall length. The values of k, He and Lw used
200 for wall specimens TUE (positions A, B, and D) and TUF (positions F and G) are given in Table 1.

11
201 Constantin (2016) (denoted CON16 in the corresponding figures) derived a plastic hinge length expression
202 (Equation 2) specifically for RC U-shaped walls detailed for high ductilities and calibrated using both
203 experimental and numerical results.

𝐿𝑝 = [α1 𝐻𝑒 + α2 𝐿𝑤 𝜈](1 + α3 𝐴𝐿𝑅) 2

204 where the α parameters can be read from Table 1 (dependent on the direction of the wall) and ν is related to the
205 average shear stress ratio, which can be calculated using Equation 3.
𝜏 3
ν=( )
0.17√𝑓𝑐′

206 where τ is the average shear stress parameter, which can be calculated from a sectional analysis (i.e., moment-
207 curvature analysis) or can be estimated by using a simplified approach, which involves dividing the base shear
208 (Vb) of the wall by the gross cross-sectional area of the wall (Ag) (Krolicki et al., 2011).
209 Table 1 Alpha (α) parameters for determining the Lp of RC U-shaped walls from Constantin (2016) and Hoult et al. (2018a)

Priestley et al. (2007) Constantin (2016) Hoult et al. (2018a)


Position k He (mm) Lw (mm) α1 α2 α3 α1 α2 α3 α4 α5
A-B 0.041 4650 1300 0.055 0.056 2.5 0.1 0.013 13 7.0 0.8
D 0.041 4250 1050 0.059 0.129 3.6 0.5 0.015 3 1.6 0.1
F-G 0.044 4250 1300 0.058 0.058 2.7 0.1 0.013 13 7.0 0.8

210 It should be noted that in Equation 2, Constantin (2016) has taken the length of the flange or web to correspond
211 to Lw depending on the direction of loading. For example, the Ag used to calculate ν in Equation 3 for the wall
212 bending about the major axis (position A and B) is taken as the length of the web (Lweb) multiplied by the thickness
213 of the wall (tw) and neglecting the cross-section from the flanges. Moreover, the average shear stress for bending
214 about the minor axes was calculated using the Ag corresponding to the area of both the flanges (and neglecting
215 the web of the wall), where the ‘shear force in the direction of the flanges [were] equally distributed between the
216 flanges’ (Constantin, 2016). For the wall loaded to position F or G (Figure 2b, diagonal direction), the average
217 shear stress on the section was taken as the mean between the web and the flange and the wall length (Lw) was
218 taken as the web length.

219 More recently, an expression from Hoult et al. (2018a) (denoted HOU18 in the corresponding figures) (Equation
220 4) was derived using the results from numerical analyses on RC U-shaped walls with inferior detailing commonly
221 found in low-to-moderate seismic regions (i.e., unconfined and low amounts of longitudinal reinforcement).

𝐿𝑝 = (α1 𝐿𝑤 − α2 𝐻𝑒 )(1 − α3 ALR)(α4 𝑒 −α5 𝜈 ) 4

12
222 where the α parameters can be read from Table 1 (dependent on the direction of the wall) and ν is related to the
223 average shear stress capacity, which can be calculated using Equation 3. It should be noted that the α parameters
224 for bending about the major axis (positions A and B) have been used for estimating Lp in the diagonal direction
225 from Hoult et al. (2018a), as the web of the wall will typically carry larger shear stress than the flange in this
226 direction (Constantin, 2016).

227 The resulting values of Lp for wall specimens TUE and TUF at various positions are given in Table 2 using the
228 expressions discussed above (and the parameters given in Table 1). It should be noted that the ultimate shear force
229 in each direction was adopted from Hoult et al. (2020), which is needed to calculate the shear parameter (ν) for
230 two of the three expressions used here.
231 Table 2 Values of Lp for specimens TUE and TUF at various flexural positions

Lp (mm)
Position Priestley et al. (2007) Constantin (2016) Hoult et al. (2018a)
A-B 449 256 487
D 383 251 738
F-G 445 247 522
232

233 Using the vertical strain distributions from specimens TUE and TUF derived from the DIC data, curvature
234 distributions can be calculated, and equivalent plastic hinge lengths are determined for different loading positions
235 and drift levels to ultimately compare to the expressions given above. The curvature distribution up the height of
236 a cantilever wall can be used to determine the rotation at the base. As is well known from the first moment area
237 theorem, the rotation is equal to the area under the curvature distribution. The elastic rotation can be obtained
238 using Equation 5:

1 𝐻
𝜃𝑒 = ∫ 𝛷𝑦 (𝑥) 𝑑𝑥 5
2 0
239 where θe is the elastic rotation and H is the height of the wall.

240 The plastic rotation can be determined with Equation 6:


𝐿𝑦
𝜃𝑝 = ∫ |𝛷(𝑥) − 𝛷𝑦 (𝑥)| 𝑑𝑥 6
0

241 where Φ(x) is the curvature found at a height along the cantilever wall of x and Φy(x) is the curvature at yield at
242 a height along the cantilever wall of x.

243 The equivalent plastic hinge length (Lp) can be calculated by assuming that the plastic rotation (θp) is equivalent
244 to an area with height of Lp and width Φp (Equation 7):

13
𝜃𝑝 𝜃𝑝 𝐿𝑦
1
𝐿𝑝 = = = ∫ |𝛷(𝑥) − 𝛷𝑦 (𝑥)| 𝑑𝑥 7
𝛷𝑝 𝛷𝑢 − 𝛷𝑦 𝛷𝑢 − 𝛷𝑦 0
245 where Φu is the curvature at the ultimate limit state.
246 The equivalent plastic hinge lengths (Lp) were derived from the curvature distribution up the height of the wall
247 using the experimental DIC data in Figure 11 for specimens TUE and TUF as a function of drift. It should be
248 emphasised here that the additional strain due to strain penetration into the foundation have not been included in
249 the calculations. This is because only the displacement field from the speckle pattern on the wall was analysed,
250 where the targets attached to the foundation were not included. Furthermore, plastic hinge lengths were not
251 derived for the wall at positions C, E, and H, as drifts were limited to approximate yield in these directions. The
252 equivalent plastic hinge lengths (Lp) derived experimentally (“DIC” in the corresponding legends) for specimen
253 TUE at positions A and B (Figure 11a) appear to reduce as a function of drift, from a length between 1500 - 2000
254 mm at 0.5% drift to a length between 250 – 500 mm at 2.0% drift. This reduction of Lp as a function of drift is
255 consistent with the findings in Beyer (2007). The Lp results from RC U-shaped wall specimens TUA and TUB
256 (Beyer, 2007) at positions A and B have been superimposed in Figure 11a, which show similar resulting Lp values
257 for the given range of drift. It is worth noting that the cross-sectional dimensions of TUA and TUB are similar to
258 the wall specimens TUE and TUF tested here. Similarly, the Lp for TUE at position D (Figure 11b) reduces as a
259 function of drift, where the Lp derived from the East flange and West flange provide similar values. The Lp results
260 from Beyer (2007) for TUA and TUB at position D are superimposed in Figure 11b, also show comparable results
261 in comparison to the Lp for TUE derived from the West flange. For specimen TUF with diagonal loading (Figure
262 11c), the experimental Lp derived at position F rapidly reduces after yield to a more consistent Lp value of
263 approximately 400 mm. The Lp values obtained from Beyer (2007) for TUA and TUB at position F are also
264 consistent with the values derived here. The Lp at position G also decreases rapidly to a more consistent value
265 (Figure 11c), the length of which is calculated to be larger than those calculated for position F. The expressions
266 from Hoult et al. (2018a) (“HOU18”) and Priestley et al. (2007) (“PCK07”) predict reasonable values of Lp in
267 comparison to the experimentally derived Lp for TUE and TUF at the positions considered here and at large drift
268 levels. In comparison, the expressions for Lp from Constantin (2016) (“CON16”) primarily provide more
269 conservative estimates in comparison to the experimental results here, the use of which could be warranted in the
270 design of an RC wall. It should also be emphasised that the expression from Constantin (2016) was derived
271 specifically for walls with detailing to achieve high ductilities, which contrasts the walls examined here with
272 inferior detailing.

14
273 Figure 11 Equivalent plastic hinge lengths derived from experimental (“DIC”) curvature profiles and expressions available in the literature
274 for specimen TUE (a) East-West direction (positions A-B), (b) North direction (position D), and (c) specimen TUF for the diagonal directions
275 (position F and G)

276 3.3 LONGITUDINAL STRAINS AT THE BASE FROM IN-PLANE AND DIAGONAL LOADING

277 In this section, vertical strains at the base of the wall were derived around the perimeters of wall specimens TUE
278 and TUF using the DIC data. One of the primary focuses was to determine if shear lag effects were prominent
279 during testing. A linear strain profile across the wall length at the base is typically assumed in the design of an
280 RC wall using a displacement-based design procedure (Thomsen & Wallace, 2004) and is inherent in moment-
281 curvature estimates using a sectional analysis. This assumption comes from the elementary theory of bending, or
282 the Bernoulli-Euler assumption, that plane sections remain plane after bending. This theory is only approximately
283 and is only appropriate when there is no shear force or when the structural element has infinite shear stiffness
15
284 (Kwan, 1996). For non-rectangular sections, such as the U-shaped sections analysed in this research investigation,
285 the “plane section remains plane” assumption is not necessarily justifiable (Moffatt & Dowling, 1975), and the
286 effects due to shear lag can be substantial (Arabzadeh & Galal, 2018). The shear lag phenomenon is best described
287 as a shear flow that develops where the longitudinal strains in the centre of the web and flanges lag behind the
288 strains at the web-flange intersection of the wall (Kwan, 1996). For example, the two RC U-shaped wall tests
289 from Constantin and Beyer (2016) showed that the vertical strains captured at the base of the wall deviated
290 substantially from the Bernoulli-Euler assumption. The numerical investigation from Hoult (2019) focusing on
291 RC U-shaped walls found prominent shear lag effects in the wall section that was perpendicular to the loading
292 direction. Furthermore, the effects of shear lag were found to be worsened with increasing axial load on the wall
293 (Hoult, 2019).

294 Using the DIC data, the average longitudinal strains across the wall components and using a vertical base length
295 of 100 mm were calculated for specimens TUE and TUF. It is worth noting here again that the base crack,
296 corresponding to the strain penetration into the foundation, was not included in these calculations. Figure 12
297 shows an example of a heatmap of the resulting strain values for specimen TUE at position B for a drift of 1.0%.
298 As one could expect, the tensile (positive) strains accumulate at the open flexural-shear cracks. Similarly, the
299 largest compressions strains appear to be concentrated at diagonal cracks, now closed, that had been open on
300 previous load cycles. The average strains using a base length of 100 mm in Figure 12b produce more reasonable
301 values of the strain in comparison to Figure 12a with a base length of 20 mm. However, the trends of strain
302 distribution remain the same, where diagonal cracking appears to fluctuate the strain values across the wall
303 section. This contrasts with previous RC U-shaped wall tests, where the maximum concrete strains, for example,
304 accumulate at the very base of the wall (Constantin & Beyer, 2016). It is possible that this behaviour is observed
305 due to the absence of axial load, and thus no pre-compression.

16
306 Figure 12 Heatmap of strain values for the web of TUE at position B for a δ of 1.0% (a) base length of 20 mm and (b) base
307 length of 100 mm

308 This fluctuation of strain is further illustrated in Figure 13a, where the average strain distribution across the web
309 at position B for a drift of 1.0% if presented at four different vertical distances up from the base: 0-100 mm, 100-
310 200 mm, 200-300 mm, 300-400 mm. It should be noted that these strains have been averaged horizontally using
311 a length of 100 mm, similar to the vertical base length of 100 mm. If the maxima tensile or compressive strain is
312 found within the 400 mm vertical distance from the base and at each 100 mm increment across the wall section,
313 the strain distribution in Figure 13b is produced. This calculated strain distribution is compared to an idealised
314 strain distribution from sectional analysis (i.e., plane sections remain plane). Thus, for this research, the vertical
315 strain distribution at the base of the wall is defined as the maximum strains obtained within a height of 400 mm
316 from the base of the wall. It should be noted that the sign (i.e., positive or negative) of the average of these strains
317 within a height 400 mm from the base determines whether to take the maximum of the tension (positive) or the
318 minimum of the compression (negative) strains at each point across the wall length.

319

17
320 Figure 13 The web of TUE at position B for δ of 1.0% (a) strain distribution along different vertical distances from the base and (b) resulting
321 strain distribution using the maximum (or minimum) strain values

322 Figure 14 presents the vertical strain profiles for TUE and TUF at the wall base along the perimeter for all of the
323 positions used experimentally and for two different levels of drift. It should be noted that for the directions of the
324 wall where drifts were limited (positions C, E, and H), the first drift cycles of 0.2% and 0.4% are presented here.
325 Furthermore, for the diagonal loading directions (positions E, F, G, and H), the SRSS drifts are indicated. Some
326 nonlinear strain distributions can be observed from Figure 14. For instance, at positions A and B (Figure 14b and
327 Figure 14c, respectively) a nonlinear strain distribution can be observed in the flanges, particularly for the flange
328 in tension. This appears to corroborate the findings in Hoult (2019) that the effects due to shear lag, resulting in
329 the nonlinear distribution of strain, are more prominent in wall sections perpendicular to the loading direction.
330 The strain distribution results for TUF with diagonal loading (positions E, F, F, and H) are similar to those given
331 in Constantin and Beyer (2016) for a U-shaped wall with the same loading positions. However, the shear lag
332 effects and the nonlinear strain distribution from the walls tested in Constantin and Beyer (2016) appear to be
333 more distinguished in comparison to the results for the walls presented here. As no axial load was applied to TUE
334 or TUF during testing, any shear lag effects could be less notable compared to the same walls with axial load, as
335 was found in Hoult (2019). Some previous experimental testing on non-planar walls that have shown shear lag
336 effects to be prominent were tested with axial load ratios within the range of 3% to 15% (Brueggen et al., 2017;
337 Constantin & Beyer, 2016). The axial load subjected to wall specimens has been shown to also increase the shear
338 demand (Beyer et al., 2011; Hannewald, 2013; Hines et al., 2004), making any shear lag effects more prominent
339 for non-planar walls.

18
340 Figure 14 Vertical strains at the base of walls TUE and TUF determined from digital image correlation techniques using a baselength of 100
341 mm above the foundation on the outer perimeter of the wall. Tensile strains (positive) are plotted on the outside of the wall, whereas
342 comrpessive strains (negative) are on the inside.

343 3.4 LONGITUDINAL STRAINS AT THE BASE FROM TORSIONAL LOADING

344 An open wall section, such as a U-shaped wall, will primarily resist an applied torque through warping torsion,
345 unlike a closed section that is governed by circulatory torsion. The resulting idealised strain distribution at the
346 base of a U-shaped wall subjected to a clockwise torque is illustrated in Figure 15a, which is compared to the

19
347 idealised strain distribution in Figure 15b resulting from the wall being pushed towards position C (i.e., bending
348 about the minor axis with web in tension). As argued in Hoult (2020), given the similarities in strain distributions,
349 it is possible to determine the approximate torsional capacity of a RC U-shaped wall from the moment capacity
350 of the section bending about the minor axis with web in tension. However, rather than focusing on capacity here,
351 the torque applied to specimen TUF during testing was strategically low to simply provide information on the
352 decay of torsional stiffness with drift. Some observations of the decay of torsional stiffness of RC U-shaped walls
353 has been reported in previous research by the authors (Hoult et al., 2020; Hoult & Beyer, 2020). The focus of this
354 section is on the longitudinal strain distribution at the base caused by pure torsion and a combination of flexure
355 and torsion. While there has been some experimental research focusing on the strain distribution of RC U-shaped
356 beams with fixed-fixed conditions and subjected to torsion (Chen et al., 2016; Krpan & Collins, 1981; Xu et al.,
357 2018), to the knowledge of the author’s there has been no experimental research focusing on the resulting strain
358 distribution of RC U-shaped cantilever walls subjected to torsion.
359

360 Figure 15 Idealised strain distribution for a U-shaped wall subjected to (a) torsion (b) translational (in-plane) loading about the minor axis
361 with web in tension (i.e., to position C)

362 In the previous section, longitudinal strains at the base of the wall were calculated from the DIC data for different
363 in-plane and diagonal positions and for drift levels. Using the same calculations, the strains at the base of the wall
364 can be derived for wall TUF subjected to a small twist at three different positions: O, G, and H. Figure 16 provides
365 the strain distribution results at the base corresponding to the load steps when the wall experiences the maximum
366 clockwise and anticlockwise torque. For example, at position O and after a SRSS diagonal drift of 2.0% is
367 achieved, a clockwise and anticlockwise torque of 13 kNm was applied to TUF, which is shown in Figure 16a
368 and Figure 16b respectively. Thus, the drifts indicated in the corresponding legends of Figure 16 are the SRSS of
369 the North-South (NS) and East-West (EW) drifts that are achieved by the wall prior to the applied twist. It should
370 be noted that the strain distributions resulting from the torque applied at positions G and H (Figure 16c-f) were
371 derived by subtracting the strains achieved by the wall from the applied diagonal loading prior to the twist.
372 At position O (Figure 16a and b), when the wall is subjected to pure torsion, the resulting strain distribution
373 corresponds to that expected (i.e., Figure 15a), where larger variability occurs when the wall is more damaged

20
374 after achieving greater diagonal drifts. Interestingly, a difference in the expected distribution of strain is observed
375 at position G (Figure 16c and d) in the east flange, which is under net tension due to the diagonal loading (Figure
376 14h), when subjected to an anticlockwise torque that results in that same flange experiencing further tensile
377 strains. Similarly, when an anticlockwise torque is applied at position H (Figure 16f), an unexpected increase in
378 the magnitude of tensile strain can be observed in the sections of the wall already subjected to tension due to the
379 diagonal loading (i.e., intersection of the west flange and web). This type of behaviour was also observed
380 numerically in Hoult and Beyer (2020) for RC U-shaped walls with no applied axial load that were pushed to an
381 in-plane, flexural position (i.e., position A) and subjected to torque. This behaviour is due to the combination of
382 longitudinal stresses that are caused by the flexural loading and zero axial load, which allows the wall section in
383 tension to rotate more freely in comparison to the wall section in compression for the same applied force (Hoult
384 & Beyer, 2020).

21
385 Figure 16 Longitudinal strains at the base of TUF subjected to torsion (a) at position O, clockwise torque (b) position O, counter clockwise
386 torque (b) position G, clockwise torque (c) position G, counter clockwise torque (d) position H, clockwise toque, and (e) position H, counter
387 clockwise torque

388 4 CRACKING WIDTHS AND DISTRIBUTIONS

389 The displacement fields or strain maps that are derived from the DIC measurements have been used in previous
390 research to produce crack patterns from tested RC specimens (Destrebecq et al., 2010; Hoult et al., 2016). More
391 recently, an automated crack detection and crack measurement (ACDM) procedure was developed by Gehri et
392 al. (2020), which uses the principal tensile strain to focus on the cracked area and estimate crack widths. Using
393 the ACDM procedure, crack patterns and crack widths were determined for wall specimens TUE and TUF at the
394 eight different loading positions used experimentally (Figure 2b). The corresponding drift of the walls at these
395 positions was 2.0% for the in-plane (i.e., along the principal axes) loading directions of TUE and SRSS 2.5% for
396 the diagonal loading directions of TUF. At these drift levels, the displacement applied to the walls in the EW
397 direction (i.e., parallel to the web of the walls) are comparable. It should be noted that for the wall directions with
398 a limit imposed on the applied drift (position C, E, and H), the cracking distributions and cracks widths were
399 determined at a drift of 0.4%. It is acknowledged in Gehri et al. (2020) that at high crack widths the precision can
400 be lost locally near the cracks and can result in “incoherent” cracking distributions. For example, at high drift
401 levels imposed on TUE and TUF, a large amount of paint (with the speckle pattern for DIC) had spalled off the
402 walls around the cracks. However, the drifts levels used here were deemed to be appropriate in providing
403 sufficient information regarding the cracking distribution and cracking widths.

404 Figure 17 through to Figure 24 illustrate the cracking distributions using the ACDM procedure (Gehri et al.,
405 (2020) for position A through H, respectively. A combination of flexural and shear cracks are present in the wall
406 at positions A and B (Figure 17 and Figure 18) for TUE and positions F and G (Figure 22 and Figure 23) for TUF,
407 where the steepest cracks appear to be present in the web of these walls. This is in contrast to some previous tests
408 on RC U-shaped walls, where the loading positions of E and H have resulted in the steepest cracks in the flanges

22
409 and the web (Beyer et al., 2008; Constantin & Beyer, 2016). However, the drifts were limited in the directions of
410 E and H for specimen TUF, as with position C for specimen TUE. The observation of steep cracks at positions A
411 and B was first acknowledged in Hoult et al. (2020) using some preliminary DIC data, where the authors suggest
412 that the larger shear span ratio of specimens TUE and TUF, in comparison to previous test units, were partly
413 responsible for the different cracking distributions observed. The large shear crack distribution (and widths)
414 emphasise the increased shear demand that was imposed on these walls, particularly for these directions of
415 loading. In contrast, the cracking distributions observed in rectangular walls with similar shear span ratios are
416 primarily governed by flexural cracks (Dazio et al., 2009; Lu et al., 2017). The largest cracks widths, which are
417 indicated above some of the cracks (in millimetres), at the in-plane positions A and B (Figure 17 and Figure 18)
418 occur in the web of the wall, whereas the diagonal loading at positions F and G (Figure 22 and Figure 23) results
419 in the largest cracks widths occurring in the flanges. Note that the largest crack widths at these positions (A, B,
420 F, and G) appear to be located closer to the middle of the flange or the web, rather than towards the boundary
421 ends of the flanges. These locations of the largest cracks also tend to correspond with the locations of the largest
422 strains in the distributions given in Figure 14. Flexural cracks appear to dominate the web of TUE at position C
423 (Figure 19), with the primary cracks spaced vertically at approximately 220 mm. For position D (Figure 20), the
424 largest crack widths appear to be located at the ends of the flanges of TUE (i.e., extreme tension fibre region),
425 indicating the walls performance being governed by flexural actions for this direction of loading. At positions E
426 and H (Figure 21 and Figure 24), flexural cracks appear in the flanges, whereas the web of the wall at these
427 positions is dominated by a combination of flexure and shear cracks.

428

23
429 Figure 17 Cracking distributions for wall TUE at position A, δ=2.0%, where the largest crack widths are indicated above and next to a
430 crack in millimetres

431
432 Figure 18 Cracking distributions for wall TUE at position B, δ=2.0%, where the largest crack widths are indicated above and next to a crack
433 in millimetres

434
435 Figure 19 Cracking distributions for wall TUE at position C, δ=0.4%, where the largest crack widths are indicated above and next to a
436 crack in millimetres

24
437
438 Figure 20 Cracking distributions for wall TUE at position D, δ=2.0%, where the largest crack widths are indicated above and next to a
439 crack in millimetres

440
441 Figure 21 Cracking distributions for wall TUF at position E, SRSS δ=0.4%, where the largest crack widths are indicated above and next to a
442 crack in millimetres

25
443
444 Figure 22 Cracking distributions for wall TUF at position F, SRSS δ=2.5%, where the largest crack widths are indicated above the cracks in
445 millimetres

446
447 Figure 23 Cracking distributions for wall TUF at position G, SRSS δ=2.5%, where the largest crack widths are indicated above the cracks in
448 millimetres

26
449
450 Figure 24 Cracking distributions for wall TUF at position H, SRSS δ=0.4%, where the largest crack widths are indicated above the cracks in
451 millimetres

452 5 OUT-OF-PLANE DEFORMATION

453 As explained in Section 2.3, three-dimensional DIC data was captured throughout testing of specimens TUE and
454 TUF, which presents the opportunity to calculate the out-of-plane deformations of the wall. In Section 2.4.1 and
455 Section 2.4.2, the local out-of-plane failures of TUE and TUF (respectively) were described and illustrated in
456 Figure 5 and Figure 6. As the walls investigated are non-planar (i.e., non-rectangular), it is worth defining that
457 out-of-plane deformations here represent deformations perpendicular to the planar-sections of the wall (e.g., the
458 displacements perpendicular to the direction of the length of flange). As discussed in Dashti et al. (2020), the
459 recent experimental testing on RC walls or boundary elements, also summarised in Rosso et al. (2016), had the
460 common feature of maximum out-of-plane deformation when the specimens were centred (i.e., at 0% drift). Thus,
461 all of the following out-of-plane deformations and profiles were calculated when the wall was approximately
462 centred (i.e., at position O, Figure 2b) and prior to the local out-of-plane buckling failures observed
463 experimentally.

464 Out-of-plane deformations of the flanges were calculated using the DIC data for specimens TUE and TUF. Of
465 interest were the deformations of the flanges prior to the local out-of-plane buckling failures observed during
466 testing. Figure 25 presents the out-of-plane deformations mapped across the perimeter of the flanges of TUE and
467 TUF, where the colours of red and blue represent the maximum and minimum deformations calculated. It should
468 be noted that positive deformations coincide with displacement towards the lens of the cameras. All of the out-
469 of-plane buckling failures were observed to occur towards the west and for both specimens. Furthermore, as the

27
470 East flange of specimen TUE failed during loading to position B (i.e., flexural loading towards the east), the true
471 out-of-plane deformation is difficult to calculate as the displacement profile due to flexure would need to be
472 estimated and removed. Thus, the out-of-plane failure of the East flange of TUE has not been presented here. The
473 West flange of TUE (Figure 25a) and the East flange of TUF (Figure 25b) appear to have out-of-plane
474 deformations concentrated within a small, longitudinal region towards the ends of the flange. This is in contrast
475 to the West flange of specimen TUF (Figure 25c), where the out-of-plane deformations appear to spread further
476 up the height of the boundary ends from the base. The west flange of TUF also appears to be able to resist higher
477 out-of-plane deformations prior to the rupturing of the flange ends in comparison to the east flange of TUF, which
478 failed first.

479 Figure 25 Out-of-plane deformation measurements (in millimetres) by the DIC system at position O (a) west flange of TUE returning from
480 position D δ=2.5% (b) east flange of TUF returning from position G SRSS δ=2.5% (c) west flange of TUF returning from position F SRSS
481 δ=2.5%

482 These observations are further substantiated by analysing the out-of-plane displacement profile up the wall height
483 at the boundary ends obtained from the infrared LEDs. It should be noted that three LEDs across the thickness of
484 the flange ends were used to compute an average out-of-plane displacement profile originally in Hoult et al.
485 (2020), whereas the LEDs on the outer faces towards the west (i.e., LED grids A and D) were used here to find
486 the maximum out-of-plane deformations prior to failure. Figure 26 present the out-of-plane displacement profiles
487 of the flange ends (towards the West) of TUE and TUF prior to the local failures observed using two different
488 experimental sources (LEDs and DIC). It should be noted that while the approximate shear-span of the wall
489 (dependent on the direction of loading) is greater than four metres, a maximum height of two meters is used here
490 to coincide with the limiting height mapped using the DIC technique. The experimental results in Figure 26a and
491 Figure 26b clearly show large out-of-plane deformations concentrated within a height of approximately 1000 mm
492 from the base of the wall and within the yielding region of the wall. Conversely, the out-of-plane displacement
28
493 profile shape of the boundary end of the west flange of TUF (Figure 26c) is more consistent with a global response
494 of the wall prior to the local buckling failure observed experimentally (Figure 6b).

495 Figure 26 Out-of-plane deformations recorded by the DIC and LEDs prior to local failure for (a) TUE, west flange (b) TUF, east flange, and
496 (c) TUF, west flange (where positive deformations correspond to displacement towards the west).

497 In Section 3.1, a yielding length of approximately 1200 mm can be observed from the corresponding tensile strain
498 profiles of the flange boundary ends for TUE at positions D and TUF at positions F and G. Interestingly, the out-
499 of-plane deformations in Figure 26a and Figure 26b are restricted within this yielding zone length. A buckling
500 length is commonly needed as input into the various expressions that predict the critical tensile strain prior to out-
501 of-plane instability (Chai & Elayer, 1999; Dashti et al., 2020; Paulay & Priestley, 1993). It is common to assume
502 a buckling length equal to the equivalent plastic hinge length. However, as acknowledged in Dashti et al. (2020),
503 some recent walls tested by several researchers (Dashti et al., 2018; Johnson, 2010; Rosso et al., 2016) observed
504 buckling lengths that exceeded the plastic hinge length by more than a factor of two. Given that the equivalent
505 plastic hinge length is typically taken to be approximately half of the yielding length (Bohl & Adebar, 2011;
506 Kazaz, 2013), the buckling length results found here, which corroborate with some of the observations from tests
507 listed above, indicate that the yielding length (or 2Lp) might be a better parameter for the buckling length used in
508 the various expressions for out-of-plane instability, particularly for assessment. A predicted buckling length of Lp
509 might still be prudent for design purposes. The buckling length in Figure 26c, corresponding to the West flange
510 of TUF is an exception to this, where the length suggested in Rosso et al. (2016) of 75% of the unsupported
511 height, which is approximately 3 metres for wall TUF, would be a better estimate in comparison to the profile
512 given in the figure. However, it should be emphasised that the observed out-of-plane failure was restricted to the
513 yielding length of the wall, even for the West flange of TUF, as shown in Figure 6b.

29
514 6 CONCLUSION

515 This research paper presented new findings for the seismic performance of RC U-shaped walls using state-of-the-
516 art instrumentation with the digital image correlation technique. Two slender RC U-shaped walls were tested at
517 the Earthquake Engineering and Structural Dynamics Laboratory (EESD Lab), École Polytechnique Fédérale de
518 Lausanne (EPFL), Switzerland. Specimen TUE was subjected to in-plane loading and specimen TUF was
519 subjected to diagonal loading. Both wall specimens were observed to fail in a similar mode, with local out-of-
520 plane buckling failures first occurring in the flange boundary end regions of the wall followed by global failure
521 corresponding to concrete crushing in the corners of the web-flange region of the wall.

522 A summary of the main research findings presented here are given below:

523  The strain profiles up the height of the wall in the boundary ends of the flanges were found to be typical,
524 with larger strains observed within a yielding region towards the base of the wall at larger drift levels
525 when the wall was performing inelastically.

526  The length of the yielding region was found to be approximately equal to twice the equivalent plastic
527 hinge lengths that were calculated in Section 4.2.

528  The Lp expression from Hoult et al. (2018a), which was derived specifically for RC U-shaped walls with
529 “inferior” detailing, and Priestley et al. (2007) were found to predict reasonable estimates of the plastic
530 hinge lengths of these wall specimens at large drift levels. The expression from Constantin (2016)
531 provided somewhat conservative estimates, which could be warranted in the design of an RC U-shaped
532 wall.

533  Some shear lag effects could be observed, corresponding to nonlinear strain distributions at the base of
534 the wall when subjected to in-plane and diagonal loading. The authors suggest that these effects would be
535 more distinguished had an axial load been subjected to these walls during testing.

536  Under pure torsion (i.e., at position O), the strain distribution at the base of wall TUF corresponded to the
537 theoretical and expected profile for an open section governed by warping torsion. However, the
538 distribution becomes more complex when the wall is pushed to different flexural positions prior to the
539 applied twist, as was observed here with the diagonal loading subjected to TUF. In the event of an
540 earthquake, it is possible for these types of walls to experience both flexural and torsional loading
541 simultaneously. Thus, given the paucity of research in this area, further research on RC U-shaped walls
542 subjected to torsion is recommended.

543  The steepest cracks appeared to be located in the web of the walls at positions A, B, F, and G, which is
544 dissimilar to steep crack observations from some previous test campaigns on RC U-shaped walls. The
30
545 larger shear span ratio of the walls presented here, as well as the limiting drifts towards positions E and
546 H, are likely to be responsible for the different crack distributions observed.

547  The largest crack widths appeared to be located in the flanges for the diagonal loading of TUF at positions
548 F and G, whereas the largest cracks are observed to be in the web for the in-plane loading of TUE at
549 positions A and B. The location of these largest cracks at these positions tended to be located closer to the
550 centre of the sections of the web or flange, depending on loading direction, rather than the extreme tension
551 fibre regions of the wall. An exception to this is the crack distributions and widths observed when pushing
552 the wall to position D.

553  Using the DIC data, the out-of-plane deformations of the flange boundary ends were found to be
554 concentrated within a small region within the yielding zone, confirming the initial hypothesis in Hoult et
555 al. (2020) that these were local, and not global, failures of the wall. While the boundary end of the West
556 flange of specimen TUF appeared to have an out-of-plane displacement profile typical of a global response
557 of the wall, the buckling failure was ultimately restricted to the yielding region of the wall.

558  The out-of-plane local buckling failures observed from these test specimens corroborate with other tests,
559 where the buckling length was found to be approximately equal to the yielding length, or two times the
560 equivalent plastic hinge length. It is suggested that a buckling length of the lesser of the yielding length
561 or 75% of the unsupported height might be a better indicator of the buckling length for RC walls and input
562 for several out-of-plane instability models in the literature, rather than the equivalent plastic hinge length.

563 7 NUMERICAL DATA AND REPRODUCIBILITY

564 The experimental and numerical dataset from the numerical research investigation undertaken in this paper can
565 be downloaded from the publicly accessible platform Zenodo, DOI: 10.5281/zenodo.3994111. The structure of
566 the data folders is described in the report ‘Data_organization_Hoult_and_Beyer_2020.pdf’, also available for
567 download. The following data is provided: (1) DIC processed index files for TUE and TUF (2) MATLAB files
568 for reading the data, and (3) MATLAB files for reproducing graphs and figures. This readily available dataset
569 allows transparency of the data and files that were used to compile the results in this paper. Furthermore, the
570 dataset also allows reproducibility studies to be conducted with some ease. The dataset also allows greater
571 opportunities for sharing and reusing the research data produced from this investigation, which may help future
572 studies focusing on similar topics.

573 8 ACKNOWLEDGEMENTS

574 The authors would like to acknowledge the support of the Swiss Government Excellence Postdoctoral Scholarship
575 for the year 2018/2019 and the funding provided through the Swiss Bilateral Research Programs for the Latin
31
576 American Region through the one-year research merger scheme. The authors would also like to thank our partners
577 from Colombia, including Carlos Blandón Uribe from Universidad EIA, Carlos Arteta and Gustavo Rodriguez
578 from Universidad del Norte, Ricardo Bonett from Universidad de Medellín, Ana B. Acevedo Jaramillo from
579 Universidad EAFIT, and Mario E. Rodriguez from Universidad Nacional Autónoma de México.

580 9 REFERENCES

581 Almeida, J. P., Tarquini, D., & Beyer, K. (2016). Modelling Approaches for Inelastic Behaviour of RC Walls:
582 Multi-level Assessment and Dependability of Results. Archives of Computational Methods in
583 Engineering, 23(1), 69-100.
584 Arabzadeh, H., & Galal, K. (2018). Seismic-Response Analysis of RC C-Shaped Core Walls Subjected to
585 Combined Flexure, Shear, and Torsion. 144(10), 04018165. doi:10.1061/(ASCE)ST.1943-541X.0002181
586 Behrouzi, A., Mock, A., Lehman, D., Lowes, L., & Kuchma, D. (2020). Impact of Bi-Directional Loading on the
587 Seismic Performance of C-Shaped Piers of Core Walls. Submitted to the Journal of Engineering
588 Structures.
589 Beyer, K. (2007). Seismic design of torsionally eccentric buildings with U-shaped RC Walls. (PhD), ROSE
590 School, (ROSE-2008/0X)
591 Beyer, K., Dazio, A., & Priestley, M. J. N. (2008). Quasi-Static Cyclic Tests of Two U-Shaped Reinforced
592 Concrete Walls. Journal of Earthquake Engineering, 12(7), 1023-1053.
593 doi:10.1080/13632460802003272
594 Beyer, K., Dazio, A., & Priestley, M. J. N. (2011). Shear Deformations of Slender Reinforced Concrete Walls
595 under Seismic Loading. ACI Structural Journal, 108(2), 167-177.
596 Blandón, C., & Bonett, R. (2020). Thin slender concrete rectangular walls in moderate seismic regions with a
597 single reinforcement layer. Journal of Building Engineering, 28, 101035. doi:10.1016/j.jobe.2019.101035
598 Blandón, C. A., Arteta, C. A., Bonett, R. L., Carrillo, J., Beyer, K., & Almeida, J. P. (2018). Response of thin
599 lightly-reinforced concrete walls under cyclic loading. Engineering Structures, 176, 175-187.
600 doi:10.1016/j.engstruct.2018.08.089
601 Bohl, A., & Adebar, P. (2011). Plastic Hinge Lengths in High-Rise Concrete Shear Walls. ACI Structural Journal,
602 108(2), 148-157.
603 Brueggen, B. L., French, C. E., & Sritharan, S. (2017). T-Shaped RC Structural Walls Subjected to
604 Multidirectional Loading: Test Results and Design Recommendations. Journal of Structural Engineering,
605 143(7), 04017040.
606 Chai, Y. H., & Elayer, D. T. (1999). Lateral Stability of Reinforced Concrete Columns under Axial Reversed
607 Cyclic Tension and Compression. Structural Journal, 96(5). doi:10.14359/732
608 Chen, S., Diao, B., Guo, Q., Cheng, S., & Ye, Y. (2016). Experiments and calculation of U-shaped thin-walled
609 RC members under pure torsion. Engineering Structures, 106, 1-14.
610 Constantin, R. (2016). Seismic behaviour and analysis of U-shaped RC walls. École polytechnique fédérale de
611 Lausanne, Lausanne, Switzerland.
612 Constantin, R., & Beyer, K. (2016). Behaviour of U-shaped RC walls under quasi-static cyclic diagonal loading.
613 Engineering Structures, 106, 36-52. doi:10.1016/j.engstruct.2015.10.018
614 Correlated Solutions. (2019). VIC-3D (Version 7). Retrieved from http://www.correlatedsolutions.com
615 Dashti, F., Dhakal, R. P., & Pampanin, S. (2018). Evolution of out-of-plane deformation and subsequent
616 instability in rectangular RC walls under in-plane cyclic loading: Experimental observation. 47(15), 2944-
617 2964. doi:10.1002/eqe.3115
618 Dashti, F., Dhakal, R. P., & Pampanin, S. (2020). Design Recommendations to Prevent Global OOP Instability
619 of Rectangular RC Ductile Walls. Submitted to the Bulletin of the New Zealand Society for Earthquake
620 Engineering.

32
621 Dazio, A., Beyer, K., & Bachmann, H. (2009). Quasi-static cyclic tests and plastic hinge analysis of RC structural
622 walls. Engineering Structures, 31(7), 1556-1571. doi:10.1016/j.engstruct.2009.02.018
623 Destrebecq, J. F., Toussaint, E., & Ferrier, E. (2011). Analysis of Cracks and Deformations in a Full Scale
624 Reinforced Concrete Beam Using a Digital Image Correlation Technique. Experimental Mechanics, 51(6),
625 879-890. doi:10.1007/s11340-010-9384-9
626 Fenwick, R., & Dhakal, R. P. (2007). Material strain limits for seismic design of concrete structures. SESOC:
627 Journal of the NZ Structural Engineering Society, 20(1), pp. 14-28.
628 Gehri, N., Mata-Falcón, J., & Kaufmann, W. (2020). Automated crack detection and measurement based on
629 digital image correlation. Construction and Building Materials, 256, 119383.
630 doi:10.1016/j.conbuildmat.2020.119383
631 Hannewald, P. (2013). Seismic Behavior of Poorly Detailed RC Bridge Piers. (PhD thesis), EPFL, Retrieved from
632 http://infoscience.epfl.ch/record/188364
633 Hines, E. M., Restrepo, J. I., & Seible, F. (2004). Force-Displacement Characterization of Well-Confined Bridge
634 Piers. Structural Journal, 101(4). doi:10.14359/13340
635 Hoult, N. A., Dutton, M., Hoag, A., & Take, W. A. (2016). Measuring crack movement in reinforced concrete
636 using digital image correlation: Overview and application to shear slip measurements. Proceedings of the
637 IEEE, 104(8), 1561-1574. doi:10.1109/JPROC.2016.2535157
638 Hoult, R., Appelle, A., Almeida, J., & Beyer, K. (2020). Seismic Performance of Slender RC U-shaped Walls
639 with a Single-Layer of Reinforcement. Submitted to the Journal of Engineering Structures.
640 Hoult, R., Goldsworthy, H. M., & Lumantarna, E. (2018a). Plastic Hinge Length for Lightly Reinforced C-shaped
641 Concrete Walls. Journal of Earthquake Engineering, 1-32. doi:10.1080/13632469.2018.1453419
642 Hoult, R. D. (2019). Shear Lag Effects in Reinforced Concrete C-shaped Walls. Journal of Structural
643 Engineering, 145(3), 04018270. doi:10.1061/(ASCE)ST.1943-541X.0002272
644 Hoult, R. D. (2020). Torque Capacity of Reinforced Concrete U-Shaped Walls. Submitted to the ACI Structural
645 Journal.
646 Hoult, R. D., & Beyer, K. (2020). Forthcoming. Decay of Torsional Stiffness of RC U-shaped Walls When
647 Subjected Simultaneously to In-plane Translational Displacements. Journal of Structural Engineering.
648 Hoult, R. D., Goldsworthy, H. M., & Lumantarna, E. (2018b). Plastic hinge analysis for lightly reinforced and
649 unconfined concrete structural walls. Bulletin of Earthquake Engineering, 16(10), 4825-4860.
650 doi:10.1007/s10518-018-0369-x
651 Ile, N., & Reynouard, J. M. (2005). Behaviour of U-shaped Walls Subjected to Uniaxial and Biaxial Cyclic Lateral
652 Loading. Journal of Earthquake Engineering, 09(01), 67-94.
653 Johnson, B. (2010). Anchorage detailing effects on lateral deformation components of R/C shear walls.
654 (University of Minnesota), University of Minnesota,
655 Kazaz, l. (2013). Analytical Study on Plastic Hinge Length of Structural Walls. Journal of Structural
656 Engineering(11), 1938.
657 Krolicki, J., Maffei, J., & Calvi, G. M. (2011). Shear Strength of Reinforced Concrete Walls Subjected to Cyclic
658 Loading. Journal of Earthquake Engineering, 15(sup1), 30-71. doi:10.1080/13632469.2011.562049
659 Krpan, P., & Collins, M. P. (1981). Testing thin-walled open RC structure in torsion. Journal of the Structural
660 Division, 107(6), 1129-1140.
661 Kwan, A. K. H. (1996). Shear Lag in Shear/Core Walls. Journal of Structural Engineering, 122(9), 1097-1104.
662 doi:10.1061/(ASCE)0733-9445(1996)122:9(1097)
663 Lu, Y., Henry, R. S., Gultom, R., & Ma, Q. T. (2017). Cyclic Testing of Reinforced Concrete Walls with
664 Distributed Minimum Vertical Reinforcement. Journal of Structural Engineering, 143(5), 04016225.
665 doi:10.1061/(ASCE)ST.1943-541X.0001723
666 Moehle, J. P. (1992). Displacement‐ Based Design of RC Structures Subjected to Earthquakes. Earthquake
667 Spectra, 8(3), 403-428. doi:10.1193/1.1585688
668 Moffatt, K., & Dowling, P. J. (1975). Shear lag in steel box girder bridges. The Structural Engineer, 53(10).
669 Paulay, T. (1986). The Design of reinforced concrete ductile shear walls for earthquake resistance: Research
670 Report No. 81-1, Department of Civil Engineering, University of Canterbury, Christchurch, New Zealand.
33
671 Paulay, T., & Priestley, M. J. N. (1993). Stability of Ductile Structural Walls. Structural Journal, 90(4).
672 doi:10.14359/3958
673 Priestley, M. J. N., Calvi, G. M., & Kowalsky, M. J. (2007). Displacement-based seismic design of structures /
674 M. J. N. Priestley, Gian Michele Calvi, Mervyn J. Kowalsky: Pavia : IUSS Press : Fondazione Eucentre,
675 2007.
676 Priestley, M. J. N., & Park, R. (1984). Strength and Ductility of Bridge Substructures. In. National Roads Board,
677 Wellington, New Zealand: RRU Bulletin 71.
678 Priestley, M. J. N., & Park, R. (1987). Strength and Ductility of Concrete Bridge Columns Under Seismic
679 Loading. Structural Journal, 84(1). doi:10.14359/2800
680 Rosso, A., Almeida, J. P., & Beyer, K. (2016). Stability of thin reinforced concrete walls under cyclic loads: state-
681 of-the-art and new experimental findings. Bulletin of Earthquake Engineering, 14(2), 455-484.
682 Rosso, A., Jiménez-Roa, L. A., Almeida, J. P. d., Zuniga, A. P. G., Blandón, C. A., Bonett, R. L., & Beyer, K.
683 (2018). Cyclic tensile-compressive tests on thin concrete boundary elements with a single layer of
684 reinforcement prone to out-of-plane instability. Bulletin of Earthquake Engineering, 16(2), 859-887.
685 Thomsen, J., & Wallace, J. (2004). Displacement-Based Design of Slender Reinforced Concrete Structural
686 Walls—Experimental Verification. Journal of Structural Engineering, 130(4), 618-630.
687 doi:10.1061/(ASCE)0733-9445(2004)130:4(618)
688 Wallace, J. W., & Moehle, J. P. (1992). Ductility and Detailing Requirements of Bearing Wall Buildings. Journal
689 of Structural Engineering, 118(6), 1625-1644.
690 Wibowo, A., Wilson, J., Lam, N. T. K., & Gad, E. F. (2013). Seismic performance of lightly reinforced structural
691 walls for design purposes. Magazine of Concrete Research, 65, 809-828.
692 Wood, S. L. (1989). Minimum tensile reinforcement requirements in walls. ACI Structural Journal, 86(5), 582-
693 591.
694 Xu, J., Chen, S., Guo, Q., Ye, Y., Diao, B., & Mo, Y. L. (2018). Experimental and Analytical Studies of U-Shaped
695 Thin-Walled RC Beams Under Combined Actions of Torsion, Flexure and Shear. International Journal
696 of Concrete Structures and Materials, 12(1), 33. doi:10.1186/s40069-018-0245-8

34
Conflict of Interest

Declaration of interests

☒ The authors declare that they have no known competing financial interests or personal relationships
that could have appeared to influence the work reported in this paper.

☐The authors declare the following financial interests/personal relationships which may be considered
as potential competing interests:

You might also like